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 390 毫秒
1.
2.
Shokhen M  Khazanov N  Albeck A 《Proteins》2011,79(3):975-985
Various mechanisms for the reversible formation of a covalent tetrahedral complex (TC) between papain and peptidyl aldehyde inhibitors were simulated by DFT calculations, applying the quantum mechanical/self consistent reaction field (virtual solvent) [QM/SCRF(VS)] approach. Only one mechanism correlates with the experimental kinetic data. The His–Cys catalytic diad is in an N/SH protonation state in the noncovalent papain–aldehyde Michaelis complex. His159 functions as a general base catalyst, abstracting a proton from the Cys25, whereas the activated thiolate synchronously attacks the inhibitor's carbonyl group. The final product of papain inhibition is the protonated neutral form of the hemithioacetal TC(OH), in agreement with experimental data. The predicted activation barrier g = 5.2 kcal mol?1 is close to the experimental value of 6.9 kcal mol?1. An interpretation of the experimentally observed slow binding effect for peptidyl aldehyde inhibitors is presented. The calculated g is much lower than the rate determining activation barrier of hemithioacetal formation in water, g, in agreement with the concept that the preorganized electrostatic environment in the enzyme active site is the driving force of enzyme catalysis. We have rationalized the origin of the acidic and basic pKa's on the k2/KS versus pH bell‐shaped profile of papain inhibition by peptidyl aldehydes. Proteins 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

3.
The structural and energetic perturbations caused by cavity-creating mutations (Leu-41 → Val and Leu-41 → Ala) in the dimeric 4-α-helical-bundle protein ROP have been characterized by CD spectroscopy and differential scanning calorimetry (DSC). Deconvolution of the CD spectra showed a decrease in α -helicity as a result of the amino acid exchanges that follows qualitatively the overall decrease in conformational stability. Transition enthalpies are sensitive probes of the energetic change associated with point mutations. ΔH0 values at the respective transition temperatures, T1/2 (71.0, 65.3, and 52.9°C at 0.5 mg/ml) decrease from 580 ± 20 to 461 ± 20 kJ/(mol of dimmer) and 335 ± 20 kJ/(mol of dimmer) for wildtype ROP (Steif, C., Weber, P., Hinz, H.-J., Flossdorf, J., Cesareni, G., Kokkinidis, M. Biochemistry 32:3867-3876, 1993), L41V, and L41A, respectively. The conformational stabilities at 25°C expressed by the standard Gibbs energies of denaturation, ΔG, are 71.7, 61.1, and 46.1 kJ/(mol of dimmer). The corresponding transition enthalpies have been obtained from extrapolation using the c(T)and c(T) functions. Their values at 25°C are 176.3, 101.9, and 141.7 kJ/(mol of dimmer) for wild-type ROP, L41V, and L41A, respectively. When the stability perturbation resulting from the cavity creating mutations is referred to the exchange of 1 mol of CH2 group, the average ΔΔG value is ?5.0 ± 1 kJ/(mol of CH2 group). This decrease in conformation stability suggests that dimeric ROP exhibits the same susceptibility to Leu → Yal and Leu → Ala exchanges as small monomeric proteins. Careful determinations of the partial specific heat capacities of wild-type and mutated protein solutions suggest that the mutational effects are predominantly manifested in the native rather than the unfolded state. © 1995 Wiley-Liss, Inc.  相似文献   

4.
Effects of hydrophobic and electrostatic interactions on the self‐assembling process of the ionic‐complementary peptide EMK16‐II are investigated by atomic force microscopy imaging, circular dichroism spectra, light scattering, and chromatography. It is found that the hydrophobicity of the peptide promotes the aggregation in pure water even at a very low concentration, resulting in a much lower critical aggregation concentration than that of another peptide, EAK16‐II. The effect of anions in solution with different valences on electrostatic interactions is also important. Monovalent anions (Cl? and Ac?) with a proper concentration can facilitate the formation of peptide fibrils, with Cl? of smaller size being more effective than Ac? of larger size. However, only small amounts of fibrils, but plenty of large amorphous aggregates, are found when the peptide solution is incubated with multivalent anions, such as SO, C6H5O, and HPO. More importantly, by gel filtration chromatography, the citrate anion, which induces a similar effect on the self‐assembling process of EMK16‐II as that of SO and HPO, can interact with two or more positively charged residues of the peptide and reside in the amorphous aggregates. This implies a “salt bridge” effect of multivalent anions on the peptide self‐assembling process, which can interpret a previous puzzle why divalent cations inhibit the formation of ordered nanofibrils of the ionic‐complementary peptides. Thus, our results clarify the important effects of hydrophobic and electrostatic interactions on the self‐assembling process of the ionic‐complementary peptides. These are greatly helpful for us to understand the mechanism of peptides' self‐assembling process and protein folding and aggregation. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 318–329, 2010. This article was originally published online as an acceptedpreprint. The “Published Online” date corresponds to the preprintversion. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

5.
A complex of carbonic anhydrase (CA) with one of its substrates, bicarbonate, has been studied crystallographically. Human isoenzyme II was mutated at position 200 from threonine to histidine, which results in higher affinity for bicarbonate. The HCO ion binds in the active site to the zinc ion as a pseudo-bidentate ligand which gives the metal a coordination geometry between tetrahedral and trigonal bipyramide. The water/hydroxide normally bound with tetrahedral coordination to the zinc is probably replaced by the OH group of the bicarbonate ion. The importance of residues Thr-199 and Glu-106 in controlling the binding orientation of HCO is discussed as well as the catalytic mechanism. Both the complex as well as the uncomplexed mutant were studied at 1.9 Å resolution. © 1993 Wiley-Liss, Inc.  相似文献   

6.
A cellular suspension from rat submandibular glands was exposed to different concentrations of NH4Cl, and the variations of the intracellular concentration of calcium ([Ca2+]i) and the intracellular pH (pHi) were measured using fura-2 and 2′,7′-bis-(2-carboxy-ethyl)-5(6)-carboxyfluorescein. More than 5 mmol/l NH4Cl significantly increased the [Ca2+]i without affecting the response to 100 µmol/l carbachol. When exposed to 1 and 5 mmol/l NH4Cl, the cells acidified immediately. At 30 mmol/l, NH4Cl first alkalinized the cells and the pHi subsequently dropped. This drop reflects the uptake of NH ions that dissociate to NH3 and H+ in the cytosol. These protons are exchanged for extracellular sodium by the Na+/H+ exchanger because the presence of an inhibitor of the exchanger in the medium increased the acidification induced by 1 mmol/l NH4Cl. Ouabain partly blocked the uptake of NH. In the combined presence of ouabain and bumetanide (an inhibitor of the Na+-K+-2Cl cotransporter), 1 mmol/l NH4Cl alkalinized the cells. The contribution of the Na/K ATPase and the Na+-K+-2Cl cotransporter in the uptake of NH was independent of the presence of calcium in the medium. Isoproterenol increased the uptake of NH by the cotransporter. Conversely, 1 mmol/l extracellular ATP blocked the basal uptake of NH by the cotransporter. This inhibition was reversed by extracellular magnesium or Coomassie Blue. It was mimicked by benzoyl-ATP but not by CTP, GTP, UTP, ADP, or ADPβS. ATP only slightly inhibited the increase of cyclic AMP (−22%) by isoproterenol but fully blocked the stimulation of the cotransporter by the β-adrenergic agonist. ATP increased the release of 3H-arachidonic acid from prelabeled cells but SK&F 96365, an imidazole-based cytochrome P450 inhibitor, did not affect the inhibition by ATP. It is concluded that the activation of a purinoceptor inhibits the basal and the cyclic AMP-stimulated activity of the Na+-K+-2Cl cotransporter. J. Cell. Physiol. 180:422–430, 1999. © 1999 Wiley-Liss, Inc.  相似文献   

7.
Ca2+ signaling pathways are well studied in cardiac myocytes, but not in cardiac fibroblasts. The aim of the present study is to characterize Ca2+ signaling pathways in cultured human cardiac fibroblasts using confocal scanning microscope and RT‐PCR techniques. It was found that spontaneous intracellular Ca2+ (Ca) oscillations were present in about 29% of human cardiac fibroblasts, and the number of cells with Ca oscillations was increased to 57.3% by application of 3% fetal bovine serum. Ca oscillations were dependent on Ca2+ entry. Ca oscillations were abolished by the store‐operated Ca2+ (SOC) entry channel blocker La3+, the phospholipase C inhibitor U‐73122, and the inositol trisphosphate receptors (IP3Rs) inhibitor 2‐aminoethoxydiphenyl borate, but not by ryanodine. The IP3R agonist thimerosal enhanced Ca oscillations. Inhibition of plasma membrane Ca2+ pump (PMCA) and Na+–Ca2+ exchanger (NCX) also suppressed Ca oscillations. In addition, the frequency of Ca oscillations was reduced by nifedipine, and increased by Bay K8644 in cells with spontaneous Ca2+ oscillations. RT‐PCR revealed that mRNAs for IP3R1‐3, SERCA1‐3, CaV1.2, NCX3, PMCA1,3,4, TRPC1,3,4,6, STIM1, and Orai1‐3, were readily detectable, but not RyRs. Our results demonstrate for the first time that spontaneous Ca oscillations are present in cultured human cardiac fibroblasts and are regulated by multiple Ca2+ pathways, which are not identical to those of the well‐studied contractile cardiomyocytes. This study provides a base for future investigations into how Ca2+ signals regulate biological activity in human cardiac fibroblasts and cardiac remodeling under pathological conditions. J. Cell. Physiol. 223: 68–75, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

8.
Small-angle x-ray scattering of poly(γ-methyl-L -glutamate), [Glu(OMe)]n, in m-cresol and in pyridine was measured to determine the mass per unit length, Mq, and the radius of gyration of the cross section, 〈S1/2. It was confirmed from the values of Mq that [Glu(OMe)]n exists in an α-helical conformation in these solvents. It was elucidated from the calculations on 〈S1/2 that the side chains come in moderately close contact with the main chain in these solvents. It was indicated from the analysis of the outer portion of the scattering curves that the side-chain conformation varied depending on the solvent.  相似文献   

9.
The effects of altered external sodium and potassium concentrations on steady state, active Na+ + K+ transport in Ehrlich ascites tumor cells have been investigated. Membrane permeability to Na+ and K+, intracellular [Na+] and [K+], and membrane potential were measured. Active cation fluxes were calculated as equal and membrane potential were measured. Active cation fluxes were calculated as equal and opposite to the net, diffusional leak fluxes. Elevation of external K+ (6–60 Mm)by equivalent replacement of Na+ (154–91 mM) inhibits both active Na+ and K+ fluxes, but not proportionally. This results in a decrease of the coupling ratio (rp = -Jkp/J) as external K+ is increased. Elevation of external K+ (3–68 mM) at constant Na+ (92mM) inbibits J, but is without effect on J. The coupling ratio declines from 1.01 ± 0.14 to 0.07 ± 0.05, a 14-fold alteration. Reduction of external Na+ (154–25 mM) at constant K+ (6mM) depresses J, but is without effect on J. The coupling ratio increases from 0.63 ± 0.04 at 154 mM Na+ to 4.5 ± 2.04 at 25 mM Na+. The results of this investigation are consistent with the independent regulation of active cation fluxes by the transported species. Kinetic analysis of the data indicates that elevation of external sodium stimulates active sodium efflux by interacting at “modifier sites” at the outer cell surface. Similarly, external potassium inhibits active potassium influx by interaction at separate modifier sites.  相似文献   

10.
The scavenging effects of eighteen thiazolyl thiazolidine‐2,4‐dione compounds (TTCs) on superoxide radical , hydroxyl radical HO?, and 1,1‐diphenyl‐2‐picrylhydrazyl (DPPH?) radical were evaluated by the chemiluminescence technique, electron spin resonance spectrometry (ESR) and visible spectrophotometry, respectively. The examined compounds were shown to have 27–59% scavenging ability, 19–69% HO? scavenging activity and 2–32% DPPH? scavenging ability. This property of the tested compound seems to be important in the prevention of various diseases of free radicals etiology. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

11.
The effect of the Y108V mutation of human glutathione S‐transferase P1‐1 (hGST P1‐1) on the binding of the diuretic drug ethacrynic acid (EA) and its glutathione conjugate (EASG) was investigated by calorimetric, spectrofluorimetric, and crystallographic studies. The mutation Tyr 108 → Val resulted in a 3D‐structure very similar to the wild type (wt) enzyme, where both the hydrophobic ligand binding site (H‐site) and glutathione binding site (G‐site) are unchanged except for the mutation itself. However, due to a slight increase in the hydrophobicity of the H‐site, as a consequence of the mutation, an increase in the entropy was observed. The Y108V mutation does not affect the affinity of EASG for the enzyme, which has a higher affinity (Kd ~ 0.5 μM) when compared with those of the parent compounds, K ~ 13 μM, K ~ 25 μM. The EA moiety of the conjugate binds in the H‐site of Y108V mutant in a fashion completely different to those observed in the crystal structures of the EA or EASG wt complex structures. We further demonstrate that the ΔCp values of binding can also be correlated with the potential stacking interactions between ligand and residues located in the binding sites as predicted from crystal structures. Moreover, the mutation does not significantly affect the global stability of the enzyme. Our results demonstrate that calorimetric measurements maybe useful in determining the preference of binding (the binding mode) for a drug to a specific site of the enzyme, even in the absence of structural information.  相似文献   

12.
AMBER v. 4.1 force field in 1.5 ns NPT molecular dynamics simulations of murine epidermal growth factor (mEGF), human epidermal growth factor (hEGF), and human transforming growth factor-α (hTGF-α) structures with explicit TIP3P solvation were used to investigate differences in backbone stability, changes in secondary structure, interdomain flexibility, and weakly polar interactions. Backbone root mean square deviations of sections of each peptide show that the most stable regions in mEGF and hEGF are the A-, B-, and C-loops, whereas the most stable regions in hTGF-α are the A- and B-loops. The secondary structure in the B-loops of mEGF and hEGF differ significantly from the nuclear magnetic resonance (NMR) structures of mEGF and hEGF. The position and type of turns in the B-loop of mEGF and hEGF increase the interstrand distance of the antiparallel β-sheets thereby disrupting their structure. The interdomain flexibility of simulated hTGF-α structure is greater than in either mEGF or hEGF. The ϕ, ψ dihedrals of hTGF-α occupy two distinct populations of phase space corresponding to either a C or an α-helical conformation. This change in dihedral angle is stabilized by Phe15 with Arg42 and Phe17 with Arg42 N-π weakly polar interactions that are present only in hTGF-α but not in mEGF or hEGF. Proteins 33:396–407, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

13.
The antioxidant behavior of a series of new synthesized substituted thiazolyl‐thiazolidine‐2,4‐dione compounds (TZDs) was examined using chemiluminescence and electron paramagnetic resonance spin trapping techniques. 5,5‐Dimethyl‐1‐pyrroline‐N‐oxide (DMPO) was used as the spin trap. The reactivity of TZDs with superoxide anion radical (O) and hydroxyl radical (HO?) was evaluated using potassium superoxide/18‐crown‐6 ether dissolved in dimethylsulfoxide, and the Fenton‐like reaction (Fe2+ + H2O2), respectively. The results showed that TZDs efficiently inhibited light emission from the O generating system at a concentration of 0.05–1 mmol L?1 (5–94% reductions were found at 1 mmol L?1 concentration). The TZD compounds showed inhibition of HO?‐dependent DMPO–OH spin adduct formation from DMPO (the amplitude decrease ranged from 8 to 82% at 1 mmol L?1 concentration). The findings showed that examined TZDs had effective activities as radical scavengers. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

14.
It was found that the cellular Na+-concentration (C) of Lodderomyces elongisporus D is depended on the extracellular K+-concentration (C). The relationship can be described by an equation in the form The function of the natrium ion seem to be to support the utilisation rate of potassium ion at lower extracellular K+-concentration.  相似文献   

15.
Osteoblasts synthesize bone in polarized groups of cells sealed by tight junctions. Large amounts of acid are produced as bone mineral is precipitated. We addressed the mechanism by which cells manage this acid load by measuring intracellular pH (pHi) in non‐transformed osteoblasts in response to weak acid or bicarbonate loading. Basal pHi in mineralizing osteoblasts was ~7.3 and decreased by ~1.4 units upon replacing extracellular Na+ with N‐methyl‐D ‐glucamine. Loading with 40 mM acetic or propionic acids, in normal extracellular Na+, caused only mild cytosolic acidification. In contrast, in Na+‐free solutions, weak acids reduced pHi dramatically. After Na+ reintroduction, pHi recovered rapidly, in keeping with Na+/H+ exchanger (NHE) activity. Sodium‐dependent pHi recovery from weak acid loading was inhibited by amiloride with the Ki consistent with NHEs. NHE1 and NHE6 were expressed strongly, and expression was upregulated highly, by mineralization, in human osteoblasts. Antibody labeling of mouse bone showed NHE1 on basolateral surfaces of all osteoblasts. NHE6 occurred on basolateral surfaces of osteoblasts mainly in areas of mineralization. Conversely, elevated HCO alkalinized osteoblasts, and pH recovered in medium containing Cl?, with or without Na+, in keeping with Na+‐independent Cl?/HCO exchange. The exchanger AE2 also occurred on the basolateral surface of osteoblasts, consistent with Cl?/HCO exchange for elimination of metabolic carbonate. Overexpression of NHE6 or knockdown of NHE1 in MG63 human osteosarcoma cells confirmed roles of NHE1 and NHE6 in maintaining pHi. We conclude that in mineralizing osteoblasts, slightly basic basal pHi is maintained, and external acid load is dissipated, by high‐capacity Na+/H+ exchange via NHE1 and NHE6. J. Cell. Physiol. 226: 1702–1712, 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

16.
Excited‐state intramolecular proton transfer (ESIPT) and dual luminescence behaviour of 3‐hydroxyflavone (3‐HF) have been utilized to monitor its binding to liposomal membranes prepared from egg yolk phosphatydilcholine (EYPC). Additionally, absorption spectrophotometric assay has been performed to evaluate the antioxidant activity of 3‐HF against lipid peroxidation in this membrane system. When 3‐HF molecules are partitioned into EYPC liposomes, a weak long‐wavelength absorption band with λ ~410 nm appears in addition to the principal absorption at ~λ = 345 nm. Selective excitation of the 410 nm band produces the characteristic emission (λ~460 nm) of the ground‐state anionic species, whereas excitation at the higher energy absorption band leads to dual emission with predominatly ESIPT tautomer fluorescence (λ = 528 nm). Both ESIPT tautomer and the anionic species exhibit fairly high fluorescence anisotropy (r) values (r = 0.122 and 0.180, respectively). Biexponential fluorescence decay kinetics are observed for the ESIPT tautomer as well as the ground‐state anionic forms, indicating heterogeneity in the microenvironments of the corresponding emitting species. Furthermore, we demonstrate that lipid peroxidation of EYPC liposomes is significantly inhibited upon 3‐HF binding, suggesting that 3‐HF can be potentially useful as an inhibitor of peroxidative damage of cell membranes. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

17.
Complex formation between poly(U) and adenosine in solutions of salts that stabilize (Na2SO4), destabilize (NaClO4), or have little effect on the water structure (NaCl), as well as the poly(U)·poly(A) interaction in NaClO4, was studied by equilibrium dialysis and uv spectroscopy. At 3°C and neutral pH, Ado·2 poly(U) is formed in 1M NaCl and 0.33M Na2SO4. In NaClO4 solutions under the same conditions, an Ado·poly(U) was found over the whole range of salt concentration investigated (10 mM?1M), which has not been previously observed under any conditions. The Ado-poly(U) was also found in a NaCl/NaClO4 mixture, the transition from the triple- to the double-helical complex occurring within a narrow range of concentration of added NaClO4. In the presence of 1M NaCl this transition is observed on adding as little as 10 mM NaClO4, i.e., at a [ClO]/[Cl?] ratio of about 1:100. However, when NaClO4 is added to a 1M solution of the stabilizing salt Na2SO4, no transition occurs even at a [ClO]/[SO] ratio of 1:4. Investigation of melting curves and uv spectra has shown that in an equimolar mixture of the polynucleotides, only a double-helical poly(U)·poly(A) exists in 1M NaClO4 at low temperatures; this also holds for 1M NaCl. This changes to a triple-helical 2 poly(U)·poly(A) and then dissociates as the temperature increases. At low temperatures and the poly(U)/poly(A) concentration ratio of 2:1, a mixture of 2 poly(U)·poly(A) and poly(U)·poly(A) was observed in 1M NaClO4, in contrast to the case of 1M NaCl. Thus, sodium perchlorate, a strong destabilizer of water structure, promotes formation of double-helical complexes both in the polynucleotide–monomer and the polynucleotide–polynucleotide systems. Beginning with a sufficiently high ionic strength (μ ? 0.9), a further increase in the salt molarity results in an increase of the poly(U)·adenosine melting temperature in both stabilizing and neutral salts and a decrease in the destabilizing salt. In Na2SO4 concentrations higher than 1.2M Ado·2 poly(U) precipitates at room temperature. Analysis of the binding isotherms and melting profiles of the complexes between poly(U) and adenosine according to Hill's model shows that the cooperativity of binding, due to adenosine stacking on poly(U), increases in the order NaClO4 < NaCl < Na2SO4. The free energy of adenosine stacking on the template is similar to that of hydrogen bonding between adenosine and poly(U) and ranges from ?1 to ?2 kcal/mol. The values of ΔHt [the effective enthalpy of adenosine binding to poly(U) next to an occupied site, obtained from the relationship between complex melting temperature and free monomer concentration at the midpoint of the transition] are ?14.2, ?18.3, and ?16.8 kcal/mol for 1M solutions of NaClO4, NaCl, and Na2SO4, respectively. The results indicate that the effects of anions of the salts studied are related to water structure alterations rather than to their direct interaction with the complexes between poly(U) and adenosine.  相似文献   

18.
The oxygen free radical scavenging activities of 15 chromonyl‐thiazolidine‐2,4‐dione compounds (CTDs) were examined in chemical systems producing superoxide anion radicals, O (potasium superoxide–18‐crown‐6 ether–DMSO), and hydroxyl radicals, HO? (a Fenton reaction: Fe(II)–H2O2–sodium trifluoroacetate, pH 6.15). Chemiluminescence and electron spin resonance (ESR) spectroscopy using 5,5‐dimethyl‐1‐pyrroline‐1‐oxide (DMPO) as spin trap were applied to evaluate antioxidant behaviour of CTDs towards the oxygen radicals. The results indicated that 11 of the 15 tested compounds showed a significant inhibitory effect on the chemiluminescence generated from the O‐generating system, ranging from 41 to 86%, and 13 CTDs quenched the ESR signal of the DMPO–OH spin adduct by 33–86%, at a concentration of 1 mmol L?1. Our findings demonstrate that CTDs could be good free radical scavengers. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

19.
Skeletal abnormalities in neurofibromatosis type 1 syndrome (NF1) are observed in ~50% of patients. Here, we describe the phenotype of Nf1Ocl mouse model with Nf1‐deficient osteoclasts. Nf1Ocl mice with Nf1+/? or Nf1?/? osteoclasts in otherwise Nf1+/+ background were successfully generated by mating parental Nf1flox/flox and TRAP‐Cre mice. Contrary to our original hypothesis, osteoporotic or fragile bone phenotype was not observed. The µCT analysis revealed that tibial bone marrow cavity, trabecular tissue volume, and the perimeter of cortical bone were smaller in Nf1 mice compared to Nf1 control mice. Nf1 mice also a displayed narrowed growth plate in the proximal tibia. In vitro analysis showed increased bone resorption capacity and cytoskeletal changes including irregular cell shape and abnormal actin ring formation in Nf1?/? osteoclasts. Surprisingly, the size of spleen in Nf1 mice was two times larger than in controls and histomorphometric analysis showed splenic megakaryocytosis. In summary, Nf1Ocl mouse model presented with a mild but specific bone phenotype. This study shows that NF1‐deficiency in osteoclasts may have a role in the development of NF1‐related skeletal abnormalities, but Nf1‐deficiency in osteoclasts in Nf1+/+ background is not sufficient to induce skeletal abnormalities analogous to those observed in patients with NF1. J. Cell. Biochem. 113: 2136–2146, 2012. © 2012 Wiley Periodicals, Inc.  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