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.
We evaluated the anti-tumor effect of adenoviral vector-mediated p53 gene therapy on the growth of canine osteosarcoma xenografts formed in nude mice. Nude mice were subcutaneously transplanted with cells of 2 P53 mutant canine osteosarcoma cell lines, POS and CHOS. The osteosarcoma xenografts were injected with either an adenoviral vector that expresses canine wild-type P53 (AxCA-cp53) or LacZ (AxCA-LacZ). Tumor growth was significantly inhibited in the xenografts injected with AxCA-cp53 in comparison to those injected with AxCA-LacZ or PBS during the observation period of 27 days. An increase of the amount of p21(WAF1/CDKN1A) mRNA, and the number of apoptotic cells was shown in the tumors injected with AxCA-cp53 in comparison to those injected with AxCA-LacZ or PBS. The present study revealed that the adenoviral vector-mediated p53 gene transfer had an anti-tumor effect in canine osteosarcoma xenografts formed in nude mice.  相似文献   

2.
Introduction:  In the chemotherapy for treatment of lymphoid tumors in dogs, myelosuppression is a frequently encountered dose‐limiting factor. One possible approach to overcome myelosuppression is induction of chemoresistance in hematopoietic stem cells through expression of the mdr1 gene. A full‐length canine mdr1 cDNA clone was isolated in our laboratory. The present study was carried out to assess whether it confers multidrug resistance in canine cell lines.
Materials and methods:  The full‐length canine mdr1 cDNA was inserted into an expression plasmid vector. A canine mammary tumor cell line (CIPP) and osteosarcoma cell line (OOS) were transfected with the canine mdr1 expression vector. Expression of P‐gp was examined by immunoblotting. Function of ATP‐dependent drug efflux was measured by flow cytometric analysis using Rhodamine 123. Sensitivity to chemotherapeutic drugs was shown by estimation of 50% inhibitory concentrations (IC50) of vincristine or doxorubicin.
Results:  Immunoblotting of the transfected cells revealed a strong band of P‐gp detected by a monoclonal antibody directed to P‐gp. Rhodamine 123 efflux test showed an apparent drug efflux activity in the transfected cells. From the IC50 of the chemotherapeutic agents, the transfected cells showed a remarkable increase (20 to 60‐fold) in the resistance to vincristine or doxorubicin.
Conclusion:  Transfection of canine mdr1 gene induced P‐gp expression and strong drug resistance in canine cell lines.  相似文献   

3.
Introduction:  MCC, a cell wall composition prepared from Mycobacterium phlei ., inhibits the proliferation and induces apoptosis in a wide range of tumor cells. Bisphosphonates have been reported to inhibit the proliferation of canine osteosarcoma cell lines. In this study, we have determined the activity of MCC alone and in combination with the bisphosphonates alendronate and pamidronate on canine osteosarcoma cell lines.
Methods:  Canine osteosarcoma cell lines D17 and D22 were incubated with different concentrations of MCC (0.01–100 μg/ml) and suboptimal concentrations of alendronate and pamidronate for 72 hours. Cellular proliferation was measured by MTT reduction. Nuclear DNA condensation was determined using with Hoescht 33258 staining, and apoptosis by flow cytometry using active‐caspase‐3/PE and anti‐cleaved‐PARP/FITC antibodies.
Results:  MCC inhibited the proliferation of both canine osteosarcoma D17 and D22 cell lines in a concentration‐dependent manner. The IC50 for D17 cells was 3.9 μg/ml and for D22 cells 44.4 μg/ml. Cells incubated with 100 μg/ml MCC were positive for Hoescht staining, active caspase‐3 and cleaved PARP, indicative of cell death by apoptosis. The addition of alendronate or pamidronate was found to potentiate the apoptosis‐inducing activity of MCC. Maximal activity was observed when 5 μM alendronante or 10 μM pamidronate were used in combination with 100 μg/ml MCC.
Conclusion:  MCC inhibits the proliferation and induces apoptosis in canine osteosarcoma cell lines in vitro . This anticancer activity can be potentiated by the use of alendronate and pamidronate. These data support the development of MCC as a therapeutic agent for the treatment of canine osteosarcoma.  相似文献   

4.
Introduction:  Aberrant expression of the proto‐oncogene c‐Met has been noted in a variety of human cancers. In dogs, inappropriate Met expression has been identified in canine osteosarcoma (OSA) tumor samples. To better define the potential role of Met dysregulation in canine cancer, we cloned canine Met, HGF, and HGF activator and evaluated their expression patterns in a variety of canine tumor cell lines.
Methods:  Canine Met, HGF, and HGF activator were cloned from normal canine liver and canine OSA cell lines using primers based on regions of homology between mouse and human sequences as well as 5' and 3' RACE.
Results:  Inappropriate expression of Met was found in canine cell lines derived from OSAs, mast cell tumors, histiocytic sarcomas, hemangiosarcoma, and melanomas. Both HGF and HGF activator were found to be expressed in several of these tumor cell lines, providing evidence of a possible autocrine loop of Met stimulation. Incubation of canine tumor cell lines with rhHGF resulted in Met autophosphorylation and activation of the downstream signaling elements Gab1, Akt and Erk1/2. Scattering of tumor cells in response to HGF occurred under conditions of cell stress, such as serum starvation. Lastly, the Met inhibitor PHA‐665752 blocked HGF induced phosphorylation of canine Met and Gab1.
Conclusions:  These studies provide evidence that similar to the case in human tumors, aberrant Met expression may play an important role in the biology of canine cancer. As such, inhibition of Met function may represent a potentially useful novel therapeutic approach.  相似文献   

5.
Introduction:  Many dogs with lymphoid tumors develop resistance to chemotherapy. As a mechanism of drug resistance in canine lymphoma, ATP‐dependent drug efflux by P‐glycoprotein was reported, however, inhibition of apoptosis mediated by P53 inactivation has not been investigated. In this study, we investigated the relationship between p53 gene mutation and clinical drug resistance in canine lymphoid tumors.
Methods:  Tumor specimens were obtained from 44 dogs with lymphoid tumors. Mutations of p53 gene at exon 4–8 of these tumor tissues were examined by PCR‐SSCP (single strand conformational polymorphism) analysis, followed by nucleotide sequencing of the abnormal bands. The cases were treated with UW‐Madison protocol, and its response was evaluated by the tumor size or the number of peripheral leukemic cells.
Results:  Of the 44 dogs, 15 dogs (34%) had p53 mutation, whereas 29 dogs (66%) were devoid of p53 mutation, before or during the chemotherapeutic protocol. Rate of good response (CR and PR) to chemotherapy was significantly lower in the dogs with p53 mutation (20%) than those without p53 mutation (55%)(p = 0.022). Median overall survival duration after examination of p53 mutation was significantly shorter in dogs with p53 mutation (101days) than those without p53 mutation (223days)(p = 0.008).
Conclusions:  Lymphoid tumors with p53 mutations were shown to have worse prognosis than those without p53 mutation.  相似文献   

6.
Introduction:  There is a wealth of information available regarding tyrosine kinase (TK) expression in human cancer, however there is minimal information regarding the expression and function of TKs in canine melanoma, and no attempt has been made to systematically define the repertoire of TKs expressed. This study employed a molecular technique called RT‐PCR display to simultaneously evaluate the expression of up to 30 different TKs in a canine melanoma cell line.
Materials and Methods:  mRNA was extracted and reverse‐transcribed from the 17CM98 canine melanoma line, then subjected to PCR using degenerate primers coding for highly conserved regions which flank the kinase domains of many receptor and nonreceptor TKs. The resulting product was ligated into a plasmid vector and used to transform E. coli . Multiple colonies were isolated, and the cDNA inserts sequenced.
Results and Conclusions:  Sequencing 46 clones yielded canine homologs of IGF‐1R (50%), JAK1 (17%), PDGFR‐α(11%), FGFR1 (9%), Axl (7%), c‐Abl (4%), and PTK2 (2%). Interestingly, IGF‐1R, JAK1 and Axl were detected using a similar technique in human melanoma, supporting the cross‐species validity of this assay. Given the abundance of IGF‐1R clones, we sought to determine the biologic effect of rhIGF‐1 in 17CM98 cells. IGF‐1 stimulated IGF‐1R phosphorylation, cell proliferation and VEGF production in 17CM98, and protected the cells from serum starvation‐induced apoptosis. Expression of IGF‐1R mRNA was detected in 5 of 5 additional canine melanoma cell lines evaluated, suggesting that IGF‐1R expression may be common in canine melanoma cells and providing a novel target for future therapy.  相似文献   

7.
8.
Introduction:  Cyclooxygenase‐2 (Cox‐2) is the inducible form and the rate‐limiting enzyme, for conversion of arachidonic acid to prostaglandins. Cox‐2 overexpression, common in carcinomas, is associated with increased growth rate, resistance to apoptosis, angiogenesis, and overall, both local and distant aggressive behavior. Cox‐2 overexpression has been detected in human and canine mammary tumors (MTs). Histopathology of canine MT is not always predictive of biologic behavior, and anecdotally, only 50% of the malignant MTs are expected to metastasize. We hypothesize that Cox‐2 expression correlates with aggressive behavior.
Methods:  This retrospective study evaluated 48 bitches, presented for excision of MT between 2000 and 2003 at FMVZ de Botucatu‐UNESP, Brasil. Follow‐up varied from 18 months to 24 months and included physical examination and thoracic radiographs. Histopathologic examination was performed in all tumors, as well as in metastatic lesions when detected in the follow‐up period. Immunohistochemistry was used to detect expression of Cox‐2 in paraffin blocks (Rabbit polyclonal anti‐PGHS‐2. Oxford Biomedical). 10 adenomas, 10 carcinomas, 10 benign mixed tumors, 10 malignant mixed tumors and 8 cases of primary carcinomas and their metastatic lesions.
Results:  Expression of Cox‐2 varied among groups. Adenomas (32.1%), mixed benign tumors (38%), carcinomas (60.3%), malignant mixed tumors (65.8%), and metastatic carcinomas (81.25%) and their metastatic lesions (84.35%). Statistically significant differences (p < 0.05) were observed between the benign and malignant counterparts and between carcinomas and metastatic carcinomas.
Conclusions:  Cox‐2 expression correlates with both histologic and biologic behavior in mammary carcinomas, and may serve as a predictor of metastatic potential.  相似文献   

9.
The p53 gene is one of the important tumour suppressor genes that are involved with the cell survival signal pathway. One of the major functions of the p53 protein is to organize cell cycle regulation and induction of apoptosis for cellular genetic stability. It has been documented that more than 50% of all human cancers include a p53 mutation. We evaluated the difference in radiosensitivity between upregulating the expression of canine wild‐type p53 (cp53) in cultured osteosarcoma (D17) cells and naive D17 cells in vitro. We found that upregulating transfected cp53 D17 cells increased their radiation sensitivity in vitro, and there was a significant decrease (P < 0.009) in survival between cp53‐transfected D17 cells and naive D17 cells. In this experiment, a p53 enhancement ratio (p53ER) reached approximately 3.0 at high doses. The transfected cp53 D17 cells were significantly more radiosensitive at all doses evaluated than naive D17 cells, except at 1 Gy where too few data points were available. The p53ER increased rapidly at doses less than 4 Gy, achieving a maximum of about 3.0 for doses of 4 Gy and above. This study shows the enhanced radiosensitivity of the transfected p53 at clinically relevant doses.  相似文献   

10.
Introduction:  Mycobacterial cell wall‐DNA complex (MCC) is a bifunctional anticancer agent that induces cancer cell apoptosis and stimulates cytokine synthesis by immune cells. Intravesical MCC is currently being evaluated in humans with high‐grade urinary bladder cancer. Evaluation of MCC in dogs with transitional cell carcinoma (TCC) will allow mechanistic studies in a natural animal model of TCC, and a potentially beneficial therapy for dogs with this cancer. In this study, we have determined the anticancer activity of MCC against canine TCC cells in vitro .
Methods:  Canine TCC cells (K9TCC cell line) were incubated with MCC (0.05–100 μg/ml, 0.5–72 hours). Cellular proliferation was measured by MTT reduction. Cell cycle was analyzed by flow cytometry with propidium iodide. Apoptosis was identified by flow cytometry using anti‐active‐caspase‐3/PE and anti‐cleaved‐PARP/FITC antibodies. Apoptosis‐inducing activity of 100 μg/ml MCC in combination with piroxicam (0.1–1.0 uM) was evaluated.
Results:  MCC inhibited K9TCC cell proliferation in a concentration‐dependent manner (maximal activity – 45% at 100 μg/ml MCC) in association with the presence of activated caspase‐3 and cleaved PARP. Inhibition of proliferation and apoptosis‐inducing activities of MCC were independent of cell cycle phase. A thirty‐minute exposure of MCC was sufficient for optimal activity. Piroxicam (0.5 uM) enhanced apoptosis‐inducing activity of MCC.
Conclusions:  MCC induces apoptosis in K9TCC cells. This activity is potentiated by piroxicam. Following positive results in vitro , in vivo studies have been initiated. One dog, treated to date, has had a minor reduction in tumor volume following the first course of treatment with no treatment‐related toxicity.  相似文献   

11.
Introduction:  Cyclooxygenase‐2 (COX‐2) inhibitors are being used increasingly in cancer therapy. Although the effects of COX‐2 inhibitors have been evaluated extensively in carcinomas, less is known about their effects in sarcomas. Since the majority of dogs with appendicular osteosarcoma (OSA) are treated for pain with a non‐steroidal anti‐inflammatory drug (some COX‐2 selective) prior to definitive treatment, it is important to determine the effects that commonly used NSAIDS have on tumor cell growth.
Methods:  Established canine osteosarcoma (POS, HMPOS and COS31) and canine fibroblast cell lines were maintained in culture under standard conditions. Cells were incubated with either deracoxib (1 uM to 500 uM) or piroxicam (1 uM to 1000 uM). Cell viability was assessed at 72 hours by cell counts and the MTT assay. The DNA fragmentation analysis was utilized to assess for apoptosis induction.
Results:  Deracoxib concentrations ≥100 uM and piroxicam concentrations ≥500 uM significantly reduced mean cell viability of all three OSA cell lines (lowest cell viability percentages 20% and 32%, respectively). Deracoxib concentrations ≥250 uM and piroxicam concentrations ≥500 uM also reduced viability of fibroblasts; however, the cell viability percent was reduced to only 54% and 68%, respectively, of the control value. Exposure of OSA cells to cytotoxic concentrations of deracoxib and piroxicam did not result in DNA fragmentation.
Conclusions:  Deracoxib and piroxicam demonstrated a cytotoxic effect on canine osteosarcoma cells. There was no evidence of apoptosis induction at the concentrations evaluated. Further investigation will need to be performed to determine whether either drug exhibits anti‐tumor effects in vivo .  相似文献   

12.
Introduction:  Photodynamic therapy (PDT) involves the light activation of a drug within a tumor causing selective tumor cell death. Unfortunately, some photosensitizing drugs have been associated with adverse reactions in veterinary patients. Zinc phthalocyanine tetrasulfonate (ZnPcS4) is a promising second‐generation photosensitizer for use in veterinary medicine, however, it cannot be applied clinically until safety and efficacy data are available.
Methods:  Increasing intraperitoneal doses of ZnPcS4 were given to Swiss Webster mice to assess acute toxicity. Based on mouse toxicity data, a phase I clinical trial of ZnPcS4‐based PDT in tumor‐bearing dogs was designed, using an accelerated titration scheme starting at 0.5% of the minimum toxic dose in mice. 24‐hours after ZnPcS4 administration tumors were irradiated with 675 nm light and dogs were evaluated by routine hematology and serum biochemistry at regular intervals after PDT.
Results:  Doses >125 mg/kg were associated with acute toxicity and mortality in Swiss Webster mice, suggesting the minimum toxic dose is 120–125 mg/kg. One dog, a Golden retriever with a massive malignant fibrous histiocytoma, has been entered into the phase I clinical trial. No deleterious effects were noted after ZnPcS4 administration. Within 48 hours of PDT, the tumor was dark and necrotic, with no grossly visible changes to the surrounding normal tissues. Histological examination of the PDT‐treated tumor confirmed widespread necrosis and thrombosis consistent with PDT‐mediated damage. The owner reported no adverse effects after treatment.
Conclusions:  Although preliminary data are encouraging, additional evaluation of ZnPcS4‐based PDT is required to determine its role in veterinary medicine.  相似文献   

13.
Introduction:  Over‐expression of COX‐2 has been observed in several human and animal malignancies and is implicated in carcinogenesis through the conversion of arachidonic acid to PGE‐2. Use of platinum‐containing cytostatic agents and/or (non‐)specific COX‐2 inhibitors, has been reported as a treatment option for canine oral non‐tonsillar squamous cell carcinomas (ONT‐SCC). However, no study describes the effect of a combination of carboplatin and piroxicam on this tumor type.
Methods:  7 dogs with a T3 (WHO‐TNM) ONT‐SCC were treated with piroxicam and carboplatin. Five had bone involvement and no detectable metastasis. Two dogs without bone involvement had metastasis in the regional lymph nodes. Piroxicam was given orally 0.3 mg/kg s.i.d. Each dog was scheduled to receive between 6 and 12 carboplatin infusions (300 mg/m2 i.v.) at 3 week intervals. Ondansetron and metoclopramide were used as anti‐emetic agents. The dogs are planned to receive piroxicam on a lifelong basis.
Results:  Complete response (CR) without adjuvant surgery was achieved in 4 of the 7 dogs. Two dogs needed adjuvant surgery to achieve CR. One dog had progressive disease and was euthanised 231 days after start of therapy. All the others were still alive and in CR at date of analysis. Median follow‐up was 335 days (107–689 days).
Conclusions:  Our study suggests that a combination of piroxicam and carboplatin is a useful treatment option for canine ONT‐SCC. All dogs tolerated therapy well and the 57% response rate for reaching a complete and durable remission without adjuvant surgery is promising.  相似文献   

14.
Introduction:  Lymphoma is one of the most common cancers in dogs and while clinical remission can be induced using chemotherapy, very few dogs are cured. Since cytokine‐adjuvanted autologous whole‐tumor‐cell vaccines (ATCV) can induce potent antitumor immune responses against otherwise non‐immunogenic cancers we initiated a study of such an approach in dogs with lymphoma.
Methods:  Following achievement of a complete remission using a 19‐week CHOP‐based chemotherapy protocol, 51 dogs with B‐cell lymphoma were randomized to receive 8 treatments (4 weekly, then 4 additional at q2wk intervals) of vaccine or lipid‐equivalent placebo. Dogs were followed monthly for assessment of remission duration and survival. Surrogate indices of immune response (delayed‐type hypersensitivity, interferon‐γ quantitative RT‐PCR, lymphocyte proliferation, and flow cytometry for lymphoma‐specific antibodies) were also investigated before and after vaccination.
Results:  No significant difference in median remission duration was observed between dogs receiving vaccine (277 days) or placebo (258 days); the Kaplan‐Meier curves were virtually super‐imposable. No significant differences in surrogate indices of immune response were noted pre‐ and post‐vaccination.
Conclusions:  In this context, an hGM‐CSF DNA‐cationic lipid complexed ATCV vaccine did not enhance remission duration in dogs with lymphoma, likely due to lack of vaccine‐induced tumor‐specific immunity.  相似文献   

15.
Introduction:  MOPP chemotherapy is useful for relapsed canine lymphoma. This study evaluates the efficacy of this protocol after substitution of CCNU (lomustine) or BiCNU (carmustine) for mechlorethamine (C/B‐OPP).
Methods:  Patient signalment, response to chemotherapy, toxicity and survival data were abstracted from medical records of dogs from receiving C/B‐OPP between 1998 and 2004.
Results:  Fifty‐eight dogs received C/B‐OPP rescue chemotherapy during the study period. The median remission duration after initial chemotherapy, consisting of CHOP‐based therapy in 91% of dogs, was 133 days (range, 10 to 932 days). Thirty‐eight of fifty‐eight dogs (66%) responded to C/B‐OPP rescue after relapse (22 CR, 16 PR), for a median of 48 days (range, 2 to 359 days). Overall, C/B‐OPP extended survival by a median of 90 days (range, 2 to 426 days). Twenty‐four dogs (41%) experienced one or more episodes of Grade II or higher gastrointestinal toxicity. Forty‐one dogs (71%) experienced one or more episodes of Grade II or higher hematologic toxicity. Twelve dogs (20%) developed regenerative anemia with diarrhea consistent with gastrointestinal hemorrhage. Treatment delays due to hematologic toxicity occurred in 37 dogs (63%). There were 16 nonfatal treatment‐related episodes of sepsis requiring hospitalization. 5 dogs died due to sepsis and/or chemotherapy‐related complications.
Conclusions:  C/B‐OPP chemotherapy has activity against relapsed canine lymphoma which is similar to that of traditional MOPP rescue therapy. Moderate to severe hematologic toxicity was observed. Further work is warranted to optimize drug doses and scheduling.  相似文献   

16.
Introduction:  Greater than 50% of dogs with thyroid tumors present with surgically unresectable disease for which external beam radiotherapy has been reported to prolong survival. The success of 131I for control of thyroid tumors in cats and in humans suggests such therapy may also play a role in the management of canine thyroid cancer.
Methods:  Thirty‐nine dogs with WHO stage II/III (invasive or ectopic; n = 32) or IV (metastatic; n = 7) thyroid tumors were treated with 131I alone. Changes in thyroid function, 99MTc‐pertechnetate (99MTc) scintigraphic changes, and tumor response were recorded. Dogs with ventral cervical tumors were evaluated for feasibility of surgical resection following 131I.
Results:  Median overall survival was 839 days and 366 days for dogs with stage II/III and stage IV tumors, respectively. Thyroid hormone status, site and surgical resection were not associated with outcome in dogs with stage II/III tumors. Three dogs developed severe bone marrow suppression.
Conclusions:  These findings suggest 131I should be investigated more thoroughly in dogs with thyroid tumors not considered surgical candidates to more clearly characterize the indications for therapy and followup recommendations. 131I dosimetry in dogs with thyroid tumors remains problematic. Administration of 131I is currently based on empiric recommendations and, in general, the treatment is well tolerated although additional studies are indicated to optimize response and minimize toxicity.  相似文献   

17.
Introduction:  Inverse associations have been shown between E‐cadherin expression and degree of differentiation in canine mammary tumours and with the acquisition of malignancy in human epithelial cancers. The purpose of this study was to examine whether there was an association between prognosis, and the distribution or intensity of immunohistochemical staining for E‐cadherin in anal sac adenocarcinoma.
Methods:  Formalin‐fixed paraffin‐wax embedded canine anal sac adenocarcinoma specimens were obtained for 36 patients for whom clinical data, including survival statistics, were available. Canine mammary adenoma tissue was employed as the positive and negative controls with antibody diluent replacing the primary antibody in the latter. Samples were scored according to proportion of cells showing positive cytoplasmic membranous staining, intensity of cytoplasmic membranous staining and proportion of nuclei staining. As quality control, a sample of 10 slides was randomly selected for repeat evaluation; results exactly matched those obtained in the first assessment.
Results:  All anal sac adenocarcinoma cells expressed E‐cadherin to varying degrees. There was no evidence of an association between survival time and the intensity of staining or the proportion of nuclei staining. Patients whose tumours exhibited >75% positive cytoplasmic membranous staining survived significantly longer than those exhibiting <75% staining with median survival times of 1100 and 479 days respectively (log rank test: χ2 = 3.89, 1 DF, p = 0.049).
Conclusion:  Immunohistochemical evaluation of the proportion of cells in anal sac adenocarcinoma samples exhibiting positive cytoplasmic membranous staining for E‐cadherin may allow improved determination of prognoses and could aid clinical decision making.  相似文献   

18.
Canine osteosarcoma (OS) has been used as a model system for the study of cancer biology and treatment despite the lack of information regarding its pathogenesis. Expression of tumor suppressor genes known to participate in malignant transformation were studied in five different OS cell lines. Each of the cell lines exhibited properties of transformed cells, and those that were tested grew in soft agarose and formed osteoid-containing tumors when injected subcutaneously into nude mice. p53 function was determined to be defective in each cell line as indicated by the lack of induction of p53-responsive genes, p21 and mdm2, following treatment with 5-fluorouracil. p53 mRNA and protein levels were elevated in three cell lines and were extremely low in two cell lines. p53 protein overexpression correlated with the presence of mutations within the DNA binding domain. Four cell lines appeared to contain normal retinoblastoma (Rb) mRNA and Rb protein and no detectable p16 mRNA or protein. In contrast, the remaining cell line contained high levels of p16 mRNA and protein and significantly reduced levels of Rb, p107, and p130 proteins. These results underscore the importance of inactivating p53 and Rb family pathways in canine OS and suggest that unlike human OS, cells derived from canine OS contain mutations that simultaneously inactivate all three Rb family members.  相似文献   

19.
20.
Introduction:  The most common neoplasms located in the anterior mediastinum in the canine are thymoma and lymphoma. Distinguishing between the two is a diagnostic challenge. Treatment and prognosis for these diseases differs significantly. Thymomas contain a population of normally developing T cells. The majority of these T cells exhibit an immature phenotype, characterized by co‐expression of CD4 and CD8. This phenotype is rarely seen on neoplastic lymphocytes. The purpose of this study was to determine if analysis by flow cytometry could discriminate thymoma from lymphoma based on these cell surface markers.
Methods:  Fine needle aspirates were obtained from ten canine patients with mediastinal masses. Cells were analyzed by flow cytometry using a panel of T and B cell markers.
Results:  Six cases with 10% or greater CD4 + CD8+ cells were diagnosed with thymoma and were confirmed by histopathology. Four cases had fewer than 5% CD4 + CD8+ cells, having lymphocytes expressing CD4 only (3 cases) or CD21, a B cell marker(1 case). These were confirmed as lymphoma by cytology and/or a clonality assay. The sensitivity and specificity of this assay when used in the diagnostic work‐up for suspected thymoma was 100%.
Conclusion:  Flow cytometry may provide important, complementary information in the diagnostic work‐up of the canine patient with a mediastinal mass.  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