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.
Abstract

A novel screening assay for the identification of baculovirus infected cells expressing membrane receptors was developed by using a replica transfer technique. Sf9 cells were cotransfected with wild type baculoviral DNA and the transfer vector pVL941–β1 containing the coding region of the human β1-adrenergic receptor gene. Infected cells embedded in agarose were incubated with [125I]-iodocyanopindolol and transferred onto filters that were subsequently autoradiographed. This procedure resulted in the isolation of recombinant baculoviruses that expressed β1-adrenergic receptors. Binding assays carried out with [125I]-ICYP indicated that more than 600,000 receptors were expressed per cell, the highest level noted so far for this receptor in genetically engineered cells. Sf9 cells expressing the β1-AR were analysed by ligand binding, competition experiments, adenylyl cyclase stimulation and photoaffinity labeling. These cells express a homogenous population of receptors and display the known pharmacological properties of β1-AR in human tissues.  相似文献   

2.
Enteropeptidase (EC 3.4.21.9) is the glycoprotein enzyme in the small intestine that triggers the activation of the zymogens in pancreatic juice by converting trypsinogen into trypsin. Because of its physiological significance, there have been many studies on the expression, purification, and characterization of enteropeptidase from different species. The baculovirus expression system has been commonly used in research communities and scientific industries for the production of high levels of recombinant proteins, which require posttranslational modifications for functional activity. In the present study, we isolated bovine enteropeptidase catalytic subunit gene from Bos taurus indicus (GenBank accession no. KC756844), and cloned it in pFast Bac HT “A” baculovirus expression donor vector, under the polyhedrin promoter. Recombinant bovine enteropeptidase was expressed in SF-9 insect cells with high expression levels. Recombinant enteropeptidase was purified using Ni-NTA affinity chromatography. A 6-mg quantity of pure active protein was obtained from 100 mL culture using this approach. Its activity and kinetic parameters were determined by cleavage of its fluorogenic substrate Gly-(Asp) 4-Lys-β-naphthylamide. The recombinant bovine enteropeptidase showed a K m value of 0.75 ± 0.02 mM with K cat 25 ± 1 s.  相似文献   

3.
-Synuclein, a presynaptic protein of the central nervous system, has been implicated in the synaptic events such as neuronal plasticity during development and learning, and neuronal degeneration under pathological conditions. As an effort to understand the biological function of -synuclein, we examined the expression patterns of -synuclein in various human hematopoietic cells, and in Drosophila at different developmental stages. The -synuclein was ubiquitously expressed in all the tested hematopoietic cells including T cells, B cells, NK cells, and monocytes, as well as in the lymphoma cell lines, Jurkat and K562. A potential -synuclein homologue was also expressed in Drosophila, and its expression appeared to be temporally and spatially regulated during development. Our data suggest that -synuclein may function in invertebrates as well as in vertebrates and its function may not be restricted to the neuron.  相似文献   

4.
Li X  Pei J  Wu G  Shao W 《Biotechnology letters》2005,27(18):1369-1373
For the first time, a β-glucosidase gene from the edible straw mushroom, Volvariella volvacea V1-1, has been over-expressed in E. coli. The gene product was purified by chromatography showing a single band on SDS-PAGE. The recombinant enzyme had a molecular mass of 380 kDa with subunits of 97 kDa. The maximum activity was at pH 6.4 and 50 °C over a 5 min assay. The purified enzyme was stable from pH 5.6–8.0, had a half life of 1 h at 45 °C. The β-glucosidase had a Km of 0.2 mM for p-nitrophenyl-β-D-glucopyranoside.  相似文献   

5.
The glycoside hydrolase β-1,3-glucomannanase is an enzyme that specifically breaks the β-1,3 glycosidic bond of the glucomannan, the main cell wall constituent of some yeasts. In this work, a codon optimized DNA sequence of the MAN5C gene from Penicillium lilacinum ATCC 36010 was expressed in the yeast Pichia pastoris under the control of AOX1 promoter. The recombinant protein plMAN5C was purified from the shake flask culture and the stirred-tank bioreactor culture in yields of 30.0 mg/l and 224.0 mg/l, respectively. The purified protein had a specific activity of 14.6 U/mg at 37 °C, pH 4.5. Biochemical analysis showed that the optimal temperature and pH for plMAN5C were 50 °C and 4.5, respectively. The recombinant plMAN5C was efficient in lysis of the cell wall of the red yeast Rhodosporidium toruloides to form protoplast. Our work provided an effective system for heterogeneous production of β-1,3-glucomannanase, which should facilitate a more convenient application of this enzyme in biotechnology and other related areas.  相似文献   

6.
α-N-Acetylgalactosaminidase (αNAGAL, EC 3.2.1.49) purified from chicken liver has been used in seroconversion of human erythrocytes. Blood group A, defined by the terminal α-linkedN-acetylgalactosamine, can be cleavedin vitroby αNAGAL, resulting in the underlying penultimate blood group H (O) epitope structure. In order to produce sufficient quantities of recombinant αNAGAL (rαNAGAL) for such studies, we expressed the cDNA encoding chicken liver αNAGAL inPichia pastoris,a methylotrophic yeast strain. The αNAGAL coding sequence was cloned into theEcoRI site of the vector pPIC 9 such that the protein was in the same reading frame as the secretion signal of yeast α-mating factor derived from the vector. AfterP. pastoristransformation, colonies were screened for high-level expression of rαNAGAL based on enzyme activity. As a result of methanol induction of high-density cell cultures in a fermentor, enzymatically active rαNAGAL was produced and secreted into the culture medium. The recombinant enzyme was purified over 150-fold by chromatography on a cation exchange column followed by an affinity column. Its homogeneity was confirmed by Coomassie blue-stained SDS–PAGE, Western blot, and N-terminal sequencing. The purified rαNAGAL has a molecular mass of approximately 50 kDa while its native counterpart has a molecular mass of 43 kDa. This discrepancy in size was eliminated by endoglycosidase treatment, suggesting that the recombinant protein was hyperglycosylated by the hostP. pastoriscells. rαNAGAL was further characterized in terms of specific activity, pH profile, kinetic parameters, and thermostability by comparing with αNAGAL purified from chicken liver. The data presented here suggest that by overexpressing rαNAGAL inP. pastorisand purifying with affinity chromatography one can readily obtain the quantity of enzyme needed for seroconversion studies.  相似文献   

7.
The cDNA coding for Penicillium purpurogenum α-galactosidase (αGal) was cloned and sequenced. The deduced amino acid sequence of the α-Gal cDNA showed that the mature enzyme consisted of 419 amino acid residues with a molecular mass of 46,334 Da. The derived amino acid sequence of the enzyme showed similarity to eukaryotic αGals from plants, animals, yeasts, and filamentous fungi. The highest similarity observed (57% identity) was to Trichoderma reesei AGLI. The cDNA was expressed in Saccharomyces cerevisiae under the control of the yeast GAL10 promoter. Almost all of the enzyme produced was secreted into the culture medium, and the expression level reached was approximately 0.2 g/liter. The recombinant enzyme purified to homogeneity was highly glycosylated, showed slightly higher specific activity, and exhibited properties almost identical to those of the native enzyme from P. purpurogenum in terms of the N-terminal amino acid sequence, thermoactivity, pH profile, and mode of action on galacto-oligosaccharides.α-Galactosidase (αGal) (EC 3.2.1.22) is of particular interest in view of its biotechnological applications. αGal from coffee beans demonstrates a relatively broad substrate specificity, cleaving a variety of terminal α-galactosyl residues, including blood group B antigens on the erythrocyte surface. Treatment of type B erythrocytes with coffee bean αGal results in specific removal of the terminal α-galactosyl residues, thus generating serological type O erythrocytes (8). Cyamopsis tetragonoloba (guar) αGal effectively liberates the α-galactosyl residue of galactomannan. Removal of a quantitative proportion of galactose moieties from guar gum by αGal improves the gelling properties of the polysaccharide and makes them comparable to those of locust bean gum (18). In the sugar beet industry, αGal has been used to increase the sucrose yield by eliminating raffinose, which prevents normal crystallization of beet sugar (28). Raffinose and stachyose in beans are known to cause flatulence. αGal has the potential to alleviate these symptoms, for instance, in the treatment of soybean milk (16).αGals are also known to occur widely in microorganisms, plants, and animals, and some of them have been purified and characterized (5). Dey et al. showed that αGals are classified into two groups based on their substrate specificity. One group is specific for low-Mr α-galactosides such as pNPGal (p-nitrophenyl-α-d-galactopyranoside), melibiose, and the raffinose family of oligosaccharides. The other group of αGals acts on galactomannans and also hydrolyzes low-Mr substrates to various extents (6).We have studied the substrate specificity of αGals by using galactomanno-oligosaccharides such as Gal3Man3 (63-mono-α-d-galactopyranosyl-β-1,4-mannotriose) and Gal3Man4 (63-mono-α-d-galactopyranosyl-β-1,4-mannotetraose). The structures of these galactomanno-oligosaccharides are shown in Fig. Fig.1.1. Mortierella vinacea αGal I (11) and yeast αGals (29) are specific for the Gal3Man3 having an α-galactosyl residue (designated the terminal α-galactosyl residue) attached to the O-6 position of the nonreducing end mannose of β-1,4-mannotriose. On the other hand, Aspergillus niger 5-16 αGal (12) and Penicillium purpurogenum αGal (25) show a preference for the Gal3Man4 having an α-galactosyl residue (designated the stubbed α-galactosyl residue) attached to the O-6 position of the third mannose from the reducing end of β-1,4-mannotetraose. The M. vinacea αGal II (26) acts on both substrates to almost equal extents. The difference in specificity may be ascribed to the tertiary structures of these enzymes. Open in a separate windowFIG. 1Structures of galactomanno-oligosaccharides.Genes encoding αGals have been cloned from various sources, including humans (3), plants (20, 32), yeasts (27), filamentous fungi (4, 17, 24, 26), and bacteria (1, 2, 15). αGals from eukaryotes show a considerable degree of similarity and are grouped into family 27 (10).Here we describe the cloning of P. purpurogenum αGal cDNA, its expression in Saccharomyces cerevisiae, and the purification and characterization of the recombinant enzyme.  相似文献   

8.
The cDNAs coding for Mortierella vinacea α-galactosidases I and II were expressed in Saccharomyces cerevisiae under the control of the yeast GAL10 promoter. The recombinant enzymes purified to homogeneity from the culture filtrate were glycosylated, and had properties identical to those of the native enzymes except for improving the heat stability of α-galactosidase II and decreasing the specific activities of both enzymes.  相似文献   

9.
Transforming growth factor- (TGF-) isoform expression by odontoblasts leads to their sequestration within the dentine matrix, from where they may be released during caries and participate in the reparative processes. Two receptor types for TGF- have been implicated in TGF- induced signalling. The aim of this study was to characterise immunohistochemically the expression of these receptors in sound and carious human teeth to facilitate our understanding of the ability of these cells to respond to TGF- stimulation. Sound and carious human teeth were routinely processed and paraffin sections stained for TGF- receptors I and II, using the StrAviGen immunoperoxidase method. Strong specific staining for both receptor types was observed in the odontoblasts of healthy teeth with the greatest intensity seen with receptor I. Staining of weaker intensity was also observed for both receptors in the underlying cell rich area and pulp core. Similar patterns of staining were observed within carious tissues. We conclude that odontoblasts and other cells of the pulp of mature human molar teeth show the presence of both TGF- receptors I and II in health and disease with odontoblasts showing the strongest expression. Such findings may be important in the response of these cells to tissue injury.  相似文献   

10.
Human α1-antitrypsin (AAT) is the most abundant serine proteinase inhibitor (serpin) in the human plasma. Commercially available AAT for the medications of deficiency of α1-antitrypsin is mainly purified from human plasma. There is a high demand for a stable and low-cost supply of recombinant AAT (rAAT). In this study, the baculovirus expression vector system using silkworm larvae as host was employed and a large amount of highly active AAT was recovered from the silkworm serum (~?15 mg/10 ml) with high purity. Both the enzymatic activity and stability of purified rAAT were comparable with those of commercial product. Our results provide an alternative method for mass production of the active rAAT in pharmaceutical use.  相似文献   

11.
Wall-bound α-glucosidase (EC 3.2.1.20) has been solubilized from suspension-cultured rice cells with Sumyzyme C and Pectolyase Y-23 and isolated by a procedure including fractionation with ammonium sulfate, Sephadex G-100 column chromatography, CM-cellulose column chroma-tography, Sephadex G-200 column chromatography, and preparative disc gel electrophoresis. The molecular weight of the enzyme was 64,000. The enzyme readily hydrolyzed maltose, maltotriose, and amylose, but hydrolyzed isomaltose and soluble starch more slowly. The Michaelis constant for maltose of the enzyme was estimated to be 0.272 mm. The enzyme produced panose as the main α- glucosyltransferred product from maltose.  相似文献   

12.
13.
Bone morphogenic protein (BMP)-7 is a member of the transforming growth factor (TGF)-beta superfamily, which is originally identified based on its ability to induce cartilage and bone formation. In recent years, BMP-7 is also defined as a potent promoter of cell motility, invasion, and metastasis. However, there is little knowledge of the role of BMP-7 and its cellular function in chondrosarcoma cells. In the present study, we investigated the biological impact of BMP-7 on cell motility using transwell assay. In addition, the intracellular signaling pathways were also investigated by pharmacological and genetic approaches. Our results demonstrated that treatment with exogenous BMP-7 markedly increased cell migration by activating c-Src/PI3K/Akt/IKK/NF-κB signaling pathway, resulting in the transactivation of αvβ3 integrin expression. Indeed, abrogation of signaling activation, by chemical inhibition or expression of a kinase dead form of the protein attenuated BMP-7-induced expression of integrin αvβ3 and cell migration. These findings may provide a useful tool for diagnostic/prognostic purposes and even therapeutically in late-stage chondrosarcoma as an anti-metastatic agent.  相似文献   

14.
α-Amylase from the antarctic psychrophile Alteromonas haloplanktis is synthesized at 0 ± 2°C by the wild strain. This heat-labile α-amylase folds correctly when overexpressed in Escherichia coli, providing the culture temperature is sufficiently low to avoid irreversible denaturation. In the described expression system, a compromise between enzyme stability and E. coli growth rate is reached at 18°C.Psychrophilic enzymes possess specific properties, such as high activity at low temperatures and weak thermal stability, which promise to allow the use of these enzymes as industrial biocatalysts, as biotechnological tools, or for fundamental research (6, 8, 11). For instance, substantial energy savings can be obtained if heating is not required during large-scale processes which take advantage of the efficient catalytic capacity of cold-adapted enzymes in the range 0 to 20°C. The pronounced heat lability of psychrophilic enzymes also allows their selective inactivation in a complex mixture, as illustrated by an antarctic bacterial alkaline phosphatase which is available for molecular biology research (7). Finally, psychrophilic enzymes represent the lower natural limit of protein stability (3) and are useful tools for studies in the field of protein folding.Large-scale fermentation of psychrophilic microorganisms suffers from two main drawbacks, however: the low production levels of wild strains and the prohibitive cost of growing wild strains at low temperatures. A possible alternative is to overexpress the gene coding for a psychrophilic protein in a mesophilic host for which efficient expression systems have been designed. In this context, two crucial questions remain to be solved: (i) what is the folding state of an enzyme normally synthesized at 0°C when it is expressed by the mesophilic genetic machinery at higher temperatures, and (ii) is there a temperature at which a compromise can be reached between the stability of the psychrophilic enzyme and the mesophilic growth rate? To address these questions, the heat-labile α-amylase from the antarctic psychrophile Alteromonas haloplanktis (2, 4) was expressed in Escherichia coli at various temperatures.

Construction of the expression vector and α-amylase production.

The α-amylase gene (2) was cloned downstream from the lacZ promoter in pUC12 by ligating the SmaI site of the polylinker to the HpaI site located 60 nucleotides upstream from the formylmethionine codon. This construction is devoid of the C-terminal peptide cleaved by the wild strain following α-amylase secretion. The recombinant enzyme was expressed in E. coli RR1 with the constitutive assistance of lacZ (without IPTG [isopropyl-β-d-thiogalactopyranoside] induction) in a medium containing 16 g of bactotryptone, 16 g of yeast extract, 5 g of NaCl, 2.5 g of K2HPO4, 0.1 μM CaCl2, and 100 mg of ampicillin per liter. The effect of the culture temperature on α-amylase production by E. coli is illustrated in Fig. Fig.1.1. Within the range of temperatures used, maximal enzyme production was reached below 18°C, whereas higher temperatures induced a gradual decrease of α-amylase activity in cultures. Three independent cultures were pooled for the purification of the recombinant enzymes produced at 18 and 25°C. Open in a separate windowFIG. 1Temperature dependence of α-amylase production by E. coli. Results are expressed as percent mean maximal activity recorded at 18°C.

α-Amylase purification.

The gram-negative A. haloplanktis was cultivated at 4°C, and α-amylase was purified from the culture supernatants by ion-exchange chromatography on DEAE-agarose followed by gel filtration on Sephadex G-100 and Ultrogel AcA54 as previously described (2, 4). The recombinant α-amylases were purified by the protocol developed for the wild-type enzyme except that concentration by ammonium sulfate precipitation at 70% saturation was required before the first chromatographic step. Recombinant enzyme production at 18 and 25°C ranged between 60 and 100 mg/liter of culture, which corresponds to a 10-fold improvement over production by the wild strain.

Characterization of the recombinant α-amylases.

N- and C-terminal amino acid sequences (determined on an Applied Biosystems Procise analyzer and by carboxypeptidase Y digestion, respectively) of α-amylase produced at 18 and 25°C indicated that the signal peptide is correctly cleaved in E. coli and that no additional posttranslational cleavage occurred. The isoelectric point (5.5) and the molecular mass (49,340 Da as determined from the sequence and 49,342 ± 8 Da as determined from electrospray mass spectroscopy measurements) were identical to the values recorded for the wild-type enzyme. Dynamic light scattering (DynaPro-801; DLS Instruments) also showed that the purified recombinant enzymes are homogeneous, without any evidence of aggregated forms.

Comparison of the wild-type and recombinant α-amylases.

Several properties of the wild-type enzyme produced at 4°C and the recombinant α-amylase expressed in E. coli at 18°C were compared (Table (Table1).1).

TABLE 1

Kinetic parameters, dissociation constants, and free thiol groups for the wild-type and recombinant α-amylases
α-Amylasekcat (s−1)Km (μM)kcat/Km (s−1 · μM−1)Kd
Cysteinesa (mol−1)Free thiol (mol−1)
Cl (mM)Ca (M)
Wild-type (produced at 4°C)780 ± 25174 ± 84.65.9 ± 0.22.10−880.03
Recombinant (produced at 18°C)792 ± 34168 ± 144.76.1 ± 0.22.10−880.05
Recombinant (produced at 25°C)609 ± 29186 ± 223.36.0 ± 0.32.10−880.05
Open in a separate windowaFrom the amino acid sequence. 

(i) Kinetic and ion binding parameters.

4-Nitrophenyl-α-d-maltoheptaoside-4,6-O-ethylidene (EPS) was used as the substrate in a coupled assay with α-glucosidase at 25°C. The absorption coefficient for 4-nitrophenol was 8,990 M−1 · cm−1 at 405 nm, and a stoichiometric factor of 1.25 was applied for kcat (turnover number) calculation. Dissociation constants were determined by activation kinetics following Cl or Ca2+ titration of the apoenzyme obtained by dialysis against 25 mM HEPES-NaOH (pH 7.2) and 25 mM HEPES-NaOH–5 mM EGTA (pH 8.0), respectively. The saturation curves were computer fitted by a nonlinear regression analysis of the Hill equation in the form v = kcat [I]h/Kd + [I]h where [I] is the ion concentration and h is the Hill coefficient. The free calcium concentrations were set by calcium titration in the presence of 5 mM EGTA at pH 8.0. Kinetic parameters (kcat, Km and kcat/Km) for the hydrolysis of EPS as well as dissociation constants (Kd) for calcium and chloride ions were found to be identical in the wild-type and recombinant enzymes produced at 18°C (Table (Table1).1). Owing to the stringent structural requirements for functional active site and ion binding site conformation, it can be safely concluded that the recombinant enzyme is properly folded at 18°C.

(ii) Disulfide bond integrity.

Free thiol content was determined by DTNB (5,5′-dithiobis-2-nitrobenzoic acid) titration of the unfolded enzyme in 8 M urea in order to promote −SH group accessibility. The eight cysteine residues of A. haloplanktis α-amylase are engaged in disulfide linkages (4). Thus, the lack of free sulfhydryl groups, as detected by DTNB titration of both the native and the unfolded enzymes (Table (Table1),1), indicates that the four disulfide bonds are formed in the recombinant α-amylase samples.

(iii) Conformational stability.

Fluorescence intensity of α-amylases (50 μg/ml) was recorded in 30 mM MOPS (morpholinepropanesulfonic acid)–50 mM NaCl–1 mM CaCl2 (pH 7.2) at a scanning rate of 1°C/min and at an excitation wavelength of 280 nm and an emission wavelength of 347 nm with a Perkin-Elmer LS 50 spectrofluorimeter. Raw data were corrected for thermal dependence of the fluorescence by using the slopes of the pre- and posttransition regions as described elsewhere (10). The conformational stability (ΔGN⇔U) was determined by reversible, thermally induced unfolding recorded by fluorescence. Both the wild-type and the recombinant α-amylases have melting point (Tm) values of 45 ± 0.2°C and display the same cooperative transition (Fig. (Fig.2).2). Consequently, plots of ΔG as a function of T (constructed by using the relation ΔG = −RTlnK, where K = fraction unfolded/fraction folded) are similar (Fig. (Fig.2,2, inset). These results indicate that the weak interactions stabilizing the folded state of the wild-type and recombinant α-amylases are quantitatively identical. Open in a separate windowFIG. 2Heat-induced unfolding transitions of the wild-type α-amylase (•) and the recombinant enzyme produced at 18°C (○). The fraction of protein in the unfolded state (fU) was calculated as follows: fU = (yF − y)/(yF − yU), where yF and yU are the fluorescence intensities of the native and the fully unfolded states, respectively, and y is the fluorescence intensity at a given temperature. The inset shows a plot of ΔG as a function of the temperature around the melting point (Tm), where ΔG = 0.

Expression at 25 and 37°C.

When cultures of the recombinant E. coli are carried out at 25°C, all parameters determined by activation kinetics and independent of the enzyme concentration, such as Km and Kd, remain constant, as does the free sulfhydryl content (Table (Table1).1). This indicates that the native enzyme fraction is correctly folded. By contrast, the kcat of the recombinant α-amylase is reduced by about 20%, suggesting the occurrence of a corresponding inactive fraction. When expressed at 37°C, no α-amylase activity is recorded; the recombinant heat-labile enzyme could fail to fold at this high temperature, or its denaturation rate could exceed its synthesis rate. Furthermore, Western blotting with rabbit polyclonal antibodies to α-amylase detects only trace amounts of the recombinant gene product, suggesting that the denatured enzyme is quickly degraded by the E. coli cell.

Conclusions.

We have previously shown that cloning of a psychrophilic gene in E. coli and detection of the gene product can be achieved by careful control of the culture conditions: overnight incubation at 25°C of transformed cells followed by 1 to 2 days of incubation at 4°C produced halos of substrate hydrolysis on agar plates (5). The folding state of the recombinant psychrophilic enzymes (e.g., fully or partly active, native or marginal stability, etc.), however, was unknown. The results presented here demonstrate that the genuine properties of a psychrophilic enzyme are preserved when it is expressed in a mesophilic host, providing the culture temperature is sufficiently low to allow correct folding and to avoid irreversible denaturation. In our expression system, a compromise is reached between the stability of the psychrophilic enzyme and the growth rate of the mesophilic host by cultivating the recombinant E. coli at 18°C. It should be noted that commonly used E. coli strains have different growing capacities at that temperature. We found E. coli RR1, HB101, or XL1-Blue (Stratagene) suitable for these culture conditions (the generation times are about 3 h, and stationary phase is reached after approximately 30 h), whereas E. coli DH5α grows twice as slowly at 18°C.The lack of α-amylase expression at 37°C is not an isolated case: under the same conditions, lipases and proteases (1, 5, 9) from antarctic psychrophiles were not expressed in an active form. This illustrates the general heat lability of psychrophilic enzymes, which is thought to arise from their flexible conformation, allowing high catalytic activity at temperatures close to 0°C (3).  相似文献   

15.
A recombinant human consensus interferon-α mutant (cIFN) was expressed in Pichia pastoris. The maximum dry cell weight, cIFN concentration and antiviral activity were 160 g l−1, 1.24 g l−1 and 4.1 × 107 IU ml−1, respec tively. The cIFN secreted into the medium was in the form of aggregates dominantly by non-covalent interaction and partially by disulphide bond. When the fermentation supernatant was disaggregated with 6 M guanidine hydrochloride, the antiviral activity of cIFN achieved 2.2 × 108 IU ml−1.  相似文献   

16.
Membrane progestin receptors (mPRs) are responsible for mediating the rapid, nongenomic activity of progestins and belong to the G protein-coupled receptor (GPCR) family. mPRs are also considered as attractive proteins to draw a new medicinal approach. In this study, we optimized a procedure for the expression and purification of recombinant human mPRα protein (hmPRα) by a methylotropic yeast, Pichia pastoris, expression system. The protein expressed in crude membrane fractions exhibited a binding affinity of Kd = 3.8 nM and Bmax = 288.8 fmol/mg for progesterone. These results indicated that the hmPRα expressed in yeast was active. Solubilized hmPRα was purified through three column chromatography steps. A nickel-nitrilotriacetic acid (Ni-NTA) column was first used, and the mPRα proteins were then bound to cellulose resin with free amino groups (Cellufine Amino) and finally passed through an SP-Sepharose column. The optimization of expression and purification conditions resulted in a high yield of purified hmPRα (1.3–1.5 mg from 1 L culture). The purified hmPRα protein demonstrated progesterone binding (Kd = 5.2 nM and Bmax = 111.6 fmol/mg). The results indicated that we succeeded in solubilizing and purifying hmPRα in an active form. Sufficient amount of active hmPRα protein will support the establishment of applications for the screening of ligands for mPRα.  相似文献   

17.
ADENOVIRUS infection of human embryonic kidney (HEK) cultures seems to induce cellular RNA synthesis, which is preceded by a transient increase in the activities of the Mg2+-activated and Mn2+-(NH4)2SO4-activated DNA dependent RNA polymerases and in the rate of histone acetylation1. The two polymerase reactions, assayed in isolated cell nuclei, apparently reflect the activities of distinct nucleolar and nucleo-plasmic RNA polymerases2,3. We were therefore prompted to test the effect of a specific inhibitor of the mammalian DNA-dependent RNA polymerase function, α-amanitin, on the multiplication of adenovirus. α-Amanitin is a bicyclic octapeptide isolated from the poisonous mushroom Amanita phalloides4 and which blocks RNA synthesis in intact animals5,6. Nuclei isolated from the livers of such animals show a reduced activity of the RNA polymerases associated with nucleoplasm5,6 and the nucleolus6.  相似文献   

18.
The integrin α4β1(VLA4) has been expressed as a soluble, active, heterodimeric immunoglobulin fusion protein. cDNAs encoding the extracellular domains of the human α4 and β1 subunits were fused to the genomic DNA encoding the human γ1 immunoglobulin Fc domain and functional integrin fusion protein was expressed as a secreted, soluble molecule from a range of mammalian cell lines. Specific mutations were introduced into the Fc region of the molecules to promote α4β1 heterodimer formation. The soluble α4β1 Fc fusion protein exhibited divalent cation dependent binding to VCAM-1, which was blocked by the appropriate function blocking antibodies. The apparent Kd for VCAM-1 binding were similar for both the soluble and native forms of α4β1. In addition, the integrin–Fc fusion was shown to stain cells expressing VCAM-1 on their surface by FACs analysis. This approach for expressing soluble α4β1 should be generally applicable to a range of integrins.  相似文献   

19.
The α-methylserine aldolase gene from Variovorax paradoxus strains AJ110406, NBRC15149, and NBRC15150 was cloned and expressed in Escherichia coli. Formaldehyde release activity from α-methyl-L-serine was detected in the cell-free extract of E.coli expressing the gene from three strains. The recombinant enzyme from V. paradoxus NBRC15150 was purified. The V max and K m of the enzyme for the formaldehyde release reaction from α-methyl-L-serine were 1.89 μmol min?1 mg?1 and 1.2 mM respectively. The enzyme was also capable of catalyzing the synthesis of α-methyl-L-serine and α-ethyl-L-serine from L-alanine and L-2-aminobutyric acid respectively, accompanied by hydroxymethyl transfer from formaldehyde. The purified enzyme also catalyzed alanine racemization. It contained 1 mole of pyridoxal 5′-phosphate per mol of the enzyme subunit, and exhibited a specific spectral peak at 429 nm. With L-alanine and L-2-aminobutyric acid as substrates, the specific peak, assumed to be a result of the formation of a quinonoid intermediate, increased at 498 nm and 500 nm respectively.  相似文献   

20.
Alport disease in humans, which usually results in proteinuria and kidney failure, is caused by mutations to the COL4A3, COL4A4, or COL4A5 genes, and absence of collagen α3α4α5(IV) networks found in mature kidney glomerular basement membrane (GBM). The Alport mouse harbors a deletion of the Col4a3 gene, which also results in the lack of GBM collagen α3α4α5(IV). This animal model shares many features with human Alport patients, including the retention of collagen α1α2α1(IV) in GBMs, effacement of podocyte foot processes, gradual loss of glomerular barrier properties, and progression to renal failure. To learn more about the pathogenesis of Alport disease, we undertook a discovery proteomics approach to identify proteins that were differentially expressed in glomeruli purified from Alport and wild-type mouse kidneys. Pairs of cy3- and cy5-labeled extracts from 5-week old Alport and wild-type glomeruli, respectively, underwent 2-dimensional difference gel electrophoresis. Differentially expressed proteins were digested with trypsin and prepared for mass spectrometry, peptide ion mapping/fingerprinting, and protein identification through database searching. The intermediate filament protein, vimentin, was upregulated ∼2.5 fold in Alport glomeruli compared to wild-type. Upregulation was confirmed by quantitative real time RT-PCR of isolated Alport glomeruli (5.4 fold over wild-type), and quantitative confocal immunofluorescence microscopy localized over-expressed vimentin specifically to Alport podocytes. We next hypothesized that increases in vimentin abundance might affect the basement membrane protein receptors, integrins, and screened Alport and wild-type glomeruli for expression of integrins likely to be the main receptors for GBM type IV collagen and laminin. Quantitative immunofluorescence showed an increase in integrin α1 expression in Alport mesangial cells and an increase in integrin α3 in Alport podocytes. We conclude that overexpression of mesangial integrin α1 and podocyte vimentin and integrin α3 may be important features of glomerular Alport disease, possibly affecting cell-signaling, cell shape and cellular adhesion to the GBM.  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