首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The relative contribution to development of hepatocellular carcinoma of the mouse equivalent to the human p53ser249 mutation, found in human hepatocellular carcinoma associated with aflatoxin (AFB1) exposure, is compared with other major risk factors in a transgenic mouse model. Transgenic p53ser246 mice, expressing the mutant protein gene under the control of a truncated albumin promoter, were bred to mice lacking p53 (p53-/-) and to transgenic mice expressing hepatitis B surface antigen (HBsAg). AFB1 hepatocarcinogenesis was then determined in offspring with single or multiple risk factors by determination of the numbers of high-grade hepatic tumors at 13 months of age. In AFB1-treated male mice, expression of the p53ser246 mutation increases the incidence of high-grade tumors from 0% to 14% in HBsAg-negative, p53+/+ (wild-type homozygous) control mice; from 14% to 71% in HBsAg-negative, p53+/- (wild-type heterozygous) mice; and from 62% to 100% in HBsAg-positive, p53+/+ mice. Thus, whereas HBsAg expression and AFB1 together are strongly cocarcinogenic, the presence of the p53ser246 mutant not only significantly enhances this cocarcinogenic effect, it also increases tumorigenesis in AFB1-treated p53 heterozygous and homozygous mice not expressing HBsAg. The possibility that the p53ser246 mutant protein may act as a promoting agent for AFB1 hepatocarcinogenesis is discussed.  相似文献   

2.
Infection with hepadnaviruses and exposure to aflatoxin B1 (AFB1) are considered major risk factors in the development of hepatocellular carcinoma (HCC) in humans and in animals. A high rate of mutations in the p53 tumor suppressor gene in hepatocellular carcinomas of predominantly hepatitis B virus (HBV) carrier patients has been recently related to dietary aflatoxin. Another member of the hepadnavirus family, the woodchuck hepatitis virus (WHV), infects woodchucks in a manner similar to that of HBV in humans. Therefore, it was of particular interest to determine whether the p53 gene in woodchuck HCCs associated with hepadnavirus infection and with exposure to AFB1 is affected in the same manner as in human HCCs. By direct PCR-sequencing, we analyzed exons 4-9 of the p53 gene in 13 HCCs from 12 woodchucks (two uninfected, ten WHV carriers). Six WHV carrier and two uninfected woodchucks were treated with AFB1. None of the analyzed HCC samples exhibited mutations, either in p53 gene exons 4-9, or in splicing donor-acceptor sites. The present data are consistent with our previous study that indicated a low rate of p53 mutations in HCCs of AFB1-treated ground squirrels, either infected or not infected with ground squirrel hepatitis virus, and in WHV carrier woodchucks not exposed to AFB1. Overall, our findings indicate that in woodchucks and in ground squirrels exposure to aflatoxin may affect the development of p53 mutations less than in humans.  相似文献   

3.
Recent studies have revealed that a point mutation at codon 249 in the p53 gene predominates in hepatocellular carcinoma (HCC) cases from Southern Africa and China, where infection with hepatitis B virus (HBV) and contamination of aflatoxin B1 in food are risk factors for HCC. This unique mutation from G to T at the third base in codon 249 observed in human HCC cases is suggested to be linked to aflatoxin exposure. Six ducks with HCC, five of which were fed a diet containing aflatoxin B1 for 1-2 years, were analysed for the presence of point mutations at this codon of the p53 gene by polymerase chain reaction and direct nucleotide sequencing. None of the six ducks with HCC showed the change at this codon regardless of duck hepatitis B virus infection. This suggests that aflatoxin B1 itself might not be involved in the unique mutation at codon 249 in hepatocarcinogenesis, or that other factors coincident with aflatoxin may be responsible for this unique mutation.  相似文献   

4.
Sequence-dependent formation and lack of repair of polycyclic aromatic hydrocarbon-induced DNA adducts correlates well with the positions of p53 mutational hotspots in smoking-related lung cancers (Denissenko et al, 1996, 1998). The mycotoxin aflatoxin B1 (AFB1) is considered to be a major causative agent in hepatocellular carcinoma (HCC) in regions with presumed high food contamination by AFB1. A unique mutational hotspot, a G to T transversion at the third base of codon 249 of the p53 gene is observed in these tumors. To test whether a selectivity of AFB1 adduct formation is related to this peculiar mutational spectrum, we have mapped AFB1-DNA adducts at nucleotide resolution using ligation-mediated PCR and terminal transferase-dependent PCR. Human HepG2 cells were exposed to AFB1 metabolically activated in the presence of rat liver microsomes. Significant adduct formation was seen at the third base of codon 249. However, this was not the major site of AFB1 adducts and strong adduction was also observed at codons 226, 243, 244, 245 and 248 in exon 7 of the p53 gene and at several codons in exon 8. The damage at codon 249 does not consist of a unique abasic site or ring-opened aflatoxin B1 adduct but rather is consistent with the principal N7-guanine adduct of AFB1. Time course experiments indicate that, under the conditions used, AFB1 adducts are not removed in a strand-selective manner and adduct removal from the third base of codon 249 proceeds at a relatively fast rate (50% in 7 h). The incomplete correspondence between sites of persistent AFB1 damage and the specific codon 249 mutation suggests that AFB1 may not be involved in mutation of this site or that additional mechanisms such as parallel infection with hepatitis B virus may be required for selection of codon 249 mutants in HCC.  相似文献   

5.
To clarify the relative role of hepatitis C virus (HCV) and hepatitis B virus (HBV) in hepatocarcinogenesis in hepatitis B surface antigen (HBsAg)-negative hepatocellular carcinoma (HCC) in Taiwan, polymerase chain reaction (PCR) was used to detect the HCV-RNA and HBV-DNA sequences in the serum and liver tissues from 31 HBsAg-negative HCC patients. Twenty-one were positive for antibody to HCV (group 1) and 10 were negative (group 2). Hepatitis C virus-RNA was detected by PCR in the serum of 16 group 1 patients and in the liver tissue of 17; while HBV-DNA was found in the liver tissue of only four, and no HBV-DNA was found in the serum. Hepatitis C virus RNA was detected in the serum of one group 2 patient and in the liver tissue of another. In contrast, HBV viral DNA was found in the serum of four group 2 patients and in the liver tissues of five patients. This indicates that HCV plays an important role in hepatocarcinogenesis in HBsAg-negative patients in Taiwan, especially in those with antibody to HCV. In those without antibody to HCV, HBV might still be associated with the development of HCC in a significant proportion of such patients. In order to study the role of the p53 mutation in hepatocarcinogenesis, we investigated the status of the p53 mutation in 61 HCC samples from Taiwan. The exon 5 to 8 of the p53 gene in the tumour tissue of 61 HCC were amplified and sequenced. A total of 20 cases (32.8%) were found to have mutations: 36.6% (15/41) from the HBsAg-positive group and 25.0% (5/20) from the HBsAg-negative group. The corresponding normal liver showed no mutation. The mutation is widely distributed throughout the exon 5 to 8. Only four cases (6.6%), all positive for HBsAg, had a specific hotspot mutation at codon 249 with G to T transversion. These results show that scattered point mutations in p53 are not uncommon in HCC samples from Taiwan and may be important in the development of this cancer. However, the aflatoxin-related specific mutation seems much less related to the genesis of HCC in Taiwan. To study the role of telomerase activity in hepatocarcinogenesis, a total of 39 HCC tissues and the corresponding non-tumour liver tissues were analysed. The results showed that telomerase activity was detected in all the 39 tumour tissues, while it could be detected in six of the 39 non-tumour liver tissues. The high positive rate of telomerase activity in HCC samples suggests that telomerase activity is closely related to the development or progression of HCC. To determine whether exon 1 and exon 2 of the p16 gene are altered in HCC, thirty-four tumours from 30 HCC patients were examined by DNA sequencing analysis of PCR-amplified genomic DNA. Homozygous deletions of MTS1/p16/CDKN2 exon 1 were identified in 1/34 primary tumours (3%), no mutations or rearrangements were found in these specimens. These data suggest that alterations of MTS1/p16/CDKN2 gene are rarely found in HCC, and might play little role in the development of this cancer. To study the clonality of HCC, 18 patients with multiple HCC, most of them small in size, were analysed by DNA fingerprinting. In patients positive for hepatitis B surface antigen, the integration pattern of hepatitis B viral DNA in liver tissue was also analysed. The results by both methods showed that 8/9 hepatitis B surface antigen-positive patients were different in clonality. In the remaining nine patients negative for hepatitis B surface antigen, four had different band patterns in their tumours by DNA fingerprinting. This study indicated that polyclonality of multiple HCC was rather frequent and it highlighted the importance of eliminating the underlying cause of liver injury to improve the survival of these patients. Microsatellite markers were used to study the genetic changes of HCC. Thirty cases of HCC, most of them small in size, were studied. A total of 242 microsatellite markers mapping to 1-22 and X chromosomes was used. The results showed that the range of loss of het  相似文献   

6.
The tumor-suppressing phenotype of p53 is thought to be due to its accumulation in response to DNA damage and resultant cell cycle arrest or apoptosis. scid/scid mice are defective in DNA double-strand break repair due to a mutation in DNA-dependent protein kinase (DNAPK). Treatment of scid/scid mice with gamma radiation or N-ethyl-N-nitrosourea resulted in approximately 86% incidence of T-cell lymphomas, compared with <6% in wild-type mice. The incidence of other tumor types was not increased in scid/scid mice, suggesting that the types of DNA double-strand break that are unrepaired in these mice are not strongly carcinogenic. To determine whether mutations in DNAPK and p53 interact, we examined mice deficient in both genes. Both scid/scid p53-/- and scid/scid p53+/- mice spontaneously developed lymphomas at shorter latency than did mice with either defect alone. Loss of the wild-type p53 allele was observed in 100% of tumors from scid/scid p53 +/- mice, indicating strong selection against p53. In contrast, p53 was not inactivated in lymphomas from scid/scid p53+/+ mice. Exposure of these tumor-bearing mice to gamma radiation resulted in p53 protein accumulation and high levels of apoptosis in all tumors that were not observed in tumors from scid/scid p53+/- mice. Thus, there was a bifurcation of molecular pathways to tumorigenesis. When p53 was heterozygous in the germ line, loss of the wild-type allele occurred, and the tumors became apoptosis resistant. When p53 was wild type in the germ line, p53 was not inactivated, and the tumors remained highly apoptosis sensitive.  相似文献   

7.
BACKGROUND: Liver cancer mortality in Korea is the highest in the world. Hepatitis B and C viruses (HBV, HCV) are known to be the major risk factors of hepatocellular carcinoma (HCC). Cholangiocarcinoma (CLG) accounts for more than 20% of liver cancer in the Pusan area. In Korea, the different roles of known risk factors in the development of HCC or CLG have not been adequately evaluated. METHODS: Case-control studies involved 203 incident HCC cases, 406 controls matched to the HCC cases for age (+/- 4 years) and sex, and 41 CLG cases (the HCC controls were used). They were carried out from August 1990 to August 1993. RESULTS: Relative risk (RR) of HBsAg (87.4; 95% confidence interval [CI]: 22.2-344.3) and RR of anti-HCV positivity (30.3; 95% CI: 6.1-150.6) were significant for the risk of HCC after adjustment for potentially confounding factors. In contrast, RR of Clonorchis sinensis in stool (2.7; 95% CI: 1.1-6.3) and RR of heavy drinking (4.6; 95% CI: 1.4-15.2) were significant for the risk of CLG. Transfusion history, acupuncture history, and cigarette smoking were not associated with the risk of HCC or CLG. CONCLUSIONS: Strong evidence indicated that both HBV and HCV infection were independent risk factors for HCC. In contrast, C. sinensis in stools and heavy drinking were associated with the risk of CLG in Korea.  相似文献   

8.
BACKGROUND/AIMS: A great number of hepatocellular carcinoma (HCC) develop from chronic liver disease. Among a total of 23,000 deaths of HCC in 1988 in Japan, 82% had positive antibodies against HBV and/or HCV. In the present study we investigated the etiological factors involved in this process, employing patients with chronic hepatitis as controls. METHODOLOGY: In this study, alcohol consumption and cigarette smoking were investigated in 104 male patients with HCC which developed from chronic liver disease and 104 male controls with chronic liver disease without HCC (one for each case) matched for age. RESULTS: When compared with non-drinkers and non-smokers, the relative risk (RR) for developing HCC rose to 17.9 among those with both drinking and smoking habits. The risk was greater than for those in whom either habit existed alone. The RR decreased among ex-smokers who were non-drinkers or ex-drinkers, but it was still as high as 9.4. For current smokers, even if they were non- or ex-drinkers, the RR was 15.4. CONCLUSIONS: Drinking and the cigarette smoking were both risk factors, but the existence of synergism between them was also suggested. Therefore, patients with chronic liver disease should be thoroughly counseled to refrain from both drinking and smoking.  相似文献   

9.
Aberrations of the p53 gene were observed in high frequency in HCCs in China, Mozambique and Japan. Most of the mutations were G to T transversions in codon 249 of the p53 gene in HCCs in China and Mozambique, where aflatoxin B1 was a risk factor. These findings strongly suggest that aflatoxin B1 induced this type of mutations. On the other hand, in Japanese HCCs, no mutation was observed in codon 249, indicating a different cause of HCC development. Pathological analyses revealed that p53 mutations were detected only in advanced HCCs and not in early HCCs. In addition, aberrations of the RB gene were observed in only tumors carrying mutated p53 gene. These results indicate the involvement of at least two tumor suppressor genes in a late stage of hepatocarcinogenesis. Possible mechanisms of HCC formation are discussed.  相似文献   

10.
There is strong evidence that selenium protects against certain human cancers, but the underlying mechanism is unknown. Glutathione peroxidase (GPX1) and thioredoxin reductase (TR), the most abundant antioxidant selenium-containing proteins in mammals, have been implicated in this protection. We analyzed the expression of TR and GPX1 in the following model cancer systems: (1) liver tumors in TGFalpha/c-myc transgenic mice; (2) human prostate cell lines from normal and cancer tissues; and (3) p53-induced apoptosis in a human colon cancer cell line. TR was induced while GPX1 was repressed in malignancies relative to controls in transgenic mice and prostate cell lines. In the colon cell line, p53 expression resulted in elevated GPX1, but repressed TR. The data indicate that TR and GPX1 are regulated in a contrasting manner in the cancer systems tested and reveal the p53-dependent regulation of selenoprotein expression. The data suggest that additional studies on selenoprotein regulation in different cancers are required to evaluate future implementation of selenium as a dietary supplement in individuals at risk for developing certain cancers.  相似文献   

11.
BACKGROUND/AIMS: Aflatoxins (AFs) are established hepatic carcinogens in several animal species. This study was performed to establish whether aflatoxin exposure may affect the risk of developing hepatocellular carcinoma in chronic hepatitis B virus carriers. METHODS: Urinary AF metabolites were measured for 43 HCC cases and 86 matched controls nested in a cohort of 7342 men in Taiwan. Thirty hepatocellular carcinoma cases and 63 controls were also tested for AFB1-albumin adducts. RESULTS: There was a dose-response relationship between urinary AFM1 levels and risk of hepatocellular carcinoma in chronic hepatitis B virus carriers. Comparing the highest with the lowest tertile of urinary AFM1 levels, the multivariate-adjusted odds ratio (OR) was 6.0 (95% confidence interval (CI) = 1.2-29.0). The hepatocellular carcinoma risk associated with AFB1 exposure was more striking among the hepatitis B virus carriers with detectable AFB1-N7-guanine adducts in urine. Compared with chronic hepatitis B virus carriers who were negative for AFB1-albumin adducts and urinary AFB1-N7-guanine, no elevated risk was observed for those who were positive for either marker. But an extremely high risk of hepatocellular carcinoma among those having both markers was found (OR = 10.0, 95% CI = 1.6-60.9). The proportion of AFB1 converted to AFM1 decreased with the progress of liver disease, whereas the formation of AFP1 increased. The difference in patterns of AFB1 metabolite formation was an independent risk factor for hepatocellular carcinoma after adjustment for total AFB1 excretion. There was a synergistic interaction between glutathione S-transferase M1 genotype and AFB1 exposure in hepatocellular carcinoma risk. CONCLUSIONS: AFB1 intake and expression of enzymes involved in AFB1 activation/detoxification may play an important role in hepatitis B virus-related hepatocarcinogenesis.  相似文献   

12.
Patients with hepatitis C have an increased risk of developing hepatocellular carcinoma (HCC). This is related to the stage of chronic liver disease, as characterized histologically by hepatic fibrosis and architectural distortion, but it is unclear whether histological markers can define the risk of developing HCC. We conducted a case-control immunohistochemical study of Ki-67, a marker for hepatocellular proliferation, in livers of 18 patients who had developed HCC more than 2 years after the biopsy specimen had been taken. Using conditional logistic regression analysis, the results were compared with 18 selected controls, who were age-matched patients with hepatitis C of similar histological stage who had not developed HCC. We also examined livers for cellular dysplasia, p53 mutations, and bcl-2 overexpression, and assessed whether the results could be correlated with demographic and disease-related variables, such as gender, region of birth, alcohol consumption, severity of liver disease, HCV genotype, and markers of hepatitis B virus (HBV) infection. Livers from patients who developed HCC were more often positive for Ki-67 (13 of 18 [72%] v 9 of 18 [50%]; P = .06) and tended to have higher mean Ki-67 scores (6 +/- 7.5 v 3 +/- 4.4; P = .10) compared with control cases. In the HCC-predisposed group, three livers showed large cell dysplasia, two were positive for p53 mutations, and two for bcl-2 overexpression. In contrast, in the non-HCC group, only one case had dysplasia, and none were positive for immunostaining for p53 or bcl-2 mutations. With the exception of one case, all livers with large cell dysplasia or p53 mutations and bcl-2 overexpression were also positive for Ki-67. Twelve (55%) of the 22 Ki-67-positive cases were anti-HBc-positive in the serum, in contrast to 2 of 14 (14%) patients in the Ki-67-negative group (P = .01). Patients with evidence of past infection with HBV were more often Ki-67 positive than those who had no evidence of past infection (85% [11 of 13] v 45% [10 of 22]; P = .02). There were no other associations between demographic or disease-related variables and Ki-67 expression. Increased hepatocellular proliferative activity, as assessed by Ki-67 expression, may be one factor indicative of an increased risk of developing HCC among patients with chronic hepatitis C. Furthermore, past infection with HBV appears to be an important correlate of increased hepatocellular proliferation in hepatitis C.  相似文献   

13.
14.
Tumor suppressor genes are generally viewed as being recessive at the cellular level, so that mutation or loss of both tumor suppressor alleles is a prerequisite for tumor formation. The tumor suppressor gene, p53, is mutated in approximately 50% of human sporadic cancers and in an inherited cancer predisposition (Li-Fraumeni syndrome). We have analyzed the status of the wild-type p53 allele in tumors taken from p53-deficient heterozygous (p53+/-) mice. These mice inherit a single null p53 allele and develop tumors much earlier than those mice with two functional copies of wild-type p53. We present evidence that a high proportion of the tumors from the p53+/- mice retain an intact, functional, wild-type p53 allele. Unlike p53+/- tumors which lose their wild-type allele, the tumors which retain an intact p53 allele express p53 protein that induces apoptosis following gamma-irradiation, activates p21(WAF1/CIP1) and Mdm2 expression, represses PCNA expression (a negatively regulated target of wild-type p53), shows high levels of binding to oligonucleotides containing a wild-type p53 response element and prevents chromosomal instability as measured by comparative genomic hybridization. These results indicate that loss of both p53 alleles is not a prerequisite for tumor formation and that mere reduction in p53 levels may be sufficient to promote tumorigenesis.  相似文献   

15.
Using a urinary immunoassay to measure aflatoxin metabolites, we examined the associations between exposure to aflatoxin, chronic infection with the hepatitis-B virus (HBV) and background rates of hepatocellular carcinoma (HCC) mortality in a cross-sectional survey of 250 residents from 8 areas of Taiwan with a 4-fold variation in age-adjusted HCC mortality. Specimens of fasting blood and overnight urines were used to determine HBV carrier status and excretion of aflatoxin in the subjects surveyed. While the prevalence of hepatitis-B virus carriers showed moderate variability, there was a 500-fold range in urinary aflatoxin levels. Mean log-transformed levels of aflatoxin metabolites were similar in males and females and in HBV carriers and non-carriers. In the 8 townships, HCC mortality correlated positively with both area HBV carrier prevalence and mean aflatoxin levels. The primary analyses, however, were conducted at the individual level. Each subject's aflatoxin level was treated as the response variable in a multiple regression model, and the corresponding sex-specific area HCC rate was included as a predictor along with the individual's carrier status, age and sex; alcohol consumption and cigarette smoking were also considered. In these analyses, a significant association was again observed between the marker of aflatoxin exposure and the background rate of HCC mortality. In females, the slope of the regression line was somewhat steeper in HBV carriers, but this pattern was not seen in males and formal testing yielded no statistically significant evidence of an interaction. Our findings are consistent with the hypothesis that aflatoxin plays an independent role in hepatocellular carcinoma in Taiwan.  相似文献   

16.
There is strong epidemiological evidence that the hepatitis B virus (HBV) contributes to the development of hepatocellular carcinoma (HCC). In several immortalized cell lines, an in vitro transforming activity of HBV DNA and expression vectors for the viral protein X (HBx) has now been demonstrated. Furthermore, it appears as if still unknown parts of the HBV genome other than HBx contribute to the transforming activity of HBV DNA in vitro. Only one of several studies found that HBx-transgenic mouse lines develop HCC. A mouse line transgenic for the large surface protein of HBV develops HCC due to concomitant necroinflammatory infection. Growing evidence shows the importance of recombination of integrated viral DNA and cellular DNA for HCC development. A direct transforming potential of one of these viral integrates has been demonstrated. Chemical carcinogens are more effective in HBV-containing cell lines or transgenic mice.  相似文献   

17.
p53 gene mutations occur in most human cancers and result in an altered protein product that accumulates within the cell. Although the observed endogenous human CTL response to p53 is weak, high-affinity, human p53-specific CTLs have been generated from HLA A2.1 transgenic mice immunized with human CTL epitope peptides. In this study, we examine the ability of HLA A2.1-restricted and human p53-specific CTLs from HLA A2.1 transgenic mice to suppress the growth of p53-overexpressing human tumors in severe combined immunodeficient (SCID) mice. In vitro, murine p53(149-157)-specific CTLs selectively lysed the p53-overexpressing pancreatic carcinoma cell line Panc-1 but did not recognize HLA A2.1- tumor cells or HLA A2.1+ normal human fibroblasts. Furthermore, in vivo, the growth of established human tumor xenografts in SCID mice was significantly reduced and survival was prolonged after the administration of p53-specific CTLs but not after the administration of control CTLs or PBS alone. Following treatment with p53(149-157)-specific CTLs, regressing Panc-1 tumors were infiltrated by the CD8+ CTLs, as demonstrated by immunohistochemistry. These findings suggest that p53(149-157)-specific and HLA A2.1-restricted murine CTLs suppress the growth of established Panc-1 tumors following adoptive transfer into SCID hosts and prolong their survival.  相似文献   

18.
19.
To elucidate the risk factors for hepatocellular carcinoma (HCC) in hepatitis C virus (HCV)-related liver cirrhosis (LC), we examined 204 cirrhotic patients negative for hepatitis B surface antigen and positive for HCV antibodies. The independent influence of various clinical characteristics in these patients was analyzed by multiple logistic regression, and the risk factors for HCC were identified. Multiple logistic regression analysis identified and ranked the following four risk factors: male sex (P < 0.001), habitual heavy drinking (P < 0.005), hepatitis B virus antibody positivity (anti-HBs and/or anti-HBc, P < 0.05), and age greater than 60 years (P < 0.05). The odds ratio of HCC was 4.20 (95% confidence interval; CI, 1.80-9.78) in male patients, 3.27 (95% CI, 1.46-7.30) in habitual heavy drinkers, 2.01 (95% CI, 1.01-3.99) in patients positive for hepatitis B virus antibodies, and 2.06 (95% CI, 1.00-4.23) in patients older than 60 years. The cumulative occurrence rates of HCC after blood transfusion were significantly higher in habitual heavy drinkers (4.8%, 49.4%, and 74.7% at 10, 20, and 30 years, respectively) than in non-drinkers (0%, 21.0%, and 23.3% at 10, 20, and 30 years, respectively, P < 0.0003). The mean interval for progression to LC after blood transfusion was significantly shorter in the habitual heavy drinkers than in the non-drinkers (22.4 +/- 4.4 years vs 28.4 +/- 3.9 years; P < 0.0003). This multivariate analysis revealed that habitual heavy drinking and hepatitis B virus antibody positivity are significant risk factors for HCC in HCV-related liver cirrhosis.  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