Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-29T13:59:41.147Z Has data issue: false hasContentIssue false

Section 1 - Antibodies

Published online by Cambridge University Press:  27 October 2016

Runjan Chetty
Affiliation:
University of Toronto
Kumarasen Cooper
Affiliation:
University of Pennsylvania
Allen M. Gown
Affiliation:
Phenopath Laboratories, Seattle
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2016

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Selected references

Chedid, A, Chejfec, G, Eichorst, M, et al. Antigenic markers for hepatocellular carcinoma. Cancer 1990; 65: 84–7.3.0.CO;2-D>CrossRefGoogle ScholarPubMed
Fucich, LF, Cheles, MK, Thung, SN, et al. Primary versus metastatic hepatic carcinoma. An immunohistochemical study of 34 cases. Archives of Pathology and Laboratory Medicine 1994; 118: 927–30.Google Scholar
Guindi, M. Yazdi, HM, Gilliatt, MA. Fine needle aspiration biopsy of hepatocellular carcinoma. Value of immunocytochemical and ultrastructural studies. Acta Cytologica 1994; 38: 385–91.Google ScholarPubMed
Leong, AS-Y, Sormunen, RT, Tsui, WM-S, Liew, CT. Immunostaining for liver cancers. Histopathology 1998; 33: 318–24.Google Scholar

Selected references

Banerjee, SS, Harris, M. Morphological and immunophenotypic variations in malignant melanoma. Histopathology 2000; 36: 387402.CrossRefGoogle ScholarPubMed
Kinner, B, Spector, M. Smooth muscle actin expression by human articular chondrocytes and their contraction of a collagen–glycoasaminoglycan matrix in vitro. Journal of Orthopedic Research 2001; 19: 233–41.CrossRefGoogle ScholarPubMed
Kung, IT, Thallas, V, Spencer, EJ, Wilson, SM. Expression of muscle actins in diffuse mesothelioma. Human Pathology 1995; 26: 565–70.CrossRefGoogle Scholar
Kutzner, H, Mentzel, T, Kaddu, S, et al. Cutaneous myoepithelioma: an under-recognised cutaneous neoplasm composed of myoepithelial cells. American Journal of Surgical Pathology 2001; 25: 348–55.CrossRefGoogle Scholar
Leong, AS-Y, Milios, J, Leong, FJ. Patterns of basal lamina immunostaining in soft-tissue and bony tumors. Applied Immunohistochemistry 1997; 5: 17.CrossRefGoogle Scholar
Ohta, K, Mortenson, RL, Clark, RA, et al. Immunohistochemical identification and characterization of smooth muscle-like cells in idiopathic pulmonary fibrosis. American Journal of Respiratory & Critical Care Medicine 1995; 152: 1659–65.CrossRefGoogle ScholarPubMed
Raymond, WA, Leong, AS-Y. Assessment of invasion in breast lesions using antibodies to basement membrane component and myoepithelial cells. Pathology 1991; 23: 291–7.CrossRefGoogle ScholarPubMed
Santini, D, Ceccarelli, C, Leone, O, et al. Smooth muscle differentiation in normal human ovaries, ovarian stromal hyperplasia and ovarian granulosa-stromal cell tumors. Modern Pathology 1995; 8: 2530.Google Scholar

Selected references

Dabir, PD, Ottosen, P, Hoyer, S, Hamilton-Dutoit, S. Comparative analysis of three- and two-antibody cocktails to AMACR and basal cell markers for the immunohistochemical diagnosis of prostate carcinoma. Diagnostic Pathology 2012; 7: 81.CrossRefGoogle ScholarPubMed
Hameed, O, Humphrey, PA. p63/AMACR antibody cocktail restaining of prostate needle biopsy tissues after transfer to charged slides: a viable approach in the diagnosis of small atypical foci that are lost on block sectioning. American Journal of Clinical Pathology 2005; 124: 708–15.CrossRefGoogle ScholarPubMed
Jiang, Z, Woda, BA, Rock, KL, et al. P504S: a new molecular marker for the detection of prostate carcinoma. American Journal of Surgical Pathology 2001; 25: 1397–404.CrossRefGoogle ScholarPubMed
Rubin, MA, Zhou, M, Dhanasekaran, SM, et al. alpha-Methylacyl coenzyme A racemase as a tissue biomarker for prostate cancer. JAMA 2002; 287: 1662–70.CrossRefGoogle ScholarPubMed
Shi, XY, Bhagwandeen, B, Leong, AS. p16, cyclin D1, Ki-67, and AMACR as markers for dysplasia in Barrett esophagus. Applied Immunohistochemistry and Molecular Morphology 2008; 16: 447–52.CrossRefGoogle ScholarPubMed
Zhou, M, Aydin, H, Kanane, H, Epstein, JI. How often does alpha-methylacyl-CoA-racemase contribute to resolving an atypical diagnosis on prostate needle biopsy beyond that provided by basal cell markers? American Journal of Surgical Pathology 2004; 28: 239–43.CrossRefGoogle ScholarPubMed
Zhou, M, Chinnaiyan, AM, Kleer, CG, Lucas, PC, Rubin, MA. Alpha-Methylacyl-CoA racemase: a novel tumor marker over-expressed in several human cancers and their precursor lesions. American Journal of Surgical Pathology 2002; 26: 926–31.CrossRefGoogle ScholarPubMed

Selected references

Iwamoto, N, Nishiyama, E, Ohwada, J, Arai, H. Distribution of amyloid deposits in the cerebral white matter of the Alzheimer's disease brain: relationship to blood vessels. Acta Neuropathologica (Berlin) 1997; 93: 334–40.CrossRefGoogle ScholarPubMed
Jacobson, DR, Pastore, RD, Yaghoubian, R, et al. Variant-sequence transthyretin (isoleucine 122) in late-onset cardiac amyloidosis in Black Americans. New England Journal of Medicine 1997; 336: 466–73.CrossRefGoogle ScholarPubMed
Lansbury, PT. In pursuit of the molecular structure of amyloid plaque: new technology provides unexpected and critical information. Biochemistry 1992; 31: 6865–70.CrossRefGoogle ScholarPubMed
Ravid, M, Shapiro, J, Lang, R, et al. Prolonged dimethylsulphoxide treatment in 13 patients with systemic amyloidosis. Annals of Rheumatic Diseases 1982; 41: 587–92.CrossRefGoogle ScholarPubMed
Sherriff, FE, Bridges, LR, Sivaloganathan, S. Early detection of axonal injury after human head trauma using immunocytochemistry for beta-amyloid precursor protein. Acta Neuropathologica (Berlin) 1994; 87: 5562.CrossRefGoogle ScholarPubMed
Shimizu, M, Manabe, T, Matsumoto, T, et al. β2 Microglobulin haemodialysis related amyloidosis: distinctive gross features of gastrointestinal involvement. Journal of Clinical Pathology 1997; 50: 873–5.CrossRefGoogle Scholar
Sousa, MM, Cardoso, I, Fernandes, R, et al. Deposition of transthyretin in early stages of familial amyloidotic polyneuropathy: evidence for toxicity of non fibrillar aggregates. American Journal of Pathology 2001; 159: 19932000.CrossRefGoogle Scholar

Selected references

Cessna, MH, Zhou, H, Sanger, WG, et al. Expression of ALK-1 and p80 in inflammatory myofibroblastic tumor and its mesenchymal mimics: a study of 135 cases. Modern Pathology 2002; 15: 931–8.CrossRefGoogle Scholar
Chan, JK.Newly available antibodies with practical applications in surgical pathology. International Journal of Surgical Pathology 2013; 21: 553–72.CrossRefGoogle ScholarPubMed
Conklin, CM, Craddock, KJ, Have, C, et al. Immunohistochemistry is a reliable screening tool for identification of ALK rearrangement in non-small-cell lung carcinoma and is antibody dependent. Journal of Thoracic Oncology 2013; 8: 4551.CrossRefGoogle ScholarPubMed
Corao, DA, Biegel, JA, Coffin, CM, et al. ALK expression in rhabdomyosarcomas: correlation with histologic subtype and fusion status. Pediatric and Developmental Pathology 2009; 12: 275–83.CrossRefGoogle ScholarPubMed
O'Bryant, CL, Wenger, SD, Kim, M, Thompson, LA. Crizotinib: a new treatment option for ALK-positive non-small-cell lung cancer. Annals of Pharmacotherapy 2013; 47: 189–97.CrossRefGoogle ScholarPubMed
Perez-Pinera, P, Chang, Y, Astudillo, A, Mortimer, J, Deuel, TF. Anaplastic lymphoma kinase is expressed in different subtypes of human breast cancer. Biochemical and Biophysical Research Communications 2007; 358: 399403.CrossRefGoogle ScholarPubMed
Tennstedt, P, Strobel, G, Bölch, C, et al. Patterns of ALK expression in different human cancer types. Journal of Clinical Pathology 2014; 67: 477–81.CrossRefGoogle ScholarPubMed
Wasik, MA. Expression of anaplastic lymphomas kinase in non-Hodgkin's lymphomas and other malignant neoplasms: biological, diagnostic, and clinical implications. American Journal of Clinical Pathology 2002; 118 (Suppl 1): S8192.Google ScholarPubMed

Selected references

Carroll, RS, Zhang, J, Dashmner, K, et al. Androgen receptor expression in meningiomas. Journal of Neurosurgery 1995; 82: 453–60.CrossRefGoogle ScholarPubMed
Hakimi, JM, Rondinelli, RH, Schoenberg, MP, Barrack, ER. Androgen-receptor gene structure and function in prostate cancer. World Journal of Urology 1996; 14: 329–37.CrossRefGoogle ScholarPubMed
Hoang, MP, Callender, DL, Sola Gallego, JJ, et al. Molecular and biomarker analyses of salivary duct carcinomas: comparison with mammary duct carcinoma. International Journal of Oncology 2001; 19: 865–71.Google ScholarPubMed
Mertens, HJ, Heineman, MJ, Koudstaal, J, et al. Androgen receptor content in human endometrium. European Journal of Obstetrics, Gynecology and Reproductive Biology 1996; 70: 1113.CrossRefGoogle ScholarPubMed
Moriki, T, Ueta, S, Takashi, T, et al. Salivary duct carcinoma: cytologic characteristics and application of androgen receptor immunostaining for diagnosis. Cancer 2001; 93: 344–50.CrossRefGoogle ScholarPubMed
Syed, S, Martin, AM, Haupt, H, Podolski, V, Brooks, JJ. Frequent detection of androgen receptors in spindle cell lipomas: an explanation of this lesion's male predominance? Archives of Pathology and Laboratory Medicine 2008; 132: 81–3.CrossRefGoogle ScholarPubMed
Tihan, T, Harmon, JW, Wan, X, et al. Evidence of androgen receptor expression in squamous and adenocarcinoma of the esophagus. Anticancer Research 2001; 21: 3107–14.Google ScholarPubMed
Zhuang, YH, Blauer, M, Tammela, T, Tuohimaa, P. Immunodetection of androgen receptor in human urinary bladder cancer. Histopathology 1997; 30: 556–62.CrossRefGoogle ScholarPubMed

Selected references

Arends, MJ, Wyllie, AH. Apoptosis: mechanism and roles in pathology. International Reviews in Experimental Pathology 1991; 32: 223–54.CrossRefGoogle Scholar
Bukholm, IR, Bukholm, G, Nesland, JM. Reduced expression of both Bax and Bcl-2 is independently associated with lymph node metastasis in human breast carcinomas. APMIS 2002; 110: 214–20.CrossRefGoogle ScholarPubMed
Evans, JD, Cornford, PA, Dodson, A, et al. Detailed tissue expression of bcl-2, bax, bak, and bcl-x in the normal human pancreas and in chronic pancreatitis, ampullary and pancreatic ductal adenocarcinomas. Pancreatology 2001; 1: 254–62.CrossRefGoogle ScholarPubMed
Kerr, JFR, Wyllie, AH, Currie, AR. Apoptosis: a basic biological phenomenon with wide-ranging implications in tissue kinetics. British Journal of Cancer 1972; 26: 239–57.CrossRefGoogle ScholarPubMed
Krajewska, M, Moss, SF, Krajewski, S, et al. Elevated expression of Bcl-x and reduced Bak in primary colorectal adenocarcinomas. Cancer Research 1996; 56: 2422–7.Google Scholar
Kuwashima, Y, Uehara, T, Kishi, K, et al. Proliferative and apoptotic status in endometrial adenocarcinoma. International Journal of Gynecological Pathology 1995; 14: 45–9.CrossRefGoogle ScholarPubMed
Morgner, A, Sutton, P, O'Rourke, JL, et al. Helicobacter-induced expression of Bcl-X(L) in B lymphocytes in the mouse model: a possible step in the development of gastric mucosa-associated lymphoid tissue (MALT) lymphoma. International Journal of Cancer 2001; 92: 634–40.3.0.CO;2-V>CrossRefGoogle Scholar
Nagata, S, Golstein, P. The Fas death factor. Science 1995; 267: 1445–9.CrossRefGoogle ScholarPubMed
Raynal, P, Pollard, HB. Annexins. The problem of assessing the biological role for a gene family of multifunctional calcium and phospholipid-binding proteins. Journal of Biological Chemistry 1994; 265: 4923–8.Google Scholar
Sarkiss, M, Hsu, B, El-Naggar, AK, McDonnell, TJ. The clinical relevance and assessment of apoptotic cell death. Advances in Anatomic Pathology 1996; 3: 205–11.CrossRefGoogle Scholar
Schelwies, K, Sturm, I, Grabowski, P, et al. Analysis of p53/BAX in primary colorectal carcinoma: low BAX protein expression is a negative prognostic factor in UICC stage III tumors. International Journal of Cancer 2002; 99: 589–96.CrossRefGoogle ScholarPubMed
Sjostrom, J, Blomqvist, C, von Boguslawski, K, et al. The predictive value of bcl-2, bax, bcl-xL, bag–1, fas, and fasL for chemotherapy response in advanced breast cancer. Clinical Cancer Research 2002; 8: 811–16.Google ScholarPubMed
Trask, DK, Wolf, GT, Bradford, CR, et al. Expression of Bcl-2 family proteins in advanced laryngeal squamous cell carcinoma: correlation with response to chemotherapy and organ preservation. Laryngoscope 2002; 112: 638–44.CrossRefGoogle ScholarPubMed
Wyllie, AH, Kerr, JFR, Currie, AR. Cell death: the significance of apoptosis. International Reviews in Cytology 1980; 68: 251306.CrossRefGoogle ScholarPubMed

Selected references

Fatima, N, Cohen, C, Siddiqui, MT. Arginase-1: a highly specific marker separating pancreatic adenocarcinoma from hepatocellular carcinoma. Acta Cytologica 2014; 58: 83–8.CrossRefGoogle ScholarPubMed
Radwan, NA, Ahmed, NS. The diagnostic value of arginase-1 immunostaining in differentiating hepatocellular carcinoma from metastatic carcinoma and cholangiocarcinoma as compared to HepPar-1. Diagnostic Pathology 2012; 7: 149.CrossRefGoogle ScholarPubMed
Yan, BC, Gong, C, Song, J, et al. Arginase-1: a new immunohistochemical marker of hepatocytes and hepatocellular neoplasms. American Journal of Surgical Pathology 2010; 34: 1147–54.CrossRefGoogle ScholarPubMed

Selected references

Bakhshi, A, Jensen, JP, Goldman, P, et al. Cloning the chromosomal breakpoint of t(14;18) human lymphomas: clustering around Jh on chromosome 14 near a transcriptional unit on 18. Cell 1985; 41: 899906.CrossRefGoogle Scholar
Chetty, R, Dada, MA, Gatter, KC. Bcl-2: longevity personified. Advances in Anatomic Pathology 1997; 4: 134–8.CrossRefGoogle Scholar

Selected references

Onizuka, T, Moriyama, M, Yamochi, T, et al. BCL-6 gene product, a 92- to 98-kDa nuclear phosphoprotein, is highly expressed in germinal center B cells and their neoplastic counterparts. Blood 1995; 86: 2837.CrossRefGoogle Scholar
Ye, BH, Rao, PH, Chaganti, RS, Dalla-Favera, R. Cloning of BCL-6, the locus involved in chromosome translocations affecting band 3q27 in B-cell lymphoma. Cancer Research 1993; 53: 2732–5.Google ScholarPubMed

Selected reference

Latza, U, Niedobitek, G, Schwarting, R, et al. Ber-EP4: New monoclonal antibody which distinguishes epithelia from mesothelia. Journal of Clinical Pathology 1990; 43: 213–19.CrossRefGoogle Scholar

Selected reference

Bellisario, R, Carlsen, RB, Bahl, OP. Human chorionic gonadotropin. Linear amino acid sequence of the α subunit. Journal of Biological Chemistry 1973; 248: 6796–809.Google ScholarPubMed

Selected references

Advani, AS, Lim, K, Gibson, S, et al., OCT-2 expression and OCT-2/BOB.1 co-expression predict prognosis in patients with newly diagnosed acute myeloid leukemia. Leukemia and Lymphoma 2010; 51: 606–12.CrossRefGoogle ScholarPubMed
Kim, U, Qin, XF, Gong, S, et al. The B-cell-specific transcription coactivator OCA-B/OBF-1/Bob-1 is essential for normal production of immunoglobulin isotypes. Nature 1996; 383: 542–7.CrossRefGoogle ScholarPubMed

Selected references

Barresi, V, Vitarelli, E, Branca, G, et al. Expression of brachyury in hemangioblastoma: potential use in differential diagnosis. American Journal of Surgical Pathology 2012; 36: 1052–57.CrossRefGoogle ScholarPubMed
Lauer, SR, Edgar, MA, Gardner, JM, Sebastian, A, Weiss, SW. Soft tissue chordomas: a clinicopathologic analysis of 11 cases. American Journal of Surgical Pathology 2013; 37: 719–26.CrossRefGoogle ScholarPubMed
Sangoi, AR, Dulai, MS, Beck, AH, Brat, DJ, Vogel, H. Distinguishing chordoid meningiomas from their histologic mimics: an immunohistochemical evaluation. American Journal of Surgical Pathology 2009; 33: 669–81.CrossRefGoogle ScholarPubMed

Selected references

Bast, RC, Feeney, M, Lazarus, H, et al. Reactivity of a monoclonal antibody with human ovarian carcinoma. Journal of Clinical Investigation 1981; 68: 1331–7.CrossRefGoogle ScholarPubMed
Davis, AM, Zurawski, VR, Bast, RC, Klug, TL. Characterization of the CA 125 antigen associated with human epithelial ovarian carcinomas. Cancer Research 1986; 46: 6143–8.Google ScholarPubMed
Kuzuya, K, Nozaki, M, Chihara, T. Evaluation of CA 125 as a circulating tumor marker for ovarian cancer. Acta Obstetric et Gynecologica Japan. 1986; 38: 949–57.Google Scholar
O'Brien, TJ, Raymond, LM, Bannon, GA, et al. New monoclonal antibodies identify the glycoprotein carrying the CA 125 epitope. American Journal of Obstetrics and Gynecology 1991; 165: 1857–64.CrossRefGoogle ScholarPubMed

Selected references

Hedrick, L, Cho, KR, Vogelstein, B. Cell adhesion molecules as tumor suppressors. Trends in Cell Biology 1993; 3: 36–9.CrossRefGoogle Scholar
Madhavan, M, Srinivas, P, Abraham, E, et al. Cadherins as predictive markers of nodal metastasis in breast cancer. Modern Pathology 2001; 14: 423–7.CrossRefGoogle ScholarPubMed
Matsuura, K, Kawanishi, J, Jujii, S, et al. Altered expression of E-cadherin in gastric cancer tissues and carcinomatous fluid. British Journal of Cancer 1992; 66: 1122–30.CrossRefGoogle ScholarPubMed

Selected references

Al-Ahmadie, HA, Alden, D, Qin, LX, et al. Carbonic anhydrase IX expression in clear cell renal cell carcinoma. An immunohistochemical study comparing 2 antibodies. American Journal of Surgical Pathology 2008; 32: 377–82.CrossRefGoogle ScholarPubMed
Bing, Z, Lal, P, Lu, S, et al. Role of carbonic anhydrase IX, α-methylacyl coenzyme a racemase, cytokeratin 7, and galectin-3 in the evaluation of renal neoplasms: a tissue microarray immunohistochemical study. Annals of Diagnostic Pathology 2013; 17: 5862.CrossRefGoogle ScholarPubMed
Bui, MH, Visapaa, H, Seligson, D, et al. Prognostic value of carbonic anhydrase IX and KI67 as predictors of survival for renal clear cell carcinoma. Journal of Urology 2004; 171: 2461–6.Google ScholarPubMed
Genega, E, Ghebremichael, M, Najarian, R, et al. Carbonic anhydrase IX in renal neoplasms. Correlation with tumor type and grade. American Journal of Clinical Pathology 2010; 134: 873–9.CrossRefGoogle ScholarPubMed
Stillebroer, AB, Mulders, PFA, Boerman, OC, et al. Carbonic anhydrase IX in renal cell carcinoma: implications for prognosis, diagnosis and therapy. European Urology 2010; 58: 7583.CrossRefGoogle ScholarPubMed

Selected references

Allison, J, Hall, L, Maclntyre, I, Craig, RK. The construction and partial characterisation of plasmids containing complementary DNA sequences to human calcitonin precursor polyprotein. Biochemistry Journal 1981; 199: 725–31.CrossRefGoogle ScholarPubMed
Saad, MF, Ordonez, NG, Guido, JJ, Samaan, NA. The prognostic value of calcitonin immunostaining in medullary carcinoma of the thyroid. Journal of Clinical Endocrinology and Metabolism 1984; 59: 850–6.Google ScholarPubMed

Selected reference

Hisaoka, M, Wei-Qi, S, Jian, W, et al. Specific but variable expression of h-caldesmon in leiomyosarcomas. An immunohistochemical reassessment of a novel myogenic marker. Applied Immunohistochemistry and Molecular Morphology 2001; 9: 302–8.CrossRefGoogle ScholarPubMed

Selected references

Miettinen, MM, Sarloma-Rikala, M, Kovatich, AJ, Lasota, J. Calponin and h-caldesmon in soft tissue tumors: consistent h-caldesmon immunoreactivity in gastrointestinal stromal tumors indicates traits of smooth muscle differentiation. Modern Pathology 1999; 12: 756–62.Google ScholarPubMed
Savera, AT, Gown, AM, Zarbo, RJ. Immunolocalization of three novel smooth muscle-specific proteins in salivary gland pleomorphic adenoma: assessment of the morphogenetic role of myoepithelium. Modern Pathology 1997; 10: 1093–100Google ScholarPubMed

Selected references

Andrici, J, Farzin, M, Clarkson, A, et al. Mutation specific immunohistochemistry is highly specific for the presence of calreticulin mutations in myeloproliferative neoplasms. Pathology 2016; 48: 319–24.CrossRefGoogle ScholarPubMed
Kabbage, M, Trimeche, M, Bergaoui, S, et al. Calreticulin expression in infiltrating ductal breast carcinomas: relationships with disease progression and humoral immune responses. Tumour Biology 2013; 34: 1177–88.CrossRefGoogle ScholarPubMed

Selected references

Barberis, MCP, Faleri, M, Veronese, S, et al. Calretinin: a selective marker of normal and neoplastic mesothelial cells in serous effusions. Acta Cytologica 1997; 41: 1757–61.CrossRefGoogle ScholarPubMed
Cao, QJ, Jones, JG, Li, M. Expression of calretinin in human ovary, testis, and ovarian sex cord–stromal tumors. International Journal of Gynecologic Pathology 2001; 20: 346–52.CrossRefGoogle ScholarPubMed
Dei Tos, AP, Doglioni, C. Calretinin: a novel tool for diagnostic immunohistochemistry. Advances in Anatomic Pathology 1998; 5: 61–6CrossRefGoogle ScholarPubMed

Selected references

Monteagudo, C, Marcilla, A, Mormeneo, S, Llombart-Bosch, A, Sentandreu, R. Specific immunohistochemical identification of Candida albicans in paraffin-embedded tissue with a new monoclonal antibody (1B12). American Journal of Clinical Pathology 1995; 103: 130–5.CrossRefGoogle ScholarPubMed
Williams, DW, Jones, HS, Allison, RT, Potts, AJ, Lewis, MA. Immunocytochemical detection of Candida albicans in formalin fixed, paraffin embedded material. Journal of Clinical Pathology 1998; 51: 857–9.CrossRefGoogle ScholarPubMed

Selected reference

Robb, JA. Mesothelioma versus adenocarcinoma: false-positive CEA and Leu-M1 staining due to hyaluronic acid. Human Pathology 1989; 20: 400.CrossRefGoogle ScholarPubMed

Selected references

Birchmeier, W, Behrens, J. Cadherin expression in carcinomas: role in the formation of cell junctions and the prevention of invasiveness. Biochemia Biophysiology Acta 1994; 1198: 1126.Google ScholarPubMed
Hinck, L, Nathke, IS, Papkoff, J, Nelson, WJ. Dynamics of cadherin/catenin complex formation: novel protein interactions and pathways of complex assembly. Journal of Cell Biology 1994; 125: 1327–40.Google ScholarPubMed

Selected references

Boumsell, L. Cluster report: CD1. In Knapp, W, Dorken, B, Gilks, W, et al., eds. Leucocyte Typing IV: White Cell Differentiation Antigens. Oxford: Oxford University Press, 1989, pp. 251–4.Google Scholar
Krenacs, L, Tiszalvicz, LT, Krenacs, T, Boumsell, L. Immunohistochemical detection of CD1A antigen in formalin-fixed and paraffin-embedded tissue sections with monoclonal antibody 010. Journal of Pathology 1993; 171: 99104.CrossRefGoogle ScholarPubMed
Porcelli, SA, Modlin, RL. CD1 and the expanding universe of T cell antigens. Journal of Immunology 1995; 55: 3709–10.Google Scholar

Selected references

Gonzalez, L, Anderson, I, Deane, D, et al. Detection of immune system cells in paraffin wax-embedded ovine tissues. Journal of Comparative Pathology 2001; 125: 41–7.CrossRefGoogle ScholarPubMed
Knowles, DM. Lymphoid cell markers: their distribution and usefulness in the immunophenotypic analyses of lymphoid neoplasms. American Journal of Surgical Pathology 1985; 9 (Suppl.) 85108.Google Scholar

Selected references

Cabecadas, JM, Isaacson, PG. Phenotyping of T cell lymphomas in paraffin sections – which antibodies? Histopathology 1991; 19: 419–24.CrossRefGoogle ScholarPubMed
Campana, D, Thompson, JS, Amlot, P, et al. The cytoplasmic expression of CD3 antigen in normal and malignant cells of the T lymphoid lineage. Journal of Immunology 1987; 138: 648–55.CrossRefGoogle Scholar
Chetty, R, Gatter, K. CD3 structure, function, and role of immunostaining in clinical practice. Journal of Pathology 1994: 173; 303–7.Google ScholarPubMed

Selected reference

Brady, RL, Barclay, AN. The structure of CD4. Current Topics in Microbiology and Immunology 1996; 205: 118.Google ScholarPubMed

Selected reference

Arber, DA, Weiss, LM. CD5: a review. Applied Immunohistochemistry 1995; 3: 122.Google Scholar

Selected references

Lazarovits, AI, Osman, N, Le Feuvre, CE, et al. CD7 is associated with CD3 and CD45 in human T cells. Journal of Immunology 1994: 153; 3956–66.CrossRefGoogle ScholarPubMed
Saati, TA, Alibaud, L, Lamant, L, et al. A new monoclonal anti-CD7 antibody reactive in paraffin sections. Applied Immunohistochemistry and Molecular Morphology 2001; 9: 289–96.CrossRefGoogle ScholarPubMed

Selected references

Eichmann, K, Boyce, NW, Schmidt, UR, Jonsson, JI. Distinct functions of CD8 (CD4) are utilised at different stages of T lymphocyte differentiation. Immunological Reviews 1989; 109: 3975.CrossRefGoogle ScholarPubMed
Martz, E, Davignon, D, Kurzinger, K, Springer, TA. The molecular basis for cytotoxic T lymphocyte function: analysis with blocking monoclonal antibodies. Advances in Experimental Medicine and Biology 1982; 146: 447465.CrossRefGoogle Scholar
Parnes, JR. Molecular biology and function of CD4 and CD8. Advances in Immunology 1989; 44: 265311.CrossRefGoogle ScholarPubMed

Selected references

Carbone, A, Poletti, A, Manconi, R, et al. Heterogenous in situ immunotyping of follicular dendritic reticulum cells in malignant lymphomas of B cell origin. Cancer 1987; 60: 2919–26.3.0.CO;2-Z>CrossRefGoogle Scholar
Lardelli, P, Bookman, MA, Sundeen, J, et al. Lymphocytic lymphoma of intermediate differentiation: Morphologic and immunophenotypic spectrum and clinical correlations. American Journal of Surgical Pathology 1990; 14: 752–63.CrossRefGoogle ScholarPubMed
San Miguel, JF, Caballero, MD, Gonzalez, M, et al. Immunological phenotype of neoplasms involving the B cell in the last step of differentiation. British Journal of Haematology 1986; 62: 7583.CrossRefGoogle Scholar

Selected references

Anderson, KC, Bates, MP, Slaughtenhoupt, BL, et al. Expression of human B cell associated antigens on leukemias and lymphomas: a model of B cell differentiation. Blood 1984; 63: 1424–33.CrossRefGoogle Scholar
Chu, P, Arber, D. Paraffin section detection of CD10 in 505 non-hematopoietic neoplasms: frequent expression in renal cell carcinoma and endometrial stromal sarcoma. American Journal of Clinical Pathology 2000; 113: 374–82.CrossRefGoogle Scholar
Chu, P, Chang, KL, Weiss, LM, Arber, DA. Immunohistochemical detection of CD10 in paraffin sections of hematopoietic neoplasms. A comparison with flow cytometry detection in 56 cases. Applied Immunohistochemistry and Molecular Morphology 2000; 8: 257–62.Google ScholarPubMed
McCluggage, WG, Sumathi, VP, Maxwell, P. CD10 is a sensitive and diagnostically useful immunohistochemical marker of normal endometrial stroma and of endometrial stromal neoplasms. Histopathology 2001; 39: 273–8.CrossRefGoogle ScholarPubMed

Selected reference

Chadburn, A, Inghirami, G, Knowles, DM. Hairy cell leukemia-associated antigen LeuM5 (CD11c) is preferentially expressed by benign activated and neoplastic CD8 cells. American Journal of Pathology 1990; 136: 2937.Google Scholar

Selected reference

Aber, DA, Weiss, LM. CD15: a review. Applied Immunohistochemistry 1993; 1: 1730.Google Scholar

Selected references

Kehrl, JH, Riva, A, Wilson, GL, Thevenin, C. Molecular mechanisms regulating CD19, CD29 and CD22 gene expression. Immunology Today 1994; 15: 432–6.CrossRefGoogle Scholar
Scheuermann, RH, Racila, E. CD19 antigen in leukemia and lymphoma diagnosis and immunotherapy. Leukemia and Lymphoma 1995; 18: 385–97.CrossRefGoogle ScholarPubMed

Selected reference

Chang, KL, Arber, DA, Weiss, LM. CD20: a review. Applied Immunohistochemistry 1996; 4: 115.Google Scholar

Selected reference

Bagdi, E, Krenacs, L, Krenacs, T, et al. Follicular dendritic cells in reactive and neoplastic lymphoid tissues: a reevaluation of staining patterns of CD21, CD23, and CD35 antibodies in paraffin sections after wet heat-induced epitope retrieval. Applied Immunohistochemistry and Molecular Morphology 2001; 9: 117–24.CrossRefGoogle ScholarPubMed

Selected references

Armitage, RJ, Goff, LK. Functional interaction between B cell subpopulation defined by CD23 expression. European Journal of Immunology 1988; 18: 1753–60.CrossRefGoogle ScholarPubMed
DiRaimond, F, Albitar, M, Huh, Y, et al. The clinical and diagnostic relevance of CD23 expression in the chronic lymphoproliferative disease. Cancer 2002; 94: 1721–30.Google Scholar

Selected reference

Abramson, CS, Kersey, JH, LeBien, TW. A monoclonal antibody (BA-1) reactive with cells of human B lymphocyte lineage. Journal of Immunology 1981; 126: 83–8.CrossRefGoogle ScholarPubMed

Selected references

Baumgartner, C, Sonneck, K, Krauth, MT, et al. Immunohistochemical assessment of CD25 is equally sensitive and diagnostic in mastocytosis compared to flow cytometry. European Journal of Clinical Investigation 2008; 38: 326–35.CrossRefGoogle ScholarPubMed
Cerny, J, Yu, H, Ramanathan, M, et al. Expression of CD25 independently predicts early treatment failure of acute myeloid leukemia (AML). British Journal of Haematology 2013; 160: 262–6.CrossRefGoogle Scholar
Fujiwara, S, Muroi, K, Hirata, Y, et al. Clinical features of de novo CD25+ diffuse large B cell lymphoma. Hematology 2013; 18: 1419.CrossRefGoogle ScholarPubMed
Fujiwara, S, Muroi, K, Tatara, R, et al. Clinical features of de novo CD25-positive follicular lymphoma. Leukemia and Lymphoma 2014; 55: 307–13.CrossRefGoogle ScholarPubMed
Konoplev, S, Medeiros, LJ, Bueso-Ramos, CE, Jorgensen, JL. Immunophenotypic profile of lymphoplasmacytic lymphoma/Waldenström macroglobulinemia. Hematopathology 2005; 124: 414–20.Google ScholarPubMed
Lu, ZH, Chen, W, Ju, CX, et al. CD25 is a novel marker of hepatic bile canaliculus. International Journal of Surgical Pathology 2012; 20: 455–61.CrossRefGoogle ScholarPubMed
O'Malley, DP, Chizhevsky, V, Grimm, KE, Hii, A, Weiss, LM. Utility of BCL2, PD1, and CD25 immunohistochemical expression in the diagnosis of T cell lymphomas. Applied Immunohistochemistry and Molecular Morphology 2014; 22: 99104.CrossRefGoogle ScholarPubMed
Swerdlow, SH, Campo, E, Harris, NL, et al. WHO Classification of Tumours of Haematopoietic and Lymphoid Tissues. Lyon: IARC Press, 2008.Google Scholar

Selected references

Awaya, N, Mori, S, Takeuchi, H, et al. CD30 and the NPM-ALK fusion protein (p80) are differentially expressed between peripheral blood and bone marrow in primary small cell variant of anaplastic large cell lymphoma. American Journal of Hematology 2002; 69: 200–4.CrossRefGoogle ScholarPubMed
Berney, DM, Shamash, J, Pieroni, K, Oliver, RT. Loss of CD30 expression in metastatic embryonal carcinoma: the effect of therapy? Histopathology 2001; 39: 382–5.CrossRefGoogle Scholar
Chang, KL, Arber, DA, Weiss, LM. CD30: a review. Applied Immunohistochemistry 1993; 1: 244–55.Google Scholar

Selected references

Albelda, SM, Muller, WA, Buck, CA, Newman, PJ. Molecular and cellular properties of PECAM-1 (endoCAM/CD31): a novel vascular cell–cell adhesion molecule. Journal of Cell Biology 1991; 114: 1059–61.Google ScholarPubMed
DeYoung, BR, Wick, MR, Fitzgibbon, JF, et al. CD31: an immunospecific marker for endothelial differentiation in human neoplasms. Applied Immunohistochemistry 1993; 1: 97100.Google Scholar
Stokinger, H, Gadd, SJ, Eher, R, et al. Molecular characterization and functional analysis of the leukocyte surface protein CD31. Journal of Immunology 1990; 145: 3889–97.Google Scholar
Suthipintawong, C, Leong, AS-Y, Vinyuvat, S. A comparative study of immunomarkers for lymphangiomas and hemangiomas. Applied Immunohistochemistry 1995; 3: 239–44.Google Scholar
Teo, NB, Shoker, BS, Jarvis, C, et al. Vascular density and phenotype around ductal carcinoma in situ (DCIS) of the breast. British Journal of Cancer 2002; 86: 905–11.CrossRefGoogle ScholarPubMed

Selected references

Bovio, I, Allan, RW. The expression of myeloid antigens CD13 and/or CD33 is a marker of ALK+ anaplastic large cell lymphomas. American Journal of Clinical Pathology 2008; 130: 628–34.CrossRefGoogle ScholarPubMed
Golay, J, Di Gaetano, N, Amico, D, et al. Gemtuzumab ozogamicin (Mylotarg) has therapeutic activity against CD33 acute lymphoblastic leukaemias in vitro and in vivo. British Journal of Haematology 2005; 128: 310–17.CrossRefGoogle ScholarPubMed
Hoyer, JD, Grogg, KL, Hanson, CA, Gamez, JD, Dogan, A. CD33 detection by immunohistochemistry in paraffin-embedded tissues: a new antibody shows excellent specificity and sensitivity for cells of myelomonocytic lineage. American Journal of Clinical Pathology 2008; 129: 316–23.CrossRefGoogle ScholarPubMed
Jilani, I, Estey, E, Huh, Y, et al. Differences in CD33 intensity between various myeloid neoplasms. American Journal of Clinical Pathology 2002; 118: 560–6.CrossRefGoogle ScholarPubMed
Kaleem, Z, White, G. Diagnostic criteria for minimally differentiated acute myeloid leukemia (AML-M0). Evaluation and a proposal. American Journal of Clinical Pathology 2001; 115: 876–84.CrossRefGoogle ScholarPubMed
Larson, RA, Sievers, EL, Stadtmauer, EA, et al. Final report of the efficacy and safety of gemtuzumab ozogamicin (Mylotarg) in patients with CD33-positive acute myeloid leukemia in first recurrence. Cancer 2005; 104: 1442–52.CrossRefGoogle ScholarPubMed
Robillard, N, Wuillème, S, Lodé, L, et al. CD33 is expressed on plasma cells of a significant number of myeloma patients, and may represent a therapeutic target. Leukemia 2005; 19: 2021–2.CrossRefGoogle ScholarPubMed
Shin, YK Choi, EY, Kim, SH, et al. Expression of leukemia-associated antigen, JL1, in bone marrow and thymus. American Journal of Pathology 2001; 158: 1473–80.Google ScholarPubMed

Selected references

Greaves, MF, Brown, J, Molgaard, HV, et al. Molecular features of CD34: a hematopoietic progenitor cell-associated molecule. Leukemia 1992; 1: 31–6.Google Scholar
Ramani, P, Bradley, NJ, Fletcher, CMD. QBEND/10, a new monoclonal antibody to endothelium: assessment of its diagnostic utility in paraffin sections. Histopathology 1990; 17: 237–42.CrossRefGoogle ScholarPubMed

Selected references

Badgi, E, Krenacs, L, Krenacs, T, et al. Follicular dendritic cells in reactive and neoplastic lymphoid tissues: a reevaluation of staining patterns of CD21, CD23, and CD35 antibodies in paraffin sections after wet heat-induced epitope retrieval. Applied Immunohistochemistry and Molecular Morphology 2001; 9: 117–24.Google Scholar
Bettelheim, P. M8, cluster report: CD35. In Knapp, W, Dorken, B, Gilks, W, et al., eds. Leucocyte Typing IV. White Cell Differentiation Antigens. Oxford: Oxford University Press, 1989, pp. 829–30.Google Scholar

Selected references

Alessio, M, Roggero, S, Funaro, A, et al. CD38 molecule: structural and biochemical analysis on human T lymphocytes, thymocytes, and plasma cells. Journal of Immunology 1990; 145: 878–84.CrossRefGoogle Scholar
Dianzani, U, Funaro, A, DiFranco, D, et al. Interaction between endothelium and CD4+CD45RA+ lymphocytes: role of the human CD38 molecule. Journal of Immunology 1994; 153: 952–9.CrossRefGoogle ScholarPubMed

Selected reference

Carbone, A, Gloghini, A, Gruss, HJ, Pinto, A. CD40 ligand is constitutively expressed in a subset of T cell lymphomas and on the microenvironmental reactive T cells of follicular lymphomas and Hodgkin's disease. American Journal of Pathology 1995; 147: 912–22.Google Scholar

Selected references

Flavell, DJ, Flavell, SU, Jones, DB, Wright, DH. Two new monoclonal antibodies recognising T-cells (DF-T1) and B-cells (DF-B1) in formalin fixed paraffin embedded tissue sections. Journal of Pathology 1988; 155: 343A.Google Scholar
Poppema, S, Hollema, H, Visser, L, Vos, H. Monoclonal antibodies (MT1, MT2, MB1, MB2, MB3) reactive with leukocyte subsets in paraffin-embedded tissue sections. American Journal of Pathology 1987; 127: 418–29.Google ScholarPubMed

Selected reference

East, JE, Hart, IR. CD44 and its role in tumor progression and metastasis. European Journal of Cancer 1993; 29A: 1921–2.Google Scholar

Selected references

Poppema, S, Lai, R, Visser, L. Monoclonal antibody OPD4 is reactive with CD45RO but differs from UCHL1 by the absence of monocyte activity. American Journal of Pathology 1991; 139: 725–9.Google Scholar
Weiss, LM, Arber, DA, Chang, KL. CD45: a review. Applied Immunohistochemistry 1993; 1: 166–81.Google Scholar

Selected reference

Ohh, M, Takei, F. New insights into the regulation of ICAM-1 gene expression. Leukemia and Lymphoma 1996; 20: 223–8.CrossRefGoogle ScholarPubMed

Selected reference

Chan, JKC. CD56-positive putative natural killer (NK) cell lymphomas: nasal, nasal-type, blastoid, and leukemic forms. Advances in Anatomic Pathology 1997; 4: 163–72.CrossRefGoogle Scholar

Selected reference

Arber, DA, Weiss, LM. CD57. A review. Applied Immunohistochemistry 1995; 3: 137–52.Google Scholar

Selected references

Pulford, KAF, Rigney, EM, Micklem, KJ, et al. KP1: a new monoclonal antibody that detects a monocyte/macrophage associated antigen in routinely processed tissue sections. Journal of Clinical Pathology 1989; 42: 414–21.CrossRefGoogle ScholarPubMed
Pulford, KAF, Sipos, A, Cordell, JL, Stross, WP, Mason, DY. Distribution of the CD68 macrophage/myeloid associated antigen. Immunology 1990; 2: 973–80.Google ScholarPubMed

Selected references

Dong, HY, Wilkes, S, Yang, H. CD71 is selectively and ubiquitously expressed at high levels in erythroid precursors of all maturation stages: a comparative immunochemical study with glycophorin A and hemoglobin A. American Journal of Surgical Pathology 2011; 35: 723–32.CrossRefGoogle ScholarPubMed
Habashy, HO, Powe, DG, Staka, CM, et al. Transferrin receptor (CD71) is a marker of poor prognosis in breast cancer and can predict response to tamoxifen. Breast Cancer Research and Treatment 2010; 119: 283–93.CrossRefGoogle ScholarPubMed
Marsee, DK, Pinkus, GS, Yu, H. CD71 (transferrin receptor): an effective marker for erythroid precursors in bone marrow biopsy specimens. American Journal of Clinical Pathology 2010; 134: 429–35.CrossRefGoogle ScholarPubMed

Selected reference

Lazova, R, Moynes, R, May, D, Scott, G. LN-2 (CD74): a marker to distinguish atypical fibroxanthoma from malignant fibrous histiocytoma. Cancer 1997; 79: 2115–24.3.0.CO;2-N>CrossRefGoogle ScholarPubMed

Selected reference

Dunphy, CH, Polski, JM, Lance Evans, H, Gardner, IJ. Paraffin immunoreactivity of CD10, CDw75, and Bcl-6 in follicle center cell lymphoma. Leukemia and Lymphoma 2001; 41: 585–92.CrossRefGoogle ScholarPubMed

Selected references

Chu, PG, Arber, DA. CD79: a review. Applied Immunohisto-chemistry and Molecular Morphology 2001; 9: 97106.CrossRefGoogle ScholarPubMed
Mason, DY, Cordell, JL, Brown, MH, et al. CD79a: a novel marker for B cell neoplasms in routinely processed tissue samples. Blood 1995; 86: 1453–9.CrossRefGoogle ScholarPubMed

Selected references

Stevenson, AJ, Chatten, J, Bertoni, F, Miettinen, M. CD99 (p30/32MIC2) neuroectodermal/Ewing's sarcoma antigen as an immunohistochemical marker. Review of more than 600 tumors and the literature experience. Applied Immunohistochemistry 1994; 2: 231–40.Google Scholar
Vartanian, RK, Sudilovsky, D, Weidner, N. Immunostaining of monoclonal antibody 013 (anti MIC2 gene product) (CD99) in lymphomas. Impact of heat-induced epitope retrieval. Applied Immunohistochemistry 1996; 4: 4355.Google Scholar
Weidner, N, Tjoe, J. Immunohistochemical profile of monoclonal antibody 013: antibody that recognizes glycoprotein p30/32MIC2 and is useful in diagnosing Ewing's sarcoma and peripheral neuroepithelioma. American Journal of Surgical Pathology 1994; 18: 486–94.CrossRefGoogle Scholar

Selected references

Ebert, MP, Ademmer, K, Muller-Ostermeyer, F, et al. CD8+CD103+ T cells analogous to intestinal intraepithelial lymphocytes infiltrate the pancreas in chronic pancreatitis. American Journal of Gastroenterology 1998; 93: 2141–7.CrossRefGoogle ScholarPubMed
Pauls, K, Schon, M, Kubitza, RC, et al. Role of integrin αE(CD103)β7 for tissue-specific epidermal localization of CD8+ T lymphocytes. Journal of Investigative Dermatology 2001; 117: 569–75.CrossRefGoogle ScholarPubMed

Selected references

Arber, DA, Tamayo, R, Weiss, LM. Paraffin section detection of c-kit gene product (CD117) in human tissues: value in the diagnosis of mast cell disorders. Human Pathology 1998; 29: 498504.CrossRefGoogle ScholarPubMed
Ashman, LK. The biology of stem cell factor and its receptor c-kit. International Journal of Biochemistry and Cell Biology 1999; 31: 1037–51.CrossRefGoogle ScholarPubMed
Gibson, PC, Cooper, K. GD117 (KIT): a diverse protein with selective applications in surgical pathology. Advances in Anatomic Pathology 2002; 9: 65–9.CrossRefGoogle Scholar

Selected references

Del Giudice, I, Matutes, E, Morilla, R, et al. The diagnostic value of CD123 in B-cell disorders with hairy or villous lymphocytes. Haematologica 2004; 89: 303–8.Google ScholarPubMed
Julia, F, Dalle, S, Duru, G, et al. Blastic plasmacytoid dendritic cell neoplasms: clinico-immunohistochemical correlations in a series of 91 patients. American Journal of Surgical Pathology 2014; 38: 673–80.CrossRefGoogle Scholar
Kussick, SJ, Stirewalt, DL, Yi, HS, et al. A distinctive nuclear morphology in acute myeloid leukemia is strongly associated with loss of HLA-DR expression and FLT3 internal tandem duplication. Leukemia 2004; 18: 1591–8.CrossRefGoogle ScholarPubMed

Selected references

Bobardt, MD, Saphire, ACS, Hung, HC, et al. Syndecan captures, protects, and transmits HIV to T lymphocytes. Immunity 2003; 18: 2739.CrossRefGoogle ScholarPubMed
Kambham, N, Kong, C, Longacre, TA, Natkunam, Y. Utility of syndecan-1 (CD138) expression in the diagnosis of undifferentiated malignant neoplasms: a tissue microarray study of 1,754 cases. Applied Immunohisto-chemistry and Molecular Morphology 2005; 13: 304–10.Google Scholar
Kato, M, Wang, H, Kainulainen, V, et al. Physiological degradation converts the soluble syndecan-1 ectodomain from an inhibitor to a potent activator of FGF-2. Nature Medicine 1998; 4: 691–7.CrossRefGoogle ScholarPubMed
O'Connell, FP, Pinkus, JL, Pinkus, GS. CD138 (syndecan-1), a plasma cell marker immunohistochemical profile in hematopoietic and nonhematopoietic neoplasms. American Journal of Clinical Pathology 2004; 121: 254–63.CrossRefGoogle ScholarPubMed

Selected references

Lau, SK, Chu, PG, Weiss, LM. CD163: a specific marker of macrophages in paraffin-embedded tissue samples. American Journal of Clinical Pathology 2004; 122: 794801.CrossRefGoogle ScholarPubMed
Klein, JL, Nguyen, TT, Bien-Willner, GA, et al. CD163 immunohistochemistry is superior to CD68 in predicting outcome in classical Hodgkin lymphoma. American Journal of Clinical Pathology 2014; 141: 381–7.CrossRefGoogle ScholarPubMed
Kridel, R, Xerri, L, Gelas-Dore, B, et al. The prognostic impact of CD163-positive macrophages in follicular lymphoma: a study from the BC Cancer Agency and the Lumphoma Study Association. Clinical Cancer Research 2015; 21: 3428–35.CrossRefGoogle Scholar
Nguyen, TT, Schwartz, EJ, West, RB, et al. Expression of CD163 (hemoglobin scavenger receptor) in normal tissues, lymphomas, carcinomas and sarcomas is largely restricted to the monocyte/macrophage lineage. American Journal of Surgical Pathology 2005; 29: 617–24.CrossRefGoogle Scholar

Selected references

Binh, MB, Sastre-Garau, X, Guillou, L, et al. MDM2 and CDK4 immunostainings are useful adjuncts in diagnosing well-differentiated and dedifferentiated liposarcoma subtypes: a comparative analysis of 559 soft tissue neoplasms with genetic data. American Journal of Surgical Pathology 2005; 29: 1340–7.CrossRefGoogle ScholarPubMed
Coindre, JM, Hostein, I, Maire, G, et al. Inflammatory malignant fibrous histiocytomas and dedifferentiated liposarcomas: histological review, genomic profile, and MDM2 and CDK4 status favour a single entity. Journal of Pathology 2004; 203: 822–30.CrossRefGoogle Scholar
Dei Tos, AP, Doglioni, C, Piccinin, S, et al. Coordinated expression and amplification of the MDM2, CDK4, and HMGI-C genes in atypical lipomatous tumours. Journal of Pathology 2000; 190: 531–6.3.0.CO;2-W>CrossRefGoogle ScholarPubMed
Dujardin, F, Binh, MB, Bouvier, C, et al. MDM2 and CDK4 immunohistochemistry is a valuable tool in the differential diagnosis of low-grade osteosarcomas and other primary fibro-osseous lesions of the bone. Modern Pathology 2011; 24: 624–37.CrossRefGoogle ScholarPubMed
Thway, K, Flora, R, Shah, C, Olmos, D, Fisher, C. Diagnostic utility of p16, CDK4, and MDM2 as an immunohistochemical panel in distinguishing well-differentiated and dedifferentiated liposarcomas from other adipocytic tumors. American Journal of Surgical Pathology 2012; 36: 462–9.CrossRefGoogle ScholarPubMed

Selected references

Bellizzi, AM. Assigning site of origin in metastatic neuroendocrine neoplasms: a clinically significant application of diagnostic immunohistochemistry. Advances in Anatomic Pathology 2013; 20: 285314.CrossRefGoogle ScholarPubMed
da Costa, LT, He, TC, Yu, J, et al. CDX2 is mutated in a colorectal cancer with normal APC/beta-catenin signaling. Oncogene 1999; 18: 5010–14.CrossRefGoogle Scholar
Freund, JN, Domon-Dell, C, Kedinger, M, et al. The Cdx-1 and Cdx-2 homeobox genes in the intestine. Biochemistry and Cell Biology 1998; 76: 957–69.CrossRefGoogle ScholarPubMed
Jaffee, IM, Rahmani, M, Singhal, MG, et al. Expression of the intestinal transcription factor CDX2 in carcinoid tumors is a marker of midgut origin. Archives of Pathology and Laboratory Medicine 2006; 130: 1522–6.CrossRefGoogle ScholarPubMed
James, R, Erler, T, Kazenwadel, J. Structure of the murine homeobox gene cdx-2. Expression in embryonic and adult intestinal epithelium. Journal of Biological Chemistry 1994; 269: 15, 229–37.CrossRefGoogle ScholarPubMed
Li, MK, Folpe, AL. CDX-2, a new marker for adenocarcinoma of gastrointestinal origin. Advances in Anatomic Pathology 2004; 11: 101–5.CrossRefGoogle ScholarPubMed
Werling, RW, Yaziji, H, Bacchi, CE, et al. CDX2, a highly sensitive and specific marker of adenocarcinomas of intestinal origin: an immunohistochemical survey of 476 primary and metastatic carcinomas. American Journal of Surgical Pathology 2003; 27: 303–10.CrossRefGoogle ScholarPubMed

Selected references

Bobrow, LG, Happerfield, LC, Millis, RR. Comparison of immunohistological staining with different antibodies to the c-erbB-2 oncoprotein. Applied Immunohistochemistry 1996; 4: 128–34.Google Scholar
DePotter, CR. The new-oncogene: more than a prognostic indicator? Human Pathology 1994; 25: 1264–8.Google Scholar

Selected reference

Beatty, WL, Morison, RP, Byrne, GI. Immunoelectron microscopic quantitation of differential levels of chlamydial proteins in cell culture model of persistent Chlamydia trachomatis infection. Infection and Immunity 1994; 62: 4059–62.CrossRefGoogle ScholarPubMed

Selected reference

Hendy, GN, Bevan, S, Mattei, MG, Mouland, AJ. Chromogranin A. Clinical Investigative Medicine 1995; 18: 4765.Google ScholarPubMed

Selected reference

Vastrik, I, Makela, TP, Koskinen, PJ, et al. Myc protein: partners and antagonists. Critical Reviews in Oncology 1994; 5: 5968.CrossRefGoogle Scholar

Selected reference

Birembaut, P, Caron, Y, Adnet, JJ. Usefulness of basement membrane markers in tumoral pathology. Journal of Pathology 1985; 145: 283–96.CrossRefGoogle Scholar

Selected references

Burkle, A, Neidermeier, M, Schmitt-Graff, A, et al. Overexpression of the CXCR5 chemokine receptor, and its ligand, CXCL13 in B cell chronic lymphocytic leukemia. Blood 2007; 110: 3316–25.CrossRefGoogle ScholarPubMed
Dupuis, J, Boye, K, Martin, N, et al. Expression of CXCL13 by neoplastic cells in angioimmunoblastic T cell lymphoma (AITL): A new diagnostic marker providing evidence that AITL derives from follicular helper T cells. American Journal of Surgical Pathology 2006; 30: 490–4.CrossRefGoogle ScholarPubMed
Nam-Cha, SH, Roncador, G, Sanchez-Verde, L, et al. PD-1, a follicular T-cell marker useful for recognizing nodular lymphocyte-predominant Hodgkin lymphoma. American Journal of Surgical Pathology 2008; 32: 1252–7.CrossRefGoogle ScholarPubMed
Ortonne, N, Dupuis, J, Plonquet, A, et al. Characterization of CXCL13+ neoplastic T cells in cutaneous lesions of angioimmunoblastic T cell lymphoma (AITL). American Journal of Surgical Pathology 2007; 31: 1068–76.CrossRefGoogle ScholarPubMed

Selected reference

Bartkova, J, Lukas, J, Strauss, M, Bartek, J. Cell cycle-related variation and tissue-restricted expression of human cyclin D1 protein. Journal of Pathology 1994; 172: 237–45.CrossRefGoogle ScholarPubMed

Selected references

Battifora, H. Diagnostic uses of antibodies to keratins: a review and immunohistochemical comparison of seven monoclonal and three polyclonal antibodies. In Fenoglio-Preiser, CM, Wolff, M, Rilke, F, eds. Progress in Surgical Pathology, Vol. VIII. Berlin: Springer-Verlag, 1988, pp. 1015.Google Scholar
Heatley, MK. Cytokeratins and cytokeratin staining in diagnostic histopathology (commentary). Histopathology 1996; 28: 479–83.CrossRefGoogle Scholar
Miettinen, M. Keratin immunohistochemistry: update of applications and pitfalls. Pathology Annual 1993; 28 (2): 113–43.Google ScholarPubMed

Selected references

Chu, PG, Weiss, LM. Expression of cytokeratin 5/6 in epithelial neoplasms: an immunohistochemical study of 509 cases. Modern Pathology 2002; 15: 610.CrossRefGoogle ScholarPubMed
Chu, PG, Weiss, LM. Keratin expression in human tissues and neoplasms. Histopathology 2002; 40: 403–39.CrossRefGoogle ScholarPubMed

Selected reference

Moll, R, Franke, WW, Schiller, DL, Geiger, B, Krepler, R. The catalog of human cytokeratins: patterns of expression in normal epithelia, tumors and cultured cells. Cell 1982; 31: 1124.CrossRefGoogle ScholarPubMed

Selected references

Ali, A, Serra, S, Asa, SL, et al. The predictive value of CK19 and CD99 in pancreatic endocrine tumors. American Journal of Surgical Pathology 2006; 30: 1588–94.CrossRefGoogle ScholarPubMed
Chetty, R. An overview of practical issues in the diagnosis of gastroenteropancreatic neuroendocrine pathology. Archives of Pathology and Laboratory Medicine 2008; 132: 1285–9.CrossRefGoogle ScholarPubMed
Deshpande, V, Fernandez-del Castillo, C, Muzikansky, A, et al. Cytokeratin 19 is a powerful predictor of survival in pancreatic endocrine tumors. American Journal of Surgical Pathology 2004; 28: 1145–53.CrossRefGoogle ScholarPubMed
Jain, R, Fischer, S, Serra, S, et al. The use of cytokeratin 19 (CK19) immunohistochemistry in lesions of the pancreas, gastrointestinal tract, and liver. Applied Immunohistochemistry and Molecular Morphology 2010; 18: 915.CrossRefGoogle ScholarPubMed
Nechifor-Boila, A, Borda, A, Sassolas, G, et al. Immunohistochemical markers in the diagnosis of papillary thyroid carcinomas: the promising role of combined immunostaining using HBME-1 and CD56. Pathology, Research and Practice 2013; 209: 585–92.CrossRefGoogle ScholarPubMed
Rosai, J. Immunohistochemical markers of thyroid tumors: significance and diagnostic applications. Tumori 2003; 89: 517–19.CrossRefGoogle ScholarPubMed
Ryu, HS, Lee, K, Shin, E, et al. Comparative analysis of immunohistochemical markers for differential diagnosis of hepatocelluar carcinoma and cholangiocarcinoma. Tumori 2012; 98: 478–84.CrossRefGoogle ScholarPubMed

Selected reference

Chan, JKC, Suster, S, Wenig, BM, et al. Cytokeratin 20 immunoreactivity distinguishes Merkel cell (primary cutaneous neuroendocrine) carcinomas and salivary gland small cell carcinomas from small cell carcinomas of various sites. American Journal of Surgical Pathology 1997; 21: 226–34.CrossRefGoogle ScholarPubMed

Selected reference

Battifora, H. Diagnostic uses of antibodies to keratins: a review and immunohistochemical comparison of seven monoclonal and three polyclonal antibodies. In Fenoglio-Preiser, CM, Wolff, M, Rilke, F. eds. Progress in Surgical Pathology, Vol. VIII. Berlin: Springer-Verlag, 1988, pp. 1015.Google Scholar

Selected reference

Moll, R, Franke, WW, Schiller, DL, Geiger, B, Krepler, R. The catalog of human cytokeratins: patterns of expression in normal epithelia, tumors and cultured cells. Cell 1982; 31: 1124.CrossRefGoogle ScholarPubMed

Selected reference

Battifora, H. Diagnostic uses of antibodies to keratins: a review and immunohistochemical comparison of seven monoclonal and three polyclonal antibodies. In Fenoglio-Preiser, CM, Wolff, M, Rilke, F. eds. Progress in Surgical Pathology, Vol. VIII. Berlin: Springer-Verlag, 1988, pp. 1015.Google Scholar

Selected reference

Cooper, D, Schermer, A, Sun, TT. Classification of human epithelia and their neoplasms using monoclonal antibodies to keratins: strategies, applications and limitations. Laboratory Investigation 1985; 52: 243–56.Google ScholarPubMed

Selected references

Miettinen, M. Keratin immunohistochemistry: update of applications and pitfalls. Pathology Annual 1993; 28 (2): 113–43.Google ScholarPubMed
Moll, R, Franke, WW, Schiller, DL, Geiger, B, Krepler, R. The catalog of human cytokeratins: patterns of expression in normal epithelia, tumors and cultured cells. Cell 1982; 31: 1124.CrossRefGoogle ScholarPubMed

Selected references

Swenson, PD, Kaplan, MH. Rapid detection of cytomegalovirus in cell culture by indirect immunoperoxidase staining with monoclonal antibody to an early nuclear antigen. Journal of Clinical Microbiology 1985; 21: 669–73.CrossRefGoogle Scholar
Zweygberg, WB, Wirgart, B, Grillner, L. Early detection of cytomegalovirus in cell culture by a monoclonal antibody. Journal of Virological Methods 1986; 14: 65–9.Google Scholar

Selected reference

Anderson, P, Nagler-Anderson, C, O'Brien, C, et al. A monoclonal antibody reactive with a 15-kDa cytoplasmic granule associated protein defines a subpopulation of CD8+ T lymphocytes. Journal of Immunology 1990; 144: 574–82.CrossRefGoogle ScholarPubMed

Selected references

Chu, AY, Litzky, LA, Pasha, TL, Acs, G, Zhang, PJ. Utility of D2-40, a novel mesothelial marker, in the diagnosis of malignant mesothelioma. Modern Pathology 2005; 18: 105–10.CrossRefGoogle ScholarPubMed
Kahn, HJ, Bailey, D, Marks, A. Monoclonal antibody D2-40, a new marker of lymphatic endothelium, reacts with Kaposi's sarcoma and a subset of angiosarcomas. Modern Pathology 2002; 15: 434–40.CrossRefGoogle Scholar
Kalof, AN, Cooper, K. D2-40 immunohistochemistry: so far! Advances in Anatomic Pathology 2009; 16: 62–4.CrossRefGoogle ScholarPubMed

Selected references

Cordone, I, Annino, L, Masi, S, et al. Diagnostic relevance of peripheral blood immunocytochemistry in hairy cell leukaemia. Journal of Clinical Pathology 1995; 48: 955–60.CrossRefGoogle ScholarPubMed
Hounieu, H, Chittal, SM, al Saati, T, et al. Hairy cell leukemia. Diagnosis of bone marrow involvement in paraffin-embedded sections with monoclonal antibody DBA.44. American Journal of Clinical Pathology 1992; 98: 2633.Google ScholarPubMed

Selected references

Azumi, N, Ben-Ezra, J, Battifora, H. Immunophenotypic diagnosis of leiomyosarcomas and rhabdomyosarcomas with monoclonal antibodies to muscle-specific actin and desmin in formalin-fixed tissue. Modern Pathology 1988; 1: 469–74.Google ScholarPubMed
Li, Z, Colucci, E, Babinet, C, Paulin, D. The human desmin gene: a specific regulatory program in skeletal muscle both in vitro and in transgenic mice. Neuromuscular Disorders 1993; 3: 423–7.CrossRefGoogle ScholarPubMed
Nagai, J, Capetanaki, YG, Lazarides, E. Expression of the genes coding for the intermediate filament proteins vimentin and desmin. Annals of the New York Academy of Sciences 1985; 455: 144–55.Google Scholar
Pollock, L, Rampling, D, Greenwald, SE, Malone, M. Desmin expression in rhabdomyosarcoma: influence of the desmin clone and immunohistochemical method. Journal of Clinical Pathology 1995; 48: 535–8.CrossRefGoogle ScholarPubMed

Selected reference

O'Keefe, EJ, Erickson, HP, Bennett, V. Desmoplakin I and desmoplakin II: purification and characterization. Journal of Biological Chemistry 1989; 264: 8310–18.CrossRefGoogle ScholarPubMed

Selected references

Chênevert, J, Duvvuri, U, Chiosea, S, et al. DOG1: a novel marker of salivary acinar and intercalated duct differentiation. Modern Pathology 2012; 25, 919–29.CrossRefGoogle ScholarPubMed
Choi, JJ, Sinada-Bottros, L, Maker, AV, Weisenberg, E. Dedifferentiated gastrointestinal stromal tumor arising de novo from the small intestine. Pathology Research and Practice 2014; 210: 264–6.CrossRefGoogle ScholarPubMed
Espinosa, I, Lee, CH, Kim, MK, et al. A novel monoclonal antibody against DOG1 is a sensitive and specific marker for gastrointestinal stromal tumors. American Journal of Surgical Pathology 2008; 32: 210–18.CrossRefGoogle ScholarPubMed
Hemminger, J, Iwenofu, OH. Discovered on gastrointestinal stromal tumours 1 (DOG1) expression in non-gastrointestinal stromal tumour (GIST) neoplasms. Histopathology 2012; 61: 170–7.CrossRefGoogle ScholarPubMed
Liegl, B, Hornick, JL, Corless, CL, Fletcher, CD. Monoclonal antibody DOG1.1 shows higher sensitivity than KIT in the diagnosis of gastrointestinal stromal tumors, including unusual subtypes. American Journal of Surgical Pathology 2009; 33: 437–46.CrossRefGoogle ScholarPubMed
Miettinen, M, Wang, ZF, Lasota, J. DOG1 antibody in the differential diagnosis of gastrointestinal stromal tumors: a study of 1840 cases. American Journal of Surgical Pathology 2009; 33: 1401–8.CrossRefGoogle ScholarPubMed
Sah, SP, McCluggage, WG. DOG1 immunoreactivity in uterine leiomyosarcomas. Journal of Clinical Pathology 2013; 66: 40–3.CrossRefGoogle ScholarPubMed

Selected references

Lagna, G, Hata, A, Hemmati-Brivanlou, A, et al. Partnership between DPC4 and SMAD proteins in TGF-beta signalling pathways. Nature 1996; 383: 832–6.CrossRefGoogle ScholarPubMed
Maitra, A, Molberg, K, Albores-Saavedra, J, et al. Loss of Dpc4 expression in colonic adenocarcinomas correlates with the presence of metastatic disease. American Journal of Pathology 2000; 157: 1105–11.CrossRefGoogle ScholarPubMed
Wilentz, RE, Iacobuzio-Donahue, CA, Argani, P, et al. Loss of expression of Dpc4 in pancreatic intraepithelial neoplasia: evidence that DPC4 inactivation occurs late in neoplastic progression. Cancer Research 2000; 60: 2002–6.Google ScholarPubMed
Wilentz, RE, Su, GH, Dai, JL, et al. Immunohistochemical labeling for dpc4 mirrors genetic status in pancreatic adenocarcinomas : a new marker of DPC4 inactivation. American Journal of Pathology 2000; 156: 3743.CrossRefGoogle ScholarPubMed
Xia, X, Wu, W, Huang, C, et al. SMAD4 and its role in pancreatic cancer. Tumour Biology 2014; 36: 111–19.Google ScholarPubMed

Selected reference

Heyderman, E, Strudley, I, Powell, G, et al. A new monoclonal antibody to epithelial membrane antigen (EMA) – E29. A comparison of its immunocytochemical reactivity with polyclonal anti-EMA antibodies and with another monoclonal antibody HMFG-2. British Journal of Cancer 1985; 52: 355–61.CrossRefGoogle ScholarPubMed

Selected references

Fan, X, Liu, B, Xu, H, et al. Immunostaining with EGFR mutation-specific antibodies: a reliable screening method for lung adenocarcinomas harboring EGFR mutation in biopsy and resection samples. Human Pathology 2013; 44: 1499–507.CrossRefGoogle ScholarPubMed
Lindeman, NI, Cagle, PT, Beasley, MB, et al. Molecular Testing Guideline for Selection of Lung Cancer Patients for EGFR and ALK Tyrosine Kinase Inhibitors: Guideline from the College of American Pathologists, International Association for the Study of Lung Cancer, and Association for Molecular Pathology. Archives of Pathology and Laboratory Medicine 2013; 137: 828–60.CrossRefGoogle Scholar
Wen, YH, Brogi, E, Hasanovic, A, et al. Immunohistochemical staining with EGFR mutation-specific antibodies: high specificity as a diagnostic marker for lung adenocarcinoma. Modern Pathology 2013; 26: 1197–203.CrossRefGoogle ScholarPubMed
Yousem, SA.Role of molecular studies in the diagnosis of lung adenocarcinoma. Modern Pathology 2012; 25 (suppl 1): S1117.CrossRefGoogle ScholarPubMed

Selected references

Carpenter, G. Properties of the receptor for epidermal growth factor. Cell 1984; 37: 357–8.CrossRefGoogle ScholarPubMed
Chen, WS, Lazar, CS, Lund, KA, et al. Functional independence of the epidermal growth factor receptor from a domain required for ligand-induced internalization and calcium regulation. Cell 1989; 59: 3343.CrossRefGoogle ScholarPubMed

Selected references

Bruin, PCD. Detection of Epstein–Barr virus nucleic acid sequences and protein in nodal T-cell lymphomas: relation between latent membrane protein-1 positivity and clinical course. Histopathology 1993; 23: 509–18.CrossRefGoogle ScholarPubMed
Delecluse, H-J, Kremmer, E, Rouault, JP, et al. The expression of Epstein-Barr virus latent proteins is related to the pathological feature of post-transplant lymphoproliferative disorders. American Journal of Pathology 1995; 146: 1113–20.Google Scholar
Hording, U, Nielsen, HW, Albeck, H, Daugaard, S. Nasopharyngeal carcinoma: histopathological types and association with Epstein-Barr virus. European Journal of Cancer Clinical Oncology 1993; 29B: 137–9.Google ScholarPubMed
Kanavaros, P, Lecsc, M-C, Briere, J, et al. Nasal T-cell lymphoma: a clinicopathologic entity associated with peculiar phenotype and with Epstein-Barr virus. Blood 1993; 81: 2688–95.CrossRefGoogle ScholarPubMed

Selected references

Birdsey, GM, Dryden, NH, Amsellem, V, et al. Transcription factor ERG regulates angiogenesis and endothelial apoptosis through VE-cadherin. Blood 2008; 111: 3498–506.CrossRefGoogle ScholarPubMed
Falzarano, SM, Zhou, M, Carver, P, et al. ERG gene rearrangement status in prostate cancer detected by immunohistochemistry.Virchows Archiv 2011; 459: 441–7.CrossRefGoogle ScholarPubMed
Kumar-Sinha, C, Tomlins, SA, Chinnaiyan, AM. Recurrent gene fusions in prostate cancer. Nature Reviews: Cancer 2008; 8: 497511.Google ScholarPubMed
Miettinen, M, Wang, ZF, Paetau, A, et al. ERG transcription factor as an immunohistochemical marker for vascular endothelial tumors and prostatic carcinoma. American Journal of Surgical Pathology 2011; 35: 432–41.CrossRefGoogle ScholarPubMed
Miettinen, M, Wang, Z, Sarlomo-Rikala, M, et al. ERG expression in epithelioid sarcoma: a diagnostic pitfall. American Journal of Surgical Pathology 2013; 37: 1580–5.CrossRefGoogle ScholarPubMed
Sun, C, Dobi, A, Mohamed, A, et al. TMPRSS2-ERG fusion, a common genomic alteration in prostate cancer activates C-MYC and abrogates prostate epithelial differentiation. Oncogene 2008; 27: 5348–53.CrossRefGoogle ScholarPubMed
Wang, WL, Patel, NR, Caragea, M, et al. Expression of ERG, an Ets family transcription factor, identifies ERG-rearranged Ewing sarcoma. Modern Pathology 2012; 25: 1378–83.CrossRefGoogle ScholarPubMed

Selected reference

Acs, G, Acs, P, Beckwith, SM, et al. Erythropoietin and erythropoietin receptor expression in human cancer. Cancer Research 2001; 61: 3561–5.Google ScholarPubMed

Selected references

Juric, G, Zarkovic, N, Nola, M, et al. The value of cell proliferation and angiogenesis in the prognostic assessment of ovarian granulosa cell tumors. Tumori 2001; 87: 4753.CrossRefGoogle ScholarPubMed
Marder, VJ, Mannucci, PM, Firkin, BG, et al. Standard nomenclature for factor VIII and von Willebrand factor: a recommendation by the International Committee on Thrombosis and Haemostasis. Thrombosis and Hemostasis 1985; 54: 871–2.Google ScholarPubMed
Sehested, M, Hou-Jensen, K. Factor VII-related antigen as an endothelial cell marker in benign and malignant diseases. Virchows Archiv A. Pathological Anatomy and Histology 1981; 391: 217–25.CrossRefGoogle ScholarPubMed

Selected references

Alawi, F, Stratton, D, Freedman, PD. Solitary fibrous tumor of the oral soft tissues: a clinicopathologic and immunohistochemical study of 16 cases. American Journal of Surgical Pathology 2001; 25: 900–10.CrossRefGoogle Scholar
Busam, KJ, Granter, SR, Iversen, K, Jungbluth, AA. Immunohistochemical distinction of epithelioid histiocytic proliferations from epithelioid melanocytic nevi. American Journal of Dermatopathology 2000; 22: 237–41.CrossRefGoogle ScholarPubMed
Cerio, R, Spaull, J, Oliver, GF, Wilson-Jones, E. A study of factor XIIIa and MAC387 immunolabeling in normal and pathological skin. American Journal of Dermatopathology 1990; 12: 221–33.CrossRefGoogle Scholar
Fucich, LF, Cheles, MK, Thung, SN, et al. Primary versus metastatic hepatic carcinoma. An immunohistochemical study of 34 cases. Archives of Pathology and Laboratory Medicine 1994; 118: 927–30.Google Scholar
Hill, KA, Gonzalez-Crussi, F, Chou, PM. Calcifying fibrous pseudotumor versus inflammatory myofibroblastic tumor: a histological and immunohistochemical comparison. Modern Pathology 2001; 14: 784–90.CrossRefGoogle ScholarPubMed
Kraus, MD, Haley, JC, Ruiz, R, et al. “Juvenile” xanthogranuloma: an immunophenotypic study with a reappraisal of histiogenesis. American Journal of Dermatopathology 2001; 23: 104–11.CrossRefGoogle Scholar
Leong, AS-Y, Lim, MHT. Immunohistochemical characteristics of dermatofibrosarcoma protuberans. Applied Immunohistochemistry 1994; 2: 427.Google Scholar
Mentzel, T, Kutzner, H, Rutten, A, Hugel, H. Benign fibrous histiocytoma (dermatofibroma) of the face: clinicopathologic and immunohistochemical study of 34 cases associated with an aggressive clinical behaviour. American Journal of Dermatopathology 2001; 23: 419–26.CrossRefGoogle Scholar
Nemes, Z, Thomaszy, V. Factor XIIIa and the classic histiocytic markers in malignant fibrous histiocytoma. Human Pathology 1988; 9: 822–9.Google Scholar
Nestle, FO, Nickoloff, BJ. A fresh morphological and functional look at dermal dendritic cells. Journal of Cutaneous Pathology 1995; 22: 385–93.CrossRefGoogle Scholar
Nestle, FO, Nickoloff, BJ, Burg, G. Dermatofibroma: an abortive immunoreactive process mediated by dermal dendritic cells? Dermatology 1995; 190: 265–8.CrossRefGoogle ScholarPubMed
Takata, M, Imai, T, Hirone, T. Factor XIIIa-positive cells in normal peripheral nerves and cutaneous neurofibromas of type-1 neurofibromatosis. American Journal of Dermatopathology 1994; 16: 3743.CrossRefGoogle ScholarPubMed
Zelger, BW, Zelger, BG, Steiner, H, Ofner, D. Aneurysmal and hemangiopericytoma-like fibrous histiocytoma. Journal of Clinical Pathology 1996; 49: 313–18.CrossRefGoogle Scholar

Selected references

Bamberger, AM, Schulte, HM, Thuneke, I, et al. Expression of the apoptosis-inducing Fas ligand (FasL) in human first and third trimester placenta and choriocarcinoma cells. Journal of Clinical Endocrinology and Metabolism 1997; 82: 3173–5.CrossRefGoogle ScholarPubMed
De la Monte, SM, Sohn, YK, Wands, JR. Correlates of p53- and Fas (CD95)-mediated apoptosis in Alzheimer's disease. Journal of Neurological Sciences 1997; 152: 7383.CrossRefGoogle ScholarPubMed
Hellquist, HB, Olejnicka, B, Jadner, M, et al. Fas receptor is expressed in human lung squamous cell carcinomas, whereas bcl-2 and apoptosis are not pronounced: a preliminary report. British Journal of Cancer 1997; 76: 175–9.CrossRefGoogle Scholar
Kazufumi, M, Sonoko, N, Masanori, K, et al. Expression of bcl-2 protein and APO-1 (Fas antigen) in the lung tissue from patients with idiopathic pulmonary fibrosis. Microscopy Research Technology 1997; 38: 480–7.3.0.CO;2-M>CrossRefGoogle ScholarPubMed
Lee, J, Richburg, JH, Younkin, SC, Bockelheide, K. The Fas system is a key regulator of germ cell apoptosis in the testis. Endocrinology 1997; 138: 2081–8.CrossRefGoogle ScholarPubMed
Nichans, GA, Brunner, T, Frizelle, SP, et al. Human lung carcinomas express Fas ligand. Cancer Research 1997; 57: 1007–12.Google Scholar
Nonomura, N, Mild, T, Yokoyama, M, et al. Fas/APO-1-mediated apoptosis of human renal cell carcinoma. Biochemistry Biophysiology Research Communications 1996; 229: 945–51.Google ScholarPubMed
Sheard, MA, Vojtesek, B, Janakova, L, et al. Up-regulation of Fas (CD95) in human p53 wild-type cancer cells treated with ionizing radiation. International Journal of Cancer 1997; 73: 757–62.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Shukuwa, T, Katayama, I, Koji, T. Fas-mediated apoptosis of melanoma cells and infiltrating lymphocytes in human malignant melanomas. Modern Pathology 2002; 15: 387–96.CrossRefGoogle ScholarPubMed
Strater, J, Wellisch, I, Riedl, S, et al. CD95 (APO-l/Fas)-mediated apoptosis in colon epithelial cells: a possible role in ulcerative colitis. Gastroenterology 1997; 113: 160–7.CrossRefGoogle Scholar
Tachibana, O, Lampe, J, Kleihues, P, Obgaki, H. Preferential expression of Fas/APOl (CD95) and apoptotic cell death in perinecrotic cells of glioblastoma multiforme. Acta Neuropathologica (Berlin) 1996; 92: 431–4.CrossRefGoogle Scholar
Uckan, D, Steele, A, Wang, BY, et al. Trophoblasts express Fas ligand: a proposed mechanism for immune privilege in placenta and maternal invasion. Molecular Human Reproduction 1997; 3: 655–62.CrossRefGoogle ScholarPubMed

Selected references

Biddle, DA, Ro, JY, Yoon, GS, et al. Extranodal follicular dendritic cell sarcoma of the head and neck region: three new cases, with a review of the literature. Modern Pathology 2002; 15: 50–8.CrossRefGoogle ScholarPubMed
Dako Corporation. Anti-human fascin, 55K-2 (data sheet).Google Scholar
Duh, FM, Latif, F, Weng, Y, et al. cDNA cloning and expression of the human homolog of the sea urchin fascin and Drosophila singed genes which encodes an actin-bundling protein. DNA and Cell Biology 1994; 13: 821–7.CrossRefGoogle ScholarPubMed
Gaertner, EM, Tsokos, M, Derringer, GA, et al. Interdigitating dendritic cell sarcoma. A report of four cases and review of the literature. American Journal of Clinical Pathology 2001: 115: 589–97.CrossRefGoogle ScholarPubMed
Kansal, R, Singleton, TP, Ross, CW, et al. Follicular Hodgkin lymphoma: a histopathologic study. American Journal of Clinical Pathology 2002; 117: 2935.CrossRefGoogle ScholarPubMed
Kraus, MD, Haley, JC, Ruiz, R, et al. “Juvenile” xanthogranuloma: an immunophenotypic study with a reappraisal of histogenesis. American Journal of Dermatopathology 2001; 23: 104–11.CrossRefGoogle ScholarPubMed
Mosialos, G, Birkenbach, M, Ayehunie, S, et al. Circulating human dendritic cells differentially express high levels of 55-kd actin-bundling protein. American Journal of Pathology 1996; 148: 593600.Google ScholarPubMed
Pinkus, GS, Pinkus, JL, Langhoff, E, et al. Fascin, a sensitive new marker for Reed–Sternberg cells of Hodgkin's disease: evidence for a dendritic or B-cell derivation? American Journal of Pathology 1997; 150: 543–62.Google ScholarPubMed
Said, JW, Pinkus, JL, Shintaku, IP, et al. Alterations in fascin-expressing germinal center dendritic cells in neoplastic follicles of B-cell lymphomas. Modern Pathology 1998; 11: 15.Google ScholarPubMed
Yamashiro-Matsumura, S, Matsumura, F. Purification and characterization of an F-actin-bundling 55-kilodalton protein from HeLa cells. Journal of Biological Chemistry 1985; 260: 5087–97.CrossRefGoogle ScholarPubMed

Selected references

Abenoza, P, Manivel, JC, Wick, MR, Hagen, K, Dehner, LP. Hepatoblastoma: an immunohistochemical and ultrastructural study. Human Pathology 1987; 18: 1025–35.CrossRefGoogle ScholarPubMed
Carter, RL, Hall, JM, Corbett, RP. Immunohistochemical staining for ferritin in neuroblastomas. Histopathology 1991; 18: 465–8.CrossRefGoogle ScholarPubMed
Fleming, S. Immunocytochemical localization of ferritin in the kidney and renal tumors. European Urology 1987; 13: 407–11.CrossRefGoogle Scholar
Harrison, PM, Arosio, P. The ferritins: molecular properties, iron storage function and cellular regulation. Biochemia Biophysiologica Acta 1996; 1275: 161203.Google ScholarPubMed
Imoto, M, Nishimura, D, Fukuda, Y, et al. Immunohistochemical detection of alpha-fetoprotein, carcinoembryonic antigen, and ferritin in formalin-fixed sections from hepatocellular carcinoma. American Journal of Gastroenterology 1985; 80: 902–6.Google Scholar
Johnson, DE, Powers, CN, Rupp, G, et al. Immunocytochemical staining of fine needle aspiration biopsies of the liver as a diagnostic tool for hepatocellular carcinoma. Modern Pathology 1992; 5: 117–23.Google ScholarPubMed
Kaneko, Y, Kitamoto, T, Tateishi, J, Yamaguchi, K. Ferritin immunohistochemistry as a marker for microglia. Acta Neuropathologica (Berlin) 1989; 79: 129–36.CrossRefGoogle ScholarPubMed
Momotani, E, Wuscger, N, Ravisse, P, Rastogi, N. Immunohistochemical identification of ferritin, lactoferrin and transferrin in leprosy lesions of human skin biopsies. Journal of Comparative Pathology 1992; 106: 213–20.CrossRefGoogle ScholarPubMed
Navone, R, Azzoni, L, Valente, G. Immunohistochemical assessment of ferritin in bone marrow trephine biopsies: correlation with marrow hemosiderin. Acta Hematologica 1988; 80: 194–8.CrossRefGoogle ScholarPubMed
Nogales, FF, Concha, A, Plata, C, Ruiz-Avila, I. Granulosa cell tumor of the ovary with diffuse true hepatic differentiation simulating stromal luteinization. American Journal of Surgical Pathology 1993; 17: 8590.CrossRefGoogle ScholarPubMed
Ozawa, H, Nishida, A, Mito, T, Takashima, S. Immunohistochemical study of ferritin-positive cells in the cerebellar cortex with subarachnoid hemorrhage in neonates. Brain Research 1994; 65: 345–8.Google Scholar
Papadimitriou, JC, Drachenberg, CB, Brenner, DS, et al. “Thanatosomes”: a unifying morphogenetic concept for tumor hyaline globules related to apoptosis. Human Pathology 2000; 31: 1455–65.CrossRefGoogle ScholarPubMed
Penneys, NS, Zlatkiss, I. Immunohistochemical demonstration of ferritin in sweat gland and sweat gland neoplasms. Journal of Cutaneous Pathology 1990; 17: 32–6.CrossRefGoogle ScholarPubMed
Tuccari, G, Rizzo, A, Crisafulli, C, Barresi, G. Iron-binding proteins in human colorectal adenomas and carcinomas: an immunohistochemical investigation. Histology and Histopathology 1992; 7: 543–7.Google Scholar

Selected references

Bini, A, Kudryk, BJ. Fibrinogen and fibrin in the arterial wall. Thrombosis Research 1994; 75: 337–41.CrossRefGoogle ScholarPubMed
Blomback, B. Fibrinogen structure, activation and polymerization and fibrin gel structure. Thrombosis Research 1994; 75: 327–8.CrossRefGoogle ScholarPubMed
Blomback, B. Fibrinogen and fibrin: proteins with complex roles in hemostasis and thrombosis. Thrombosis Research 1996; 83: 175.CrossRefGoogle ScholarPubMed
Bonsib, SM. Differential diagnosis in nephropathology: an immunofluorescence-driven approach. Advances in Anatomic Pathology 2002; 9: 101–14.CrossRefGoogle ScholarPubMed
Dowling, JP. Immunohistochemistry of renal diseases and tumours. In Leong, AS-Y, ed. Applied Immunohistochemistry for Surgical Pathologists. London: Edward Arnold, 1993, pp. 210–59.Google Scholar
Gaffhey, PJ. Structure of fibrinogen and degradation products of fibrinogen and fibrin. British Medical Bulletin 1997; 33: 245–51.Google Scholar
Imokawa, S, Sato, A, Hayakawa, H, et al. Tissue factor expression and fibrin deposition in the lungs of patients with idiopathic pulmonary fibrosis and systemic sclerosis. American Journal of Respiratory and Critical Care Medicine 1997; 156: 631–6.CrossRefGoogle ScholarPubMed
Kahng, HC, Chin, NW, Opitz, LM, Pahuja, M, Goldberg, SL. Cellular angiolipoma of the breast: immunohistochemical study and review of the literature. Breast Journal 2002; 8: 47–9.CrossRefGoogle ScholarPubMed
Lorand, L. Physiological roles of fibrinogen and fibrin. Federation Proceedings 1965; 24: 784–93.Google ScholarPubMed
Mosessan, MW. The roles of fibrinogen and fibrin in hemostasis and thrombosis. Seminars in Hematology 1992; 29: 177–88.Google Scholar
Mosessan, MW. Fibrinogen and fibrin polymerization: appraisal of the binding events that accompany fibrin generation and fibrin clot assembly. Blood Coagulation and Fibrinolysis 1997; 8: 257–67.Google Scholar
Takahashi, H, Shibata, Y, Fujita, S, Okabe, H. Immunohistochemical findings of arterial fibrinoid necrosis in major and lingual minor salivary glands of primary Sjogren's syndrome. Anatomical and Cellular Pathology 1996; 12: 145–57.Google ScholarPubMed

Selected references

Blomback, B. Fibrinogen and fibrin -proteins with complex roles in hemostasis and thrombosis. Thrombosis Research 1996; 83: 175.CrossRefGoogle ScholarPubMed
Dowling, JP. Immunohistochemistry of renal diseases and tumours. In Leong, AS-Y, ed. Applied Immunohistochemistry for Surgical Pathologists. London: Edward Arnold, 1993, pp. 210–59.Google Scholar
Gaffney, PJ. Structure of fibrinogen and degradation products of fibrinogen and fibrin. British Medical Bulletin 1997; 33: 245–51.Google Scholar
Henschen, A. On the structure of functional sites in fibrinogen. Thrombosis Research 1983; 5 (Suppl): 2739.CrossRefGoogle ScholarPubMed
Mosesson, MW. The roles of fibrinogen and fibrin in hemostasis and thrombosis. Seminars in Hematology 1992; 29: 177–88.Google ScholarPubMed
Mosesson, MW. Fibrinogen and fibrin polymerization: appraisal of the binding events that accompany fibrin generation and fibrin clot assembly. Blood Coagulation and Fibrinolysis 1997; 8: 257–67.CrossRefGoogle ScholarPubMed
Shafer, JA, Higgins, DL. Human fibrinogen. Critical Reviews in Clinical Laboratory Science 1988; 26: 141.Google ScholarPubMed
Stewart, FA, Te Poele, JA, Van der Wal, AF, et al. Radiation nephropathy: the link between functional damage and vascular mediated inflammatory and thrombotic change. Acta Oncologica 2001; 40: 952–7.Google Scholar

Selected references

D’ Arderme, AJ, Burns, J, Skyes, BC, Bennett, MK. Fibronectin and type III collagen in epithelial neoplasms of gastrointestinal tract and salivary gland. Journal of Clinical Pathology 1983; 36: 756–63.Google Scholar
D'Ardenne, AJ, Kirkpatrick, P, Sykes, BC. The distribution of laminin, fibronectin and interstitial collagen type III in soft tissue tumors. Journal of Clinical Pathology 1984; 37: 895904.CrossRefGoogle Scholar
D'Ardenne, AJ, Kirkpatrick, P, Wells, CA, Davies, JD. Laminin and fibronectin in adenoid cystic carcinoma. Journal of Clinical Pathology 1986; 39: 138–44.CrossRefGoogle ScholarPubMed
Kirkpatrick, P, d'Ardenne, AJ. Effects of fixation and enzymatic digestion on the immunohistochemical demonstration of laminin and fibronectin in paraffin embedded tissue. Journal of Clinical Pathology 1984; 37: 639–44.CrossRefGoogle ScholarPubMed
Laurie, GW, Leblond, CP, Martin, GR. Localization of type IV collagen, laminin, heparan sulfate proteoglycan and fibronectin to the basal lamina of basement membranes. Journal of Cell Biology 1982; 95: 340–4.Google Scholar
Mosher, DF, Fiocht, L. Fibronectin: review of its structure and possible functions. Journal of Investigative Dermatology 1981; 77: 175–80.CrossRefGoogle ScholarPubMed
Ortiz-Rey, JA, Suarez-Penaranda, JM, da Silva, EA, et al. Immunohistochemical detection of fibronectin and tenascin in incised human skin injuries. Forensic Science International 2002; 126: 118–22.CrossRefGoogle ScholarPubMed
Stenman, S, Vaheri, A. Distribution of a major connective tissue protein, fibronectin in normal human tissues. Journal of Experimental Medicine 1978; 147: 1054–64.CrossRefGoogle Scholar
Verbeke, S, Gotteland, M, Fernandez, M, et al. Basement membrane and connective tissue proteins in intestinal mucosa of patients with celiac disease. Journal of Clinical Pathology 2002; 55: 440–5.CrossRefGoogle Scholar

Selected references

Cuda, J, Mirzamani, N, Kantipudi, R, et al. Diagnostic utility of Fli-1 and D2–40 in distinguishing atypical fibroxanthoma from angiosarcoma. American Journal of Dermatopathology 2013; 35: 316–18.CrossRefGoogle ScholarPubMed
Folpe, AL, Chand, EM, Goldblum, JR, Weiss, SW. Expression of Fli-1, a nuclear transcription factor, distinguishes vascular neoplasms from potential mimics. American Journal of Surgical Pathology 2001; 25: 1061–6.CrossRefGoogle ScholarPubMed
Folpe, AL, Hill, CE, Parham, DM, et al. Immunohistochemical detection of FLI-1 protein expression: A study of 132 round cell tumors with emphasis on CD99-positive mimics of Ewing's sarcoma/primitive neuroectodermal tumor. American Journal of Surgical Pathology 2000; 24: 1657–62.CrossRefGoogle ScholarPubMed
Nilsson, G, Wang, M, Wejde, J, et al. Detection of EWS/FLI-1 by immunostaining. An adjunctive tool in diagnosis of Ewing's sarcoma and primitive neuroectodermal tumor on cytological samples and paraffin-embedded archival material. Sarcoma 1999; 3: 2532.CrossRefGoogle ScholarPubMed
Turc-Carel, C, Aurias, A, Mugneret, F, et al. Chromosomes in Ewing's sarcoma. I: an evaluation of 85 cases of remarkable consistency of t(11; 22)(q24; q12). Cancer Genetics and Cytogenetics 1988; 32: 229–38.CrossRefGoogle Scholar
Zucman, J, Delattre, O, Desmaze, C, et al. Cloning and characterization of the Ewing's sarcoma and peripheral neuroepithelioma t(11; 22) translocation breakpoints. Genes Chromosomes and Cancer 1992; 5: 271–7.CrossRefGoogle ScholarPubMed

Selected references

Catovsky, D, Brooks, D. Bradley, J, et al. Heterogeneity of B-cell leukemias demonstrated by the monoclonal antibody FMC-7. Blood 1981; 58: 406–8.CrossRefGoogle Scholar
Drexler, HG, Menon, M, Gaedicke, G, Minowada, J. Expression of FMC7 antigen and tartrate-resistant acid phosphatase isoenzyme in cases of B-lymphoproliferative diseases. European Journal of Cancer and Clinical Oncology 1987; 23: 61–8.CrossRefGoogle ScholarPubMed
Garcia, DP, Rooney, MT, Ahmad, E, Davis, BH. Diagnostic usefulness of CD23 and FMC-7 antigen expression patterns in B-cell lymphoma classification. American Journal of Clinical Pathology 2001; 115: 258–65.CrossRefGoogle ScholarPubMed
Hubl, W, Iturraspe, J, Braylan, RC. FMC7 antigen expression on normal and malignant B-cells can be predicted by expression of CD20. Cytometry 1998; 34: 71–4.3.0.CO;2-F>CrossRefGoogle ScholarPubMed
Huh, YO, Pugh, WC, Kantarjian, HM, et al. Detection of subgroups of chronic B-cell leukemias as FMC7 monoclonal antibody. American Journal of Clinical Pathology 1994; 101: 283–9.CrossRefGoogle ScholarPubMed
Menon, M. Drexler, HG, Minowada, J. Heterogeneity of marker expression in B-cell leukemias and its diagnostic significance. Leukemia Research 1986; 10: 25–8.CrossRefGoogle ScholarPubMed
Zola, H, McNamara, PJ, Moore, HA, et al. Maturation of human B lymphocytes: studies with a panel of monoclonal antibodies against membrane antigens. Clinical and Experimental Immunology 1983; 52: 655–64.Google Scholar
Zola, H, Moore, HA, Hohmann, A, Hunter, IK. The antigen of mature human B cells detected by the monoclonal antibody FMC7: studies on the nature of the antigen and modulation of its expression. Journal of Immunology 1984; 133: 321–6.CrossRefGoogle Scholar

Selected references

Al-Agha, OM, Huwait, HF, Chow, C, et al. FOXL2 is a sensitive and specific marker for sex cord–stromal tumors of the ovary. American Journal of Surgical Pathology 2011; 35: 484–94.CrossRefGoogle ScholarPubMed
Shah, SP, Kobel, M, Senz, J, et al. Mutation of FOXL2 in granulosa-cell tumors of the ovary. New England Journal of Medicine 2009; 360: 2719–29.CrossRefGoogle ScholarPubMed
Stewart, CJ, Alexiadis, M, Crook, ML, Fuller, PJ. An immunohistochemical and molecular analysis of problematic and unclassified ovarian sex cord–stromal tumors. Human Pathology 2013; 44: 2774–81.CrossRefGoogle ScholarPubMed

Selected references

Chang, A, Amin, A, Gabrielson, E, et al. Utility of GATA3 immunohistochemistry in differentiating urothelial carcinoma from prostate adenocarcinoma and squamous cell carcinomas of the uterine cervix, anus, and lung. American Journal of Surgical Pathology 2012; 36: 1472–6.CrossRefGoogle ScholarPubMed
Ellis, CL, Chang, AG, Cimino-Mathews, A, et al. GATA-3 immunohistochemistry in the differential diagnosis of adenocarcinoma of the urinary bladder. American Journal of Surgical Pathology 2013; 37: 1756–60.CrossRefGoogle ScholarPubMed
Liu, H, Shi, J, Prichard, JW, Gong, Y, Lin, F. Immunohistochemical evaluation of GATA-3 expression in ER-negative breast carcinomas. American Journal of Clinical Pathology 2014; 141: 648–55.CrossRefGoogle ScholarPubMed
Liu, H, Shi, J, Wilkerson, ML, Lin, F. Immunohistochemical evaluation of GATA3 expression in tumors and normal tissues: a useful immunomarker for breast and urothelial carcinomas. American Journal of Clinical Pathology 2012; 138: 5764.CrossRefGoogle ScholarPubMed
Ordonez, NG. Value of GATA3 immunostaining in tumor diagnosis: a review. Advances in Anatomic Pathology 2013; 20: 352–60.Google ScholarPubMed
Simon, MC. Gotta have GATA. Nature Genetics 1995; 11: 911.CrossRefGoogle ScholarPubMed
Yoon, NK, Maresh, EL, Shen, D, et al. Higher levels of GATA3 predict better survival in women with breast cancer. Human Pathology 2010; 41: 1794–801.CrossRefGoogle ScholarPubMed

Selected reference

Inagaki, M, Nakamura, Y, Takeda, M, et al. Glial fibrillary acidic protein: dynamic property and regulation by phosphorylation. Brain Pathology 1994; 4: 239–43.CrossRefGoogle ScholarPubMed

Selected references

Brown, RS, Wahl, RL. Overexpression of Glut-1 glucose transporter in human breast cancer: an immunohistochemical study. Cancer 1993; 72: 2979–85.3.0.CO;2-X>CrossRefGoogle ScholarPubMed
Haber, RS, Weiser, KR, Pritsker, A, Reder, I, Burstein, DE. GLUT1 glucose transporter expression in benign and malignant thyroid nodules. Thyroid 1997; 7: 363–7.CrossRefGoogle ScholarPubMed
Kato, Y, Tsuta, K, Seki, K, et al. Immunohistochemical detection of GLUT-1 can discriminate between reactive mesothelium and malignant mesothelioma. Modern Pathology 2007; 20: 215–20.CrossRefGoogle ScholarPubMed
Lyons, LL, North, PE, Mac-Moune Lai, F, et al. Kaposiform hemangioendothelioma: a study of 33 cases emphasizing its pathologic, immunophenotypic, and biologic uniqueness from juvenile hemangioma. American Journal of Surgical Pathology 2004; 28: 559–68.CrossRefGoogle ScholarPubMed
Mentzel, T, Kutzner, H. Reticular and plexiform perineurioma: clinicopathological and immunohistochemical analysis of two cases and review of perineurial neoplasms of skin and soft tissues. Virchows Archiv 2005; 447: 677–82.CrossRefGoogle ScholarPubMed
Sakashita, M, Aoyama, N, Minami, R, et al. Glut1 expression in T1 and T2 stage colorectal carcinomas: its relationship to clinicopathological features. European Journal of Cancer 2001; 37: 204–9.CrossRefGoogle ScholarPubMed

Selected references

Akutsu, N, Yamamoto, H, Sasaki, S, et al. Association of glypican-3 expression with growth signaling molecules in hepatocellular carcinoma. World Journal of Gastroenterology 2010; 16: 3521–8.CrossRefGoogle ScholarPubMed
Capurro, M, Wanless, IR, Sherman, M, et al. Glypican-3: a novel serum and histochemical marker for hepatocellular carcinoma. Gastroenterology 2003; 125: 8997.CrossRefGoogle ScholarPubMed
Capurro, MI, Xiang, YY, Lobe, C, et al. Glypican-3 promotes the growth of hepatocellular carcinoma by stimulating canonical Wnt signaling. Cancer Research 2005; 65: 6245–54.CrossRefGoogle ScholarPubMed
He, H, Fang, W, Liu, X, et al. Frequent expression of glypican-3 in Merkel cell carcinoma: an immunohistochemical study of 55 cases. Applied Immunohistochemistry and Molecular Morphology 2009; 17: 40–6.CrossRefGoogle ScholarPubMed
Maeda, D, Ota, S, Takazawa, Y, et al. Glypican-3 expression in clear cell adenocarcinoma of the ovary. Modern Pathology 2009; 22: 824–32.CrossRefGoogle ScholarPubMed
Ou-Yang, RJ, Hui, P, Yang, XJ, et al. Expression of glypican 3 in placental site trophoblastic tumor. Diagnostic Pathology 2010; 5: 64.CrossRefGoogle ScholarPubMed
Shafizadeh, N, Kakar, S. Diagnosis of well-differentiated hepatocellular lesions: role of immunohistochemistry and other ancillary techniques. Advances in Anatomic Pathology 2011; 18: 438–45.CrossRefGoogle ScholarPubMed
Ushiku, T, Uozaki, H, Shinozaki, A, et al. Glypican 3-expressing gastric carcinoma: distinct subgroup unifying hepatoid, clear-cell, and alpha-fetoprotein-producing gastric carcinomas. Cancer Science 2009; 100: 626–32.CrossRefGoogle ScholarPubMed
Zynger, DL, McCallum, JC, Luan, C, et al. Glypican 3 has a higher sensitivity than alpha-fetoprotein for testicular and ovarian yolk sac tumour: immunohistochemical investigation with analysis of histological growth patterns. Histopathology 2010; 56: 750–7.CrossRefGoogle Scholar

Selected references

Fiel, MI, Cernainu, G, Burstein, DE, Batheja, N. Value of GCDFP-15 (BRST-2) as a specific immunocytochemical marker for breast carcinoma in cytologic specimens. Acta Cytologica 1996; 40: 637–41.CrossRefGoogle ScholarPubMed
Haagensen, DE, Dilley, WG, Mazoujian, G, Wells, SA. Review of GCDFP-15. An apocrine marker protein. Annals of the New York Academy of Sciences 1990; 586: 161–73.CrossRefGoogle ScholarPubMed

Selected references

Ascoli, V, Carnovale-Scalzo, C, Taccogna, S, Nardi, F. Utility of HBME-1 immunostaining in serous effusions. Cytopathology 1997; 8: 328–35.CrossRefGoogle ScholarPubMed
Attanoos, RL, Goddard, H, Gibbs, AR. Mesothelioma-binding antibodies: thrombomodulin, OV 632 and HBME-1 and their use in the diagnosis of malignant mesothelioma. Histopathology 1996; 29: 209–15.CrossRefGoogle ScholarPubMed

Selected references

Cartun, RW, Kryzmowski, GA, Pedersen, CA, et al. Immunocytochemical identification of H. pylori in formalin-fixed gastric biopsies. Modern Pathology 1991; 4: 498502.Google ScholarPubMed
Genta, RM, Robason, GO, Graham, DY. Simultaneous visualization of Helicobacter pylori and gastric morphology: a new stain. Human Pathology 1994; 25: 221–6.CrossRefGoogle ScholarPubMed

Selected references

Fasano, M, Theise, ND, Nalesnik, M, et al. Immunohistochemical evaluation of hepatoblastomas with use of the hepatocyte-specific marker, hepatocyte paraffin 1, and the polyclonal anti-carcinoembryonic antigen. Modern Pathology 1998; 11: 934–8.Google ScholarPubMed
Leong, AS-Y, Sormunen, RT, Tsui, WM-S, Liew, CT. Immunostaining for liver cancers. Histopathology 1998; 33: 318–24.Google Scholar

Selected reference

Burns, J. Immunoperoxidase localization of hepatitis B antigen (HB) in formalin-paraffin processed liver tissue. Histochemistry 1975; 44: 133–5.CrossRefGoogle ScholarPubMed

Selected reference

Gudat, F, Bianchi, L. HGsAg: a target antigen on the liver cell? In Popper, H, Bianchi, L, Reutter, W, eds. Membrane Alterations as Basis of Liver Injury. Lancaster: MTP Press, 1977, pp. 171–8.Google Scholar

Selected reference

Miliauskas, J, Leong, AS-Y. Localized herpes simplex lymphadenitis: report of three cases and review of the literature. Histopathology 1991; 19: 355–60.CrossRefGoogle ScholarPubMed

Selected references

Cesarman, E, Chang, Y, Moore, PS, Said, JW, Knowles, DM. Kaposi's sarcoma-associated herpesvirus-like DNA sequences in AIDS-related body-cavity-based lymphomas. New England Journal of Medicine 1995; 332: 1186–91.CrossRefGoogle ScholarPubMed
Kazakov, DV, Schmid, M, Adams, V, et al. HHV-8 DNA sequences in the peripheral blood and skin lesions of an HIV-negative patient with multiple eruptive dermatofibromas: implications for the detection of HHV-8 as a diagnostic marker for Kaposi's sarcoma. Dermatology 2003; 206: 217221.CrossRefGoogle ScholarPubMed
McDonagh, DP, Liu, J, Gaffey, MJ, et al. Detection of Kaposi's sarcoma-associated herpesvirus-like DNA sequence in angiosarcoma. American Journal of Pathology 1996; 149: 1363–8.Google ScholarPubMed
Soulier, J, Grollet, L, Oksenhendler, E, et al. Kaposi's sarcoma-associated herpesvirus-like DNA sequences in multicentric Castleman's disease. Blood. 1995; 86: 1276–80.CrossRefGoogle ScholarPubMed

Selected references

Crumpton, MJ, Bodmer, JC, Bodmer, WF, et al. Biochemistry of class II antigens: workshop report. In Albert, ED, Mayr, WR, eds. Histocompatibility Testing. Berlin: Springer-Verlag, 1984, pp. 2937.Google Scholar
Jendro, M, Goronzy, JJ, Weyand, CM. Structural and functional characterization of HLA-DR molecules circulating in the serum. Autoimmunity 1991; 8: 289–96.CrossRefGoogle ScholarPubMed

Selected reference

Bacchi, CE, Bonetti, Pea M, Martignoni, G, Gown, AM. HMB-45. A review. Applied Immunohistochemistry 1996; 4: 7385.Google Scholar

Selected references

Jiricny, J. Eukaryotic mismatch repair: an update. Mutation Research 1998; 409: 107–21.CrossRefGoogle ScholarPubMed
Lanza, G, Gafa, R, Maestri, I, et al. Immunohistochemical pattern of MLH1/MSH2 expression is related to clinical and pathological features in colorectal adenocarcinomas with microsatellite instability. Modern Pathology 2002; 15: 741–9.CrossRefGoogle ScholarPubMed

Selected references

Daugharty, H, Long, EG, Swisher, BC, et al. Comparative study with in situ hybridization and immunocytochemistry in detection of HIV-1 in formalin-fixed paraffin-embedded cell cultures. Journal of Clinical Laboratory Analysis 1990; 4: 283–8.CrossRefGoogle ScholarPubMed
Kaluza, G, Willems, WR, Lohmeyer, J, et al. A monoclonal antibody that recognizes a formalin-resistant epitope on the p24 core protein of HIV-1. Pathology Research and Practice 1992; 188: 91–6.CrossRefGoogle ScholarPubMed

Selected references

Cooper, K, McGee, J O'D. Human papillomavirus, integration and cervical carcinogenesis: a clinicopathological perspective. Molecular Pathology 1997; 50: 13.CrossRefGoogle ScholarPubMed
Patel, D, Shepherd, PS, Naylor, JA, McCance, DJ. Reactivities of polyclonal and monoclonal antibodies raised to the major capsid protein of human papillomavirus type 16. Journal of Virology 1989; 70: 6977.Google Scholar

Selected references

Brown, KE, Anderson, SM, Young, NS. Erythrocyte P antigen: cellular receptor of B19 parvovirus. Science 1993; 262: 114–17.CrossRefGoogle ScholarPubMed
Liu, W, Ittmann, MD, Liu, J, et al. Human parvovirus B19 in bone marrows from adults with acquired immunodeficiency syndrome: a comparative study using in situ hybridization and immunohistochemistry. Human Pathology 1997; 28: 760–6.CrossRefGoogle ScholarPubMed
Morey, AL, O'Neill, HJ, Coyle, PV, et al. Immunohistological detection of human parvovirus B19 in formalin-fixed, paraffin-embedded tissues. Journal of Pathology 1992; 166: 105–8.CrossRefGoogle ScholarPubMed

Selected references

Brescia, RJ, Kurman, RJ, Main, CS, et al. Immunocytochemical localization of chorionic gonadotropin, placental lactogen, and placental alkaline phosphatase in the diagnosis of complete and partial hydatidiform moles. International Journal of Gynecological Pathology 1987; 6: 213–29.CrossRefGoogle ScholarPubMed
Cheah, PL, Looi, LM. Expression of placental proteins in complete and partial hydatidiform moles. Pathology 1994; 26: 115–18.CrossRefGoogle ScholarPubMed

Selected references

Leong, AS-Y, Forbes, IJ. Immunological and histochemical techniques in the study of the malignant lymphomas: A review. Pathology 1982; 14: 247–54.CrossRefGoogle Scholar
Merz, H, Pickers, O, Schrimel, S, et al. Constant detection of surface and cytoplasmic immunoglobulin heavy and light chain expression in formalin-fixed and paraffin-embedded material. Journal of Pathology 1993; 170: 257–64.CrossRefGoogle ScholarPubMed

Selected references

Deshpande, V, Zen, Y, Chang, JK, et al. Consensus statement on the pathology of IgG4-related disease. Modern Pathology 2012; 25: 1181–92.CrossRefGoogle ScholarPubMed
Wallace, ZS, Deshpande, V, Mattoo, H, et al. IgG4-related disease: clinical and laboratory features in one hundred and twenty-five patients. Arthritis and Rheumatology 2015; 67: 2466–75.CrossRefGoogle ScholarPubMed

Selected reference

Flemming, P, Wellman, A, Maschek, H, Lang, H, Georgii, A. Monoclonal antibodies against inhibin represent key markers of adult granulosa cell tumors of the ovary even in their metastases. A report of three cases with late metastasis, being previously misinterpreted as hemangiopericytoma. American Journal of Surgical Pathology 1995; 19: 927–33.CrossRefGoogle ScholarPubMed

Selected references

Agaimy, A, Erlenbach-Wünsch, K, Konukiewitz, B, et al. ISL1 expression is not restricted to pancreatic well-differentiated neuroendocrine neoplasms, but is also commonly found in well and poorly differentiated neuroendocrine neoplasms of extrapancreatic origin. Modern Pathology 2013; 26: 9951003.CrossRefGoogle Scholar
Ahlgren, U, Pfaff, SL, Jessell, TM, Edlund, T, Edlund, H. Independent requirement for ISL1 in formation of pancreatic mesenchyme and islet cells. Nature 1997; 385: 257–60.CrossRefGoogle ScholarPubMed
Bellizzi, AM. Assigning site of origin in metastatic neuroendocrine neoplasms: a clinically significant application of diagnostic immunohistochemistry. Advances in Anatomic Pathology 2013; 20: 285314.CrossRefGoogle ScholarPubMed
Graham, RP, Shrestha, B, Caron, BL, et al. Islet-1 is a sensitive but not entirely specific marker for pancreatic neuroendocrine neoplasms and their metastases. American Journal of Surgical Pathology 2013; 37: 399404.CrossRefGoogle Scholar
Hermann, G, Konukiewitz, B, Schmitt, A, Perren, A, Klöppel, G. Hormonally defined pancreatic and duodenal neuroendocrine tumors differ in their transcription factor signatures: expression of ISL1, PDX1, NGN3, and CDX2. Virchows Archiv 2011; 459: 147–54.CrossRefGoogle ScholarPubMed
Schmitt, AM, Riniker, F, Anlauf, M, et al. Islet 1 (Isl1) expression is a reliable marker for pancreatic endocrine tumors and their metastases. American Journal of Surgical Pathology 2008; 32: 420–5.CrossRefGoogle ScholarPubMed

Selected reference

Brown, DC, Gatter, KC. Monoclonal antibody Ki-67: Its use in histopathology. Histopathology 1990; 17: 489503.CrossRefGoogle ScholarPubMed

Selected references

Leong, AS-Y, Vinyuvat, S, Suthipintawong, C, Leong, FJ. Patterns of basal lamina immunostaining in soft-tissue and bony tumors. Applied Immunohistochemistry 1997; 5: 17.CrossRefGoogle Scholar
Liotta, LA. Tumor invasion and metastases: Role of the basement membrane. Warner-Lambert Parke-Davis Award Lecture. American Journal of Pathology 1984; 117: 339–48.Google ScholarPubMed

Selected references

Bishop, JA, Yonescu, R, Batista, D, et al. Utility of mammaglobin immunohistochemistry as a proxy marker for the ETV6-NTRK3 translocation in the diagnosis of salivary mammary analogue secretory carcinoma. Human Pathology 2013; 44: 1982–8.CrossRefGoogle ScholarPubMed
Leygue, E, Snell, L, Dotzlaw, H, et al. Mammaglobin, a potential marker of breast cancer nodal metastasis. Journal of Pathology 1999; 189: 2833.3.0.CO;2-H>CrossRefGoogle ScholarPubMed
Reyes, C, Gomez-Fernandez, C, Nadji, M. Metaplastic and medullary mammary carcinomas do not express mammaglobin. American Journal of Clinical Pathology 2012; 137: 747–52.CrossRefGoogle Scholar
Sasaki, E, Tsunoda, N, Hatanaka, Y, et al. Breast-specific expression of MGB1/mammaglobin: an examination of 480 tumors from various organs and clinicopathological analysis of MGB1-positive breast cancers. Modern Pathology 2007; 20: 208–14.CrossRefGoogle ScholarPubMed

Selected reference

Busam, KJ, Jungbluth, AA. Melan-A, a new melanocytic differentiation marker. Advances in Anatomic Pathology 1999; 6: 182–7.CrossRefGoogle ScholarPubMed

Selected reference

Gelsleichter, L, Gown, AM, Zarbo, RJ, et al. P53 and mdm-2 expression in malignant melanoma: an immunocytochemical study of expression of p53, mdm-2, and markers of cell proliferation in primary versus metastatic tumors. Modern Pathology 1995; 8: 530–5.Google ScholarPubMed

Selected reference

Allen, IV, McQuaid, S, McMahon, J, et al. The significance of measles virus antigen and genome distribution in the CNS in SSPE for mechanisms of viral spread and demyelination. Journal of Neuropathology and Experimental Neurology 1996; 55: 471–80.CrossRefGoogle ScholarPubMed

Selected references

Albelda, SM, Muller, WA, Buck, CA, Newman, PJ. Molecular and cellular properties of PECAM-1 (endoCAM/CD31): a novel vascular cell–cell adhesion molecule. Journal of Cell Biology 1991; 114: 1059–68.Google ScholarPubMed
Kuzu, I, Bicknell, R, Fletcher, CDM, Gatter, KC. Expression of adhesion molecules on the endothelium of normal tissue vessels and vascular tumors. Laboratory Investigation 1993; 69: 322–8.Google ScholarPubMed

Selected references

Chang, K, Pastan, I. Molecular cloning of mesothelin, a differentiation antigen present on mesothelium, mesotheliomas, and ovarian cancers. Proceedings of the National Academy of Sciences of the USA 1996; 93: 136–40.CrossRefGoogle ScholarPubMed
Ordóñez, NG. Value of mesothelin immunostaining in the diagnosis of mesothelioma.Modern Pathology 2003; 16: 192–7.CrossRefGoogle ScholarPubMed

Selected references

Jasani, B, Schmid, KW. Significance of metallothionein overexpression in human tumors. Histopathology 1997; 31: 211–14.CrossRefGoogle Scholar
Kagi, JHR. Overview of methallothionein. Metallobiochemistry Part B: metallothionein and related molecules. Methods in Enzymology 1993; 205: 613–26.Google Scholar

Selected reference

Busam, KJ, Iversen, K, Copian, KC, Jungbluth, AA. Analysis of microphthalmia transcription factor expression in normal tissues and tumors, and comparison of its expression with S-100 protein, gp100, and tyrosinase in desmoplastic malignant melanoma. American Journal of Surgical Pathology 2001; 25: 197204.CrossRefGoogle ScholarPubMed

Selected reference

BiogenexLaboratories. Monoclonal antibody to mitochondrial antigen (data sheet).Google Scholar

Selected reference

Edwards, C, Oates, J. OV 632 and MOC 31 in the diagnosis of mesothelioma and adenocarcinoma: an assessment of their use in formalin fixed and paraffin wax embedded material. Journal of Clinical Pathology 1995; 48: 626–30.CrossRefGoogle ScholarPubMed

Selected references

Azumi, N, Ben-Erza, J, Battifora, H. Immunophenotypic diagnosis of leiomyosarcomas and rhabdomyosarcomas with monoclonal antibodies to muscle specific actin and desmin in formalin-fixed tissue. Modern Pathology 1988; 1: 469–74.Google ScholarPubMed
Rangdaeng, S, Truong, LD. Comparative immunohistochemical staining for desmin and muscle specific actin: a study of 576 cases. American Journal of Clinical Pathology 1991; 96: 3245.CrossRefGoogle ScholarPubMed

Selected references

Falini, B, Fizzotti, M, Pucciarini, A, et al. A monoclonal antibody (MUM1p) detects expression of the MUM1/IRF4 protein in a subset of germinal center B cells, plasma cells, and activated T cells. Blood 2000; 95: 2084–92.CrossRefGoogle Scholar
Gualco, G, Weiss, LM, Bacchi, CE. MUM1/IRF4: a review. Applied Immunohistochemistry and Molecular Morphology 2010; 18: 301–10.CrossRefGoogle ScholarPubMed
Klein, U, Casola, S, Cattoretti, G, et al. Transcription factor IRF4 controls plasma cell differentiation and class-switch recombination. Nature Immunology 2006; 7: 773–82.CrossRefGoogle ScholarPubMed
Shaffer, AL, Emre, NC, Lamy, L, et al. IRF4 addiction in multiple myeloma. Nature 2008; 454: 226–31.CrossRefGoogle ScholarPubMed

Selected references

Andrulis, M., Penzel, R., Weichert, W., von Deimling, A., Capper, D. Application of a BRAF V600E mutation-specific antibody for the diagnosis of hairy cell leukemia. American Journal of Surgical Pathology 2012; 36: 1796–800.CrossRefGoogle ScholarPubMed
Koperek, O, Kornauth, C, Capper, D, et al. Immunohistochemical detection of the BRAF V600E-mutated protein in papillary thyroid carcinoma. American Journal of Surgical Pathology 2012; 36: 844–50.CrossRefGoogle ScholarPubMed
Long, GV, Wilmott, JS, Capper, D, et al. Immunohistochemistry is highly sensitive and specific for the detection of V600E BRAF mutation in melanoma. American Journal of Surgical Pathology 2013; 37: 61–5.CrossRefGoogle ScholarPubMed
Preusser, M, Capper, D, Berghoff, AS, et al. Expression of BRAF V600E mutant protein in epithelial ovarian tumors. Applied Immunohistochemistry and Molecular Morphology 2013; 21: 159–64.CrossRefGoogle ScholarPubMed
Sinicrope, FA, Smyrk, TC, Tougeron, D, et al. Mutation-specific antibody detects mutant BRAFV600E protein expression in human colon carcinomas. Cancer 2013; 119: 2765–70.CrossRefGoogle ScholarPubMed
Wan, PT, Garnett, MJ, Roe, SM, et al. Mechanism of activation of the RAF-ERK signaling pathway by oncogenic mutations of B-RAF. Cell 2004; 116: 855–67.CrossRefGoogle ScholarPubMed

Selected references

Byers, R, Hornick, JL, Tholouli, E, Kutok, J, Rodig, SJ.Detection of IDH1 R132H mutation in acute myeloid leukemia by mutation-specific immunohistochemistry. Applied Immunohistochemistry and Molecular Morphology 2012; 20: 3740.CrossRefGoogle ScholarPubMed
Kato, Y. Specific monoclonal antibodies against IDH1/2 mutations as diagnostic tools for gliomas. Brain Tumor Pathology 2015; 32: 311.CrossRefGoogle ScholarPubMed
Liu, X, Kato, Y, Kaneko, MK, et al. Isocitrate dehydrogenase 2 mutation is a frequent event in osteosarcoma detected by a multi-specific monoclonal antibody MsMab-1. Cancer Medicine 2013; 2: 803–14.CrossRefGoogle ScholarPubMed
Mauzo, SH, Lee, M, Petros, J, et al. Immunohistochemical demonstration of isocitrate dehydrogenase 1 (IDH1) mutation in a small subset of prostatic carcinomas. Applied Immunohistochemistry and Molecular Morphology 2014; 22: 284–7.CrossRefGoogle Scholar
Preusser, M, Capper, D, Hartmann, C; Euro-CNS Research Committee. IDH testing in diagnostic neuropathology: review and practical guideline article invited by the Euro-CNS Research Committee. Clinical Neuropathology 2011; 30: 217–30.CrossRefGoogle ScholarPubMed
Takano, S, Tian, W, Matsuda, M, et al. Detection of IDH1 mutation in human gliomas: comparison of immunohistochemistry and sequencing. Brain Tumor Pathology 2011; 28: 115–23.Google ScholarPubMed

Selected reference

Carabias, E, Palenque, E, Serrano, R, et al. Evaluation of an immunohistochemical test with polyclonal antibodies raised against mycobacteria used in formalin-fixed tissue compared with mycobacterial specific culture. APMIS 1998; 106: 385–8.CrossRefGoogle ScholarPubMed

Selected references

Mason, DY, Taylor, CR. The distribution of muramidase (lysozyme) in human tissues. Journal of Clinical Pathology 1975; 28: 124–32.CrossRefGoogle ScholarPubMed
Pinkus, GS, Pinkus, JL. Myeloperoxidase: a specific marker for myeloid cells in paraffin sections. Modern Pathology 1991; 4: 733–41.Google ScholarPubMed

Selected references

Cessna, MH, Zhou, H, Perkins, SL, et al. Are myogenin and myoD1 expression specific for rhabdomyosarcoma? A study of 150 cases, wioth emphasis on spindle cell mimics. American Journal of Surgical Pathology 2001; 25: 1150–7.CrossRefGoogle Scholar
Wesche, WA, Fletcher, CDM, Dias, E, et al. Immunohistochemistry of MyoD1 in adult pleomorphic soft tissue sarcomas. American Journal of Surgical Pathology 1995; 19: 261–9.CrossRefGoogle ScholarPubMed

Selected reference

Flope, AL. MyoD1 and myogenin expression in human neoplasia: a review and update. Advances in Anatomic Pathology 2002; 9: 198203.CrossRefGoogle Scholar

Selected references

Kunishige, M, Mitsui, T, Akaike, M, et al. Localisation and amount of myoglobin and myoglobin mRNA in ragged-red fiber of patients with mitochondrial encephalomyopathy. Muscle and Nerve 1996; 19: 175–82.3.0.CO;2-B>CrossRefGoogle ScholarPubMed
Zhang, JM, Riddick, L. Cytoskeleton immunohistochemical study of early ischemic myocardium. Forensic Science International 1996; 80: 229–38.CrossRefGoogle ScholarPubMed

Selected references

Hirano, T, Gong, Y, Yoshida, K, et al. Usefulness of TA02 (napsin A) to distinguish primary lung adenocarcinoma from metastatic lung adenocarcinoma. Lung Cancer 2003; 41: 155–62.CrossRefGoogle Scholar
Schauer-Vukasinovic, V, Bur, D, Kling, D, Grüninger, F, Giller, T. Human napsin A: expression, immunochemical detection, and tissue localization. FEBS Letters 1999; 462: 135–9.CrossRefGoogle ScholarPubMed
Suzuki, A, Shijubo, N, Yamada, G, et al. Napsin A is useful to distinguish primary lung adenocarcinoma from adenocarcinomas of other organs. Pathology, Research and Practice 2005; 201: 579–86.CrossRefGoogle ScholarPubMed

Selected reference

Gotow, T. Neurofilaments in health and disease. Medical Electron Microscopy 2000; 33: 173–99.CrossRefGoogle ScholarPubMed

Selected references

Ohlsson, K, Olsson, I. The neutral proteases of human granulocytes. Isolation and partial characterization of granulocyte elastases. European Journal of Biochemistry 1974; 42: 519–27.CrossRefGoogle ScholarPubMed
Pulford, KAF, Erber, WN, Crick, JA, et al. Monoclonal antibody against human neutrophil elastase for the study of normal and leukaemic myeloid cells. Journal of Clinical Pathology 1988; 41: 853–60.CrossRefGoogle Scholar

Selected references

Graham, AN, Maxwell, P, Mulholland, K, et al. Increased nm23 immunoreactivity is associated with selective inhibition of systemic tumour cell dissemination. Journal of Clinical Pathology 2002; 55: 184–9.CrossRefGoogle ScholarPubMed
Urano, T, Furukawa, K, Shiku, H. Expression of nm23/NDP kinase proteins on the cell surface. Oncogene 1993; 8: 1371–6.Google ScholarPubMed

Selected references

Bishop, JA, Westra, WH. NUT midline carcinomas of the sinonasal tract. American Journal of Surgical Pathology 2012; 36: 1216–21.CrossRefGoogle ScholarPubMed
French, CA, Miyoshi, I, Kubonishi, I, et al. BRD4-NUT fusion oncogene: a novel mechanism in aggressive carcinoma. Cancer Research 2003; 63: 304–7.Google ScholarPubMed
Haack, H, Johnson, LA, Fry, CJ, et al. Diagnosis of NUT midline carcinoma using a NUT-specific monoclonal antibody. American Journal of Surgical Pathology 2009; 33: 984–91.CrossRefGoogle ScholarPubMed
Stelow, EB, Bellizzi, AM, Taneja, K, et al. NUT rearrangement in undifferentiated carcinomas of the upper aerodigestive tract. American Journal of Surgical Pathology 2008; 32: 828–34.CrossRefGoogle ScholarPubMed

Selected references

Gibson, SE, Dong, HY, Advani, AS, Hsi, ED. Expression of the B cell-associated transcription factors PAX5, OCT-2, and BOB.1 in acute myeloid leukemia: associations with B-cell antigen expression and myelomonocytic maturation. American Journal of Clinical Pathology 2006; 126: 916–24.CrossRefGoogle Scholar
Marafioti, T, Ascani, S, Pulford, K, et al. Expression of B-lymphocyte-associated transcription factors in human T-cell neoplasms. American Journal of Pathology 2003; 162: 861–71.CrossRefGoogle ScholarPubMed
Nasr, MR, Rosenthal, N, Syrbu, S. Expression profiling of transcription factors in B- or T-acute lymphoblastic leukemia/lymphoma and Burkitt lymphoma: usefulness of PAX5 immunostaining as pan-Pre-B-cell marker. American Journal of Clinical Pathology 2010; 133: 41–8.CrossRefGoogle ScholarPubMed

Selected references

Cheng, L, Sung, MT, Cossu-Rocca, P, et al. OCT4: biological functions and clinical applications as a marker of germ cell neoplasia. Journal of Pathology 2007; 211: 19.CrossRefGoogle ScholarPubMed
de Jong, J, Stoop, H, Dohle, GR, et al. Diagnostic value of OCT3/4 for pre-invasive and invasive testicular germ cell tumours. Journal of Pathology 2005; 206: 242–9.CrossRefGoogle ScholarPubMed
Jones, TD, Ulbright, TM, Eble, JN, Baldridge, LA, Cheng, L. OCT4 staining in testicular tumors: a sensitive and specific marker for seminoma and embryonal carcinoma. American Journal of Surgical Pathology 2004; 28: 935–40.CrossRefGoogle ScholarPubMed
Jones, TD, Ulbright, TM, Eble, JN, Cheng, L. OCT4: a sensitive and specific biomarker for intratubular germ cell neoplasia of the testis. Clinical Cancer Research 2004; 10: 8544–7.CrossRefGoogle ScholarPubMed
Looijenga, LH, Stoop, H, de Leeuw, HP, et al. POU5F1 (OCT3/4) identifies cells with pluripotent potential in human germ cell tumors. Cancer Research 2003; 63: 2244–50.Google ScholarPubMed
Sung, MT, Jones, TD, Beck, SD, Foster, RS, Cheng, L. OCT4 is superior to CD30 in the diagnosis of metastatic embryonal carcinomas after chemotherapy. Human Pathology 2006; 37: 662–7.CrossRefGoogle ScholarPubMed

Selected references

Ligon, KL, Alberta, JA, Kho, AT, et al. The oligodendroglial lineage marker OLIG2 is universally expressed in diffuse gliomas. Journal of Neuropathology and Experimental Neurology 2004; 63: 499509.CrossRefGoogle ScholarPubMed
Marie, Y, Sanson, M, Mokhtari, K, et al. OLIG2 as a specific marker of oligodendroglial tumour cells. Lancet 2001; 358: 298300.CrossRefGoogle ScholarPubMed
Popova, SN, Bergqvist, M, Dimberg, A, et al. Subtyping of gliomas of various WHO grades by the application of immunohistochemistry. Histopathology 2014; 64: 365–79.CrossRefGoogle ScholarPubMed
Yokoo, H, Nobusawa, S, Takebayashi, H, et al. Anti-human Olig2 antibody as a useful immunohistochemical marker of normal oligodendrocytes and gliomas. American Journal of Pathology 2004; 164: 1717–25.CrossRefGoogle ScholarPubMed

Selected references

Brown, LF, Papadopoulos-Sergiou, A, Berse, B, et al. Osteopontin expression and distribution in human carcinomas. American Journal of Pathology 1994; 145: 610–23.Google ScholarPubMed
Butler, WT. The nature and significance of osteopontin. Connective Tissue Research 1989; 23: 123–36.CrossRefGoogle ScholarPubMed
Butler, WT. Structural and functional domains of osteopontin. Annals of the New York Academy of Sciences 1995; 760: 611.CrossRefGoogle ScholarPubMed
Denhardt, T, Guo, X. Osteopontin: a protein with diverse functions. FASEB Journal 1993; 7: 475–82.CrossRefGoogle ScholarPubMed

Selected references

Ansari-Lari, MA, Staebler, A, Zaino, RJ, Shah, KV, Ronnett, BM. Distinction of endocervical and endometrial adenocarcinomas: immunohistochemical p16 expression correlated with human papillomavirus (HPV) DNA detection. American Journal of Surgical Pathology 2004; 28: 160–7.CrossRefGoogle ScholarPubMed
Armes, JE, Lourie, R, De Silva, M, et al. Abnormalities of the RB1 pathway in ovarian serous papillary carcinoma as determined by overexpression of the p16INK4A protein. International Journal of Gynecological Pathology 2005; 24: 363–8.CrossRefGoogle Scholar
Bodner-Adler, B, Bodner, K, Czerwenka, K, et al. Expression of p16 protein in patients with uterine smooth muscle tumors: an immunohistochemical analysis. Gynecologic Oncology 2005; 96: 62–6.CrossRefGoogle ScholarPubMed
Dray, M, Russell, P, Dalrymple, C, et al. p16INK4a as a complementary marker of high-grade intraepithelial lesions of the uterine cervix. I: Experience with squamous lesions in 189 consecutive cervical biopsies. Pathology 2005; 37: 112–24.CrossRefGoogle ScholarPubMed
Kalof, AN, Evans, MF, Simmons-Arnold, L, Beatty, BG, Cooper, K. p16INK4A immunoexpression and HPV in situ hybridization signal patterns: potential markers of high-grade cervical intraepithelial neoplasia. American Journal of Surgical Pathology 2005; 29: 674–9.CrossRefGoogle ScholarPubMed
Klaes, R, Friedrich, T, Spitkovsky, D, et al. Overexpression of p16INK4A as a specific marker for dysplastic and neoplastic epithelial cells of the cervix uteri. International Journal of Cancer 2001; 92: 276–84.CrossRefGoogle ScholarPubMed
Qiao, X, Bhuiya, TA, Spitzer, M. Differentiating high-grade cervical intraepithelial lesion from atrophy in postmenopausal women using Ki-67, cyclin E, and p16 immunohistochemical analysis. Journal of Lower Genital Tract Disease 2005; 9: 100–7.CrossRefGoogle ScholarPubMed
Sano, T, Oyama, T, Kashiwabara, K, Fukuda, T, Nakajima, T. Expression status of p16 protein is associated with human papillomavirus oncogenic potential in cervical and genital lesions. American Journal of Pathology 1998; 153: 1741–8.CrossRefGoogle ScholarPubMed

Selected references

Hengst, L, Reed, SI. Translational control of p27kip1 accumulation during the cell cycle. Science 1996; 271: 1861–4.CrossRefGoogle ScholarPubMed
Lloyd, RV, Jin, L, Qian, X, Kulig, E. Aberrant p27kip1 expression in endocrine and other tumors. American Journal of Pathology 1997; 150: 401–7.Google ScholarPubMed
Toyoshima, H, Hunter, T. p27, a novel inhibitor of Gl cyclin-Cdk protein kinase activity is related to p21. Cell 1994; 78: 6774.CrossRefGoogle Scholar

Selected references

Alomari, AK, Glusac, EJ, McNiff, JM. p40 is a more specific marker than p63 for cutaneous poorly differentiated squamous cell carcinoma. Journal of Cutaneous Pathology 2014; 41: 839–45.CrossRefGoogle ScholarPubMed
Bishop, JA, Teruya-Feldstein, J, Westra, WH, et al. p40 (ΔNp63) is superior to p63 for the diagnosis of pulmonary squamous cell carcinoma. Modern Pathology 2012; 25: 405–15.CrossRefGoogle ScholarPubMed
Geddert, H, Kiel, S, Heep, HJ, Gabbert, HE, Sarbia, M. The role of p63 and deltaNp63 (p40) protein expression and gene amplification in esophageal carcinogenesis. Human Pathology 2003; 34: 850–6.CrossRefGoogle ScholarPubMed
Kim, SK, Jung, WH, Koo, JS. p40 (ΔNp63) expression in breast disease and its correlation with p63 immunohistochemistry. International Journal of Clinical and Experimental Pathology 2014; 7: 1032–41.Google ScholarPubMed
Righi, L, Graziano, P, Fornari, A, et al. Immunohistochemical subtyping of nonsmall cell lung cancer not otherwise specified in fine-needle aspiration cytology: a retrospective study of 103 cases with surgical correlation. Cancer 2011; 117: 3416–23.CrossRefGoogle Scholar
Tilson, MP, Bishop, JA. Utility of p40 in the differential diagnosis of small round blue cell tumors of the sinonasal tract. Head and Neck Pathology 2014; 8: 141–5.CrossRefGoogle ScholarPubMed
Zhang, HJ, Xue, WC, Siu, MK, et al. P63 expression in gestational trophoblastic disease: correlation with proliferation and apoptotic dynamics. International Journal of Gynecological Pathology 2009; 28: 172–8.CrossRefGoogle ScholarPubMed

Selected references

Batsakis, JG, El-Naggar, AK. p53: 15 years after discovery. Advances in Anatomic Pathology 1995; 2: 7188.CrossRefGoogle Scholar
Chang, F, Syrjanen, S, Tervahauta, A, Syrjanen, K. Tumorigenesis associated with the p53 tumour suppressor gene. British Journal of Cancer 1993; 68: 653–61.CrossRefGoogle Scholar

Selected references

Di Como, CJ, Urist, MJ, Babayan, I, et al. p63 expression profiles in human normal and tumor tissues. Clinical Cancer Research 2002; 8: 494501.Google ScholarPubMed
Martin, SE, Temm, CJ, Goheen, MP, et al. Cytoplasmic p63 immunohistochemistry is a useful marker for muscle differentiation: an immunohistochemical and immunoelectron microscopic study. Modern Pathology 2011; 24: 1320–6.CrossRefGoogle ScholarPubMed
Weinstein, MH, Signoretti, S, Loda, M. Diagnostic utility of immunohistochemical staining for p63, a sensitive marker of prostatic basal cells. Modern Pathology 2002; 15: 1302–8.CrossRefGoogle ScholarPubMed

Selected reference

Bouchard, S, Russo, P, Radu, AP, Adzick, NS. Expression of neuropeptides in normal and abnormal appendices. Journal of Pediatric Surgery 2001; 36: 1222–6.CrossRefGoogle ScholarPubMed

Selected reference

Cetani, F, Banti, C, Pardi, E, et al. CDC73 mutational status and loss of parafibromin in the outcome of parathyroid cancer. Endocrine Connections 2013; 2: 186–95.CrossRefGoogle ScholarPubMed

Selected references

Aldinger, KA, Hickey, RC, Ibanez, ML, Samaan, NA. Parathyroid carcinoma: a clinical study of seven cases of functioning and two cases of nonfunctioning parathyroid cancer. Cancer 1982; 49: 388–97.3.0.CO;2-F>CrossRefGoogle ScholarPubMed
Brown, EM. PTH secretion in vivo and in vitro: regulation by calcium and other secretagogues. Mineral and Electrolyte Metabolism 1982; 8: 130–50.Google ScholarPubMed
Chen, HL, Demiralp, B, Schneider, A, et al. Parathyroid hormone and parathyroid hormone-related protein exert both pro- and anti-apoptotic effects in mesenchymal cells. Journal of Biological Chemistry 2002; 277: 19, 374–81.Google ScholarPubMed

Selected references

Bouvet, M, Nardin, SR, Burton, DW, et al. Parathyroid hormone-related protein as a novel tumor marker in pancreatic adenocarcinoma. Pancreas 2002; 24: 284–90.CrossRefGoogle ScholarPubMed
Burtis, WJ, Brady, TG, Orloff, JJ, et al. Immunochemical characterization of circulating parathyroid hormone-related protein in patients with humoral hypercalcemia of cancer. New England Journal of Medicine 1990; 322: 1106–12.CrossRefGoogle ScholarPubMed

Selected references

Daniel, L, Lechevallier, E, Giorgi, R, et al. Pax-2 expression in adult renal tumors. Human Pathology 2001; 32: 282–7.CrossRefGoogle ScholarPubMed
Dressler, GR, Douglass, EC. Pax-2 is a DNA-binding protein expressed in embryonic kidney and Wilms tumor. Proceedings of the National Academy of Sciences of the USA 1992; 89: 1179–83.CrossRefGoogle ScholarPubMed
Ordóñez, NG. Value of PAX2 immunostaining in tumor diagnosis: a review and update. Advances in Anatomic Pathology 2012; 19: 401–9.Google ScholarPubMed
Ozcan, A, Liles, N, Coffey, D, Shen, SS, Truong, LD. PAX2 and PAX8 expression in primary and metastatic mullerian epithelial tumors: a comprehensive comparison. American Journal of Surgical Pathology 2011; 35: 1837–47.CrossRefGoogle ScholarPubMed
Rabban, JT, Mcalhany, S, Lerwill, MF, Grenert, JP, Zaloudek, CJ. PAX2 distinguishes benign mesonephric and mullerian glandular lesions of the cervix from endocervical adenocarcinoma, including minimal deviation adenocarcinoma. American Journal of Surgical Pathology 2010; 34: 137–46.CrossRefGoogle ScholarPubMed

Selected references

Baker, SJ, Reddy, EP. B cell differentiation: role of E2A and Pax5/BSAP transcription factors. Oncogene 1995; 11: 413–26.Google ScholarPubMed
Desouki, MM, Post, GR, Cherry, D, Lazarchick, J. PAX-5: a valuable immunohistochemical marker in the differential diagnosis of lymphoid neoplasms. Clinical Medicine and Research 2010; 8: 84–8.CrossRefGoogle ScholarPubMed
Morgenstern, DA, Hasan, F, Gibson, S, et al. PAX5 expression in nonhematopoietic tissues: reappraisal of previous studies. American Journal of Clinical Pathology 2010; 133: 407–15.CrossRefGoogle ScholarPubMed
O'Brien, P, Morin, P, Ouellette, RJ, Robichaud, GA. The Pax-5 gene: a pluripotent regulator of B-cell differentiation and cancer disease. Cancer Research 2011; 71: 7345–50.CrossRefGoogle ScholarPubMed

Selected references

Albadine, R, Schultz, L, Illei, P, et al. PAX8 (+)/p63 (–) immunostaining pattern in renal collecting duct carcinoma (CDC): a useful immunoprofile in the differential diagnosis of CDC versus urothelial carcinoma of upper urinary tract. American Journal of Surgical Pathology 2010; 34: 965–9.CrossRefGoogle ScholarPubMed
Argani, P, Hicks, J, De Marzo, AM, et al. Xp11 translocation renal cell carcinoma (RCC): extended immunohistochemical profile emphasizing novel RCC markers. American Journal of Surgical Pathology 2010; 34: 1295–303.CrossRefGoogle ScholarPubMed
Bowen, NJ, Logani, S, Dickerson, EB, et al. Emerging roles for PAX8 in ovarian cancer and endosalpingeal development. Gynecologic Oncology 2007; 104: 331–7.CrossRefGoogle ScholarPubMed
Chu, PG, Chung, L, Weiss, LM, Lau, SK. Determining the site of origin of mucinous adenocarcinoma: an immunohistochemical study of 175 cases. American Journal of Surgical Pathology 2011; 35: 1830–6.CrossRefGoogle ScholarPubMed
Danialan, R, Assaad, M, Burghardt, J, et al. The utility of PAX8 and IMP3 immunohistochemical stains in the differential diagnosis of benign, premalignant, and malignant endocervical glandular lesions. Gynecologic Oncology 2013; 130: 383–8.CrossRefGoogle ScholarPubMed
Laury, AR, Hornick, JL, Perets, R, et al. PAX8 reliably distinguishes ovarian serous tumors from malignant mesothelioma. American Journal of Surgical Pathology 2010; 34: 627–35.CrossRefGoogle ScholarPubMed
Nonaka, D, Tang, Y, Chiriboga, L, Rivera, M, Ghossein, R. Diagnostic utility of thyroid transcription factors Pax8 and TTF-2 (FoxE1) in thyroid epithelial neoplasms. Modern Pathology 2008; 21: 192200.CrossRefGoogle ScholarPubMed
Sangoi, AR, Fjiwara, M, West, RB, et al. Immunohistochemical distinction of primary adrenal cortical lesions from metastatic clear cell renal cell carcinoma: a study of 248 cases. American Journal of Surgical Pathology 2011; 35: 678–86.CrossRefGoogle ScholarPubMed

Selected references

Carreras, J, Lopez-Guillermo, A, Roncador, G, et al. High numbers of tumor-infiltrating programmed cell death 1-positive regulatory lymphocytes are associated with improved overall survival in follicular lymphoma. Journal of Clinical Oncology 2009; 27: 1470–6.CrossRefGoogle ScholarPubMed
Cetinozman, F, Koens, L, Jansen, PM, Willemze, R. Programmed death-1 expression in cutaneous B-cell lymphoma. Journal of Cutaneous Pathology 2014; 41: 1421.CrossRefGoogle ScholarPubMed
Dorfman, DM, Brown, JA, Shahsafaei, A, Freeman, GJ. Programmed Death-1 (PD-1) is a marker of germinal center-associated T cells and angioimmunoblastic T cell lymphoma. American Journal of Surgical Pathology 2006; 30: 802–10.CrossRefGoogle ScholarPubMed
Muenst, S, Dirnhofer, S, Tzankov, A. Distribution of PD-1+ lymphocytes in reactive lymphadenopathies. Pathobiology 2010; 77: 2427.CrossRefGoogle ScholarPubMed
Muenst, S, Hoeller, S, Willi, N, Dirnhofera, S, Tzankov, A. Diagnostic and prognostic utility of PD-1 in B cell lymphomas. Disease Markers 2010; 29: 4753.CrossRefGoogle ScholarPubMed
Nam-Cha, SH, Roncador, G, Sanchez-Verde, L, et al. PD-1, a follicular T-cell marker useful for recognizing nodular lymphocyte-predominant Hodgkin lymphoma. American Journal of Surgical Pathology 2008; 32: 1252–7.CrossRefGoogle ScholarPubMed

Selected references

Cordon-Cardo, C, O'Brien, JP, Boccia, J, et al. Expression of the multidrug resistance gene product (P-glycoprotein) in human normal and tumor tissues. Journal of Histochemistry and Cytochemistry 1990; 38: 1277–87.CrossRefGoogle ScholarPubMed
Lopes, JM, Bruland, OS, Bjekehagen, B, et al. Synovial sarcoma: Immunohistochemical expression of P-glycoprotein and glutathione S transferase-pi and clinical drug resistance. Pathology Research and Practice 1997; 193: 2136.CrossRefGoogle ScholarPubMed
Scheffer, GL, Pijnenborg, AC, Smit, EF, et al. Multidrug resistance related molecules in human and murine lung. Journal of Clinical Pathology 2002; 55: 332–9.CrossRefGoogle ScholarPubMed

Selected references

Juan, G, Traganos, F, James, WM, et al. Histone H3 phosphorylation and expression of cyclins A and B1 measured in individual cells during their progression through G2 and mitosis. Cytometry 1998; 32: 71–7.3.0.CO;2-H>CrossRefGoogle ScholarPubMed
Ribalta, T, McCutcheon, IE, Aldape, KD, et al. The mitosis-specific antibody anti-phosphohistone-H3 (PHH3) facilitates rapid reliable grading of meningiomas according to WHO 2000 criteria. American Journal of Surgical Pathology 2004; 28: 1532–6.CrossRefGoogle ScholarPubMed
Shibata, K, Ajiro, K. Cell cycle-dependent suppressive effect of histone H1 on mitosis-specific H3 phosphorylation. Journal of Biological Chemistry 1993; 268: 18, 431–4.CrossRefGoogle ScholarPubMed
Tetzlaff, MT, Curry, JL, Ivan, D, et al. Immunodetection of phosphohistone H3 as a surrogate of mitotic figure count and clinical outcome in cutaneous melanoma. Modern Pathology 2013; 26: 1153–60.CrossRefGoogle ScholarPubMed
Voss, SM, Riley, MP, Lokhandwala, PM, Wang, M, Yang, Z. Mitotic count by phosphohistone H3 immunohistochemical staining predicts survival and improves interobserver reproducibility in well-differentiated neuroendocrine tumors of the pancreas. American Journal of Surgical Pathology 2015; 39: 1324.CrossRefGoogle ScholarPubMed

Selected references

Kobayashi, I, Oka, H, Naritaka, H, et al. Expression of Pit-1 and growth hormone-releasing hormone receptor mRNA in human pituitary adenomas: difference among functioning, silent, and other nonfunctioning adenomas. Endocrine Pathology 2002; 13: 8398.CrossRefGoogle ScholarPubMed
Mete, O, Asa, SL. Therapeutic implications of accurate classification of pituitary adenomas. Seminars in Diagnostic Pathology 2013; 30: 158–64.CrossRefGoogle ScholarPubMed
Puy, LA, Asa, SL. The ontogeny of Pit-1 expression in the human fetal pituitary gland. Neuroendocrinology 1996; 63: 349–55.CrossRefGoogle ScholarPubMed

Selected references

Earle, KM, Dillard, SH. Pathology of adenomas of the pituitary gland. Excerpta Medica International Congress Series 1973; 303: 316.Google Scholar
Hardy, J, Vezina, JL. Transsphenoidal neurosurgery of intracranial neoplasm. Advances in Neurology 1976; 15: 261–5.Google ScholarPubMed
Robert, F. Electron microscopy of human pituitary tumors. In Tindall, GT, Collins, WF, eds. Clinical Management of Pituitary Disorders. New York: Raven Press, 1979, pp. 113–31.Google Scholar
Scheithauer, BW. Surgical pathology of the pituitary: the adenomas. Part 1. Pathology Annual 1984; 19: 317–69.Google Scholar

Selected references

Koshida, K, Uchibayashi, T, Yamamoto, H, et al. A potential use of a monoclonal antibody to placental alkaline phosphatase (PLAP) to detect lymph node metastases of seminoma. Journal of Urology 1996; 155: 337–41.CrossRefGoogle ScholarPubMed
Losch, A, Kainz, C. Immunohistochemistry in the diagnosis of the gestational trophoblastic disease. Acta Obstetrica et Gynecologica Scandinavica 1996; 75: 753–6.CrossRefGoogle ScholarPubMed
Manivel, JC, Jessurun, J, Wick, MR, Dehner, LP. Placental alkaline phosphatase immunoreactivity in testicular germ cell neoplasms. American Journal of Surgical Pathology 1987; 11: 21–9.CrossRefGoogle ScholarPubMed
Wick, MR, Swanson, PE, Manivel, JC. Placental-like alkaline phosphatase reactivity in human tumors: an immunohistochemical study of 520 cases. Human Pathology 1987; 18: 946–54.CrossRefGoogle ScholarPubMed

Selected references

Garg, K, Leitao, MM, Kauff, ND, et al. Selection of endometrial carcinomas for DNA mismatch repair protein immunohistochemistry using patient age and tumor morphology enhances detection of mismatch repair abnormalites. American Journal of Surgical Pathology 2009; 33: 925–33.CrossRefGoogle Scholar
Hall, G, Clarkson, A, Shi, A, et al. Immunohistochemistry for PMS2 and MSH6 alone can replace a four antibody panel for mismatch repair deficiency screening in colorectal adenocarcinoma. Pathology 2010; 42: 409–13.CrossRefGoogle ScholarPubMed
Hendriks, YM, Jagmohan-Changur, S, van der Klift, HM, et al. Heterozygous mutations in PMS2 cause hereditary nonpolyposis colorectal carcinoma (Lynch syndrome). Gastroenterology 2006; 130: 312–22.CrossRefGoogle ScholarPubMed
Modica, I, Soslow, RA, Black, D, et al. Utility of immunohistochemistry in predicting microsatellite instability in endometrial carcinoma. American Journal of Surgical Pathology 2007; 31: 744–51.CrossRefGoogle ScholarPubMed
Nicolaides, NC, Papadopoulos, N, Liu, B, et al. Mutations of two PMS homologues in hereditary nonpolyposis colon cancer. Nature 1994; 371: 7580.CrossRefGoogle ScholarPubMed
Shia, JR, Tang, LH, Vakiani, E, et al. Immunohistochemistry as first-line screening for detecting colorectal cancer patients at risk for hereditary nonpolyposis colorectal cancer syndrome A 2-antibody panel may be as predictive as a 4-antibody panel. American Journal of Surgical Pathology 2009; 33: 1639–45.CrossRefGoogle ScholarPubMed
Truninger, K, Menigatti, M, Luz, J, et al. Immunohistochemical analysis reveals high frequency of PMS2 defects in colorectal cancer. Gastroenterology 2005; 128: 1160–71.CrossRefGoogle ScholarPubMed

Selected reference

Amin, MB, Mezger, E, Zarbo, RJ. Detection of Pneumocystis carinii. Comparative study of monoclonal antibody and silver staining. American Journal of Clinical Pathology 1992; 98: 1318.CrossRefGoogle ScholarPubMed

Selected references

Plouzek, CA, Leslie, KK, Stephens, JK, Chou, JY. Differential gene expression in the amnion, chorion, and trophoblast of the human placenta. Placenta 1993; 14: 277–85.CrossRefGoogle ScholarPubMed
Sabet, LM, Daya, D, Stead, R, et al. Significance and value of immunohistochemical localization of pregnancy specific proteins in feto-maternal tissue throughout pregnancy. Modern Pathology 1989; 2: 227–32.Google ScholarPubMed
Waterhouse, R, Ha, C, Dvcksler, GS. Murine CD9 is the receptor for pregnancy-specific glycoprotein 17. Journal of Experimental Medicine 2002; 195: 277–82.CrossRefGoogle ScholarPubMed
Wright, C, Angus, B, Napier, J, et al. Prognostic factors in breast cancer: immunohistochemical staining for SP1 and NCRC 11 related to survival, tumour epidermal growth factor receptor and oestrogen receptor status. Journal of Pathology 1987; 153: 325–31.CrossRefGoogle ScholarPubMed

Selected references

Keshgegian, AA. Biochemically estrogen receptor-negative, progesterone receptor-positive breast carcinoma. Immunocytochemical hormone receptors and prognostic factors. Archives of Pathology and Laboratory Medicine 1994; 118: 240–4.Google ScholarPubMed
Leong, AS-Y, Milios, J. Comparison of antibodies to estrogen and progesterone receptors and the influence of microwave antigen retrieval. Applied Immunohistochemistry 1993; 1: 282–8.Google Scholar
MacGrogan, G, Soubeyran, I, De Mascarei, I, et al. Immunohistochemical detection of progesterone receptors in breast invasive ductal carcinomas: a correlative study of 942 cases. Applied Immunohistochemistry 1996; 4: 219–27.Google Scholar

Selected references

Edwards, YH, Fox, MF, Povey, S, Hinks, LJ. The gene for human neuron specific ubiquitin C-terminal hydrolase (UCHL1, PGP9.5) maps to chromosome 4p14. Annals of Human Genetics 1991; 55: 273–8.CrossRefGoogle ScholarPubMed
Giambanco, I, Bianchi, R, Ceccarelli, P, et al. “Neuron-specific” protein gene product 9.5 (PGP 9.5) is also expressed in glioma cell lines and its expression depends on the cellular growth state. FEBS Letters 1991; 290: 131–4.CrossRefGoogle ScholarPubMed
Gosney, JR, Gosney, MA, Lye, M, Butt, SA. Reliability of commercially available immunocytochemical markers for identification of neuroendocrine differentiation in bronchoscopic biopsies of bronchial carcinoma. Thorax 1995; 50: 116–20.CrossRefGoogle ScholarPubMed
Hibi, K, Westra, WH, Borges, M, et al. PGP9.5 as a candidate tumor marker for non-small-cell lung cancer. American Journal of Pathology 1999; 155: 711–15.CrossRefGoogle ScholarPubMed
Rode, J, Dhillon, AP, Doran, JF, et al. PGP 9.5, a new marker for human neuroendocrine tumors. Histopathology 1985; 9; 147–58.CrossRefGoogle Scholar

Selected references

Bergethon, K, Shaw, AT, Ou, SH, et al. ROS1 rearrangements define a unique molecular class of lung cancers. Journal of Clinical Oncology 2012; 30: 863–70.CrossRefGoogle ScholarPubMed
Charest, A, Lane, K, McMahon, K, et al. Fusion of FIG to the receptor tyrosine kinase ROS in a glioblastoma with an interstitial del(6)(q21q21). Genes, Chromosomes, and Cancer 2003; 37: 5871.CrossRefGoogle Scholar
Gu, TL, Deng, X, Huang, F, et al. Survey of tyrosine kinase signaling reveals ROS kinase fusions in human cholangiocarcinoma. PloS ONE 2011; 6: e15640.CrossRefGoogle ScholarPubMed
Hornick, JL, Sholl, LM, Lovly, CM. Expression of ROS1 predicts ROS1 gene rearrangement in inflammatory myofibroblastic tumors. Modern Pathology 2014; 27: 18A.Google Scholar
Lindeman, NI, Cagle, PT, Beasley, MB, et al. Molecular testing guideline for selection of lung cancer patients for EGFR and ALK tyrosine kinase inhibitors: guideline from the College of American Pathologists, International Association for the Study of Lung Cancer, and Association for Molecular Pathology. Archives of Pathology and Laboratory Medicine 2013; 137: 828–60.CrossRefGoogle Scholar
Sholl, LM, Sun, H, Butaney, M, et al. ROS1 immunohistochemistry for detection of ROS1-rearranged lung adenocarcinomas. American Journal of Surgical Pathology 2013; 37: 1441–9.CrossRefGoogle ScholarPubMed
Yoshida, A, Tsuta, K, Wakai, S, et al. Immunohistochemical detection of ROS1 is useful for identifying ROS1 rearrangements in lung cancers. Modern Pathology 2014; 27: 711–20.CrossRefGoogle ScholarPubMed

Selected references

May, FE, Westley, BR. Trefoil proteins: their role in normal and malignant cells. Journal of Pathology 1997; 183: 47.3.0.CO;2-5>CrossRefGoogle ScholarPubMed
Poulsom, R. Trefoil peptides. Baillières Clinical Gastroenterology 1996; 10: 113–34.Google ScholarPubMed
Wysocki, SJ, Iacopetta, BJ, Ingram, DM. Prognostic significance of pS2 mRNA in breast cancer. European Journal of Cancer 1994; 30A: 1882–4.Google ScholarPubMed

Selected reference

Bostwick, DG. Prostate-specific antigen: current role in diagnostic pathology of prostatic cancer. American Journal of Clinical Pathology 1994; 102 (Suppl 1): S31–7.Google Scholar

Selected reference

Epstein, JI. PSA and PAP as immunohistochemical markers in prostatic cancer. Urologic Clinics of North America 1993; 20: 757–70.CrossRefGoogle Scholar

Selected references

Djordjevic, B, Hennessy, BT, Li, J, et al. Clinical assessment of PTEN loss in endometrial carcinoma: immunohistochemistry outperforms gene sequencing. Modern Pathology 2012; 25: 699708.CrossRefGoogle ScholarPubMed
Foo, WC, Rashid, A, Wang, H, et al. Loss of phosphatase and tensin homolog expression is associated with recurrence and poor prognosis in patients with pancreatic ductal adenocarcinoma, Human Pathology 2013; 44: 1024–30.CrossRefGoogle ScholarPubMed
Govender, D, Chetty, R. Gene of the month: PTEN. Journal of Clinical Pathology 2012; 65: 601–3.Google ScholarPubMed
McMenamin, ME, Soung, P, Perera, S, et al. Loss of PTEN expression in paraffin-embedded primary prostate cancer correlates with high Gleason score and advanced stage. Cancer Research 1999; 59: 4291–6.Google ScholarPubMed
Perren, A, Weng, LP, Boag, AH, et al. Immunohistochemical evidence of loss of PTEN expression in primary ductal adenocarcinomas of the breast. American Journal of Pathology 1999; 155: 1253–60.CrossRefGoogle ScholarPubMed

Selected references

Feiden, W, Feiden, U, Gerhard, L, et al. Rabies encephalitis: immunohistochemical investigations. Clinical Neuropathology 1985; 4: 156–64.Google ScholarPubMed
Jogai, S, Radotra, BD, Banerjee, AK. Immunohistochemical study of human rabies. Neuropathology 2000; 20: 197203.CrossRefGoogle ScholarPubMed

Selected reference

Cordon-Cardo, C, Richon, VM. Expression of the retinoblastoma protein is regulated in normal human tissues. American Journal of Pathology 1994; 144: 500–10.Google ScholarPubMed

Selected references

Daimaru, Y, Hashimoto, H, Enjoji, M. Malignant peripheral nerve sheath tumours (malignant schwannomas). An immunohistochemical study of 29 cases. American Journal of Surgical Pathology 1985; 9: 434–44.CrossRefGoogle ScholarPubMed
Loeffel, SC, Gillespie, GY, Mirmiran, SA, et al. Cellular immunolocalisation of S100 protein within fixed tissue sections by monoclonal antibodies. Archives of Pathology and Laboratory Medicine 1985; 109: 117–22.Google Scholar
Nakajima, T, Watanabe, S, Sato, Y, et al. An immunoperoxidase study of S100 protein distribution in normal and neoplastic tissues. American Journal of Surgical Pathology 1982; 6: 715–27.CrossRefGoogle ScholarPubMed
Takahashi, K, Isobe, T, Ohtsuki, Y, et al. Immunohistochemical study on the distribution of alpha and beta subunits of S-100 protein in human neoplasm and normal tissues. Virchows Archiv B. Cell Pathology Including Molecular Pathology 1984; 45: 385–96.CrossRefGoogle Scholar

Selected references

Cao, D, Guo, S, Allan, RW, Molberg, KH, Peng, Y. SALL4 is a novel sensitive and specific marker of ovarian primitive germ cell tumors and is particularly useful in distinguishing yolk sac tumor from clear cell carcinoma. American Journal of Surgical Pathology 2009; 33: 894904.CrossRefGoogle ScholarPubMed
Cao, D, Humphrey, PA, Allan, RW. SALL4 is a novel sensitive and specific marker for metastatic germ cell tumors, with particular utility in detection of metastatic yolk sac tumors. Cancer 2009; 115: 2640–51.CrossRefGoogle ScholarPubMed
Cao, D, Li, J, Guo, CC, Allan, RW, Humphrey, PA. SALL4 is a novel diagnostic marker for testicular germ cell tumors. American Journal of Surgical Pathology 2009; 33: 1065–77.CrossRefGoogle ScholarPubMed
Liu, A, Cheng, L, Du, J, et al. Diagnostic utility of novel stem cell markers SALL4, OCT4, NANOG, SOX2, UTF1, and TCL1 in primary mediastinal germ cell tumors. American Journal of Surgical Pathology 2010; 34: 697706.CrossRefGoogle ScholarPubMed
Miettinen, M, Wang, Z, McCue, PA, et al. SALL4 expression in germ cell and non-germ cell tumors: a systematic immunohistochemical study of 3215 cases. American Journal of Surgical Pathology 2014; 38: 410–20.CrossRefGoogle ScholarPubMed
Ushiku, T, Shinozaki, A, Shibahara, J, et al. SALL4 represents fetal gut differentiation of gastric cancer, and is diagnostically useful in distinguishing hepatoid gastric carcinoma from hepatocellular carcinoma. American Journal of Surgical Pathology 2010; 34: 533–40.CrossRefGoogle ScholarPubMed
Wang, F, Liu, A, Peng, Y, et al. Diagnostic utility of SALL4 in extragonadal yolk sac tumors: an immunohistochemical study of 59 cases with comparison to placental-like alkaline phosphatase, alpha-fetoprotein, and glypican-3. American Journal of Surgical Pathology 2009; 33: 1529–39.CrossRefGoogle ScholarPubMed

Selected references

Gill, AJ, Benn, DE, Chou, A, et al. Immunohistochemistry for SDHB triages genetic testing of SDHB, SDHC, and SDHD in paraganglioma–pheochromocytoma syndromes. Human Pathology 2010; 41: 805–14.CrossRefGoogle ScholarPubMed
Gill, AJ, Toon, CW, Clarkson, A, et al. Succinate dehydrogenase deficiency is rare in pituitary adenomas. American Journal of Surgical Pathology 2014; 38: 560–66.CrossRefGoogle ScholarPubMed
Miettinen, M, Killian, JK, Wang, ZF, et al. Immunohistochemical loss of succinate dehydrogenase subunit A (SDHA) in gastrointestinal stromal tumors (GISTs) signals SDHA germline mutation. American Journal of Surgical Pathology 2013; 37: 234–40.CrossRefGoogle ScholarPubMed

Selected reference

Burke, AP, Thomas, RM, Elsayed, AM, Sobin, LH. Carcinoids of the jejunum and ileum: an immunohistochemical and clinicopathologic study of 167 cases. Cancer 1997; 79: 1086–93.3.0.CO;2-E>CrossRefGoogle ScholarPubMed

Selected references

Baldi, A, Groeger, AM, Esposito, V, et al. Expression of p21 in SV40 large T antigen positive human pleural mesothelioma: relationship with survival. Thorax 2002; 57: 353–6.CrossRefGoogle ScholarPubMed
Carbone, M, Rizzo, P, Grimley, PM, et al. Simian virus-40 large-T antigen binds p53 in human mesotheliomas. Nature Medicine 1997; 3: 908–12.CrossRefGoogle ScholarPubMed

Selected references

Biegel, JA. Molecular genetics of atypical teratoid/rhabdoid tumor. Neurosurgical Focus 2006; 20: E11.CrossRefGoogle ScholarPubMed
Bourdeaut, F, Lequin, D, Brugières, L, et al. Frequent hSNF5/INI1 germline mutations in patients with rhabdoid tumor. Clinical Cancer Research 2011; 17: 31–8.CrossRefGoogle ScholarPubMed
Calderaro, J, Moroch, J, Pierron, G, et al. SMARCB1/INI1 inactivation in renal medullary carcinoma. Histopathology 2012; 61: 428–35.CrossRefGoogle ScholarPubMed
Hollmann, TJ, Hornick, JL. INI1-deficient tumors: diagnostic features and molecular genetics. American Journal of Surgical Pathology 2011; 35: e4763.CrossRefGoogle ScholarPubMed
Hornick, JL, Dal Cin, P, Fletcher, CD. Loss of INI1 expression is characteristic of both conventional and proximal-type epithelioid sarcoma.American Journal of Surgical Pathology 2009; 33: 542–50.CrossRefGoogle ScholarPubMed
Sullivan, LM, Folpe, AL, Pawel, BR, Judkins, AR, Biegel, JA. Epithelioid sarcoma is associated with a high percentage of SMARCB1 deletions. Modern Pathology 2013; 26: 385–92.CrossRefGoogle ScholarPubMed

Selected references

Borrione, AC, Zanellato, AM, Scannapieco, G, et al. Myosin heavy-chain isoforms in adult and developing rabbit vascular smooth muscle. European Journal of Biochemistry 1989; 183: 413–17.CrossRefGoogle ScholarPubMed
Eddinger, TJ, Murphy, RA. Developmental changes in actin and myosin heavy chain isoform expression in smooth muscle. Archives of Biochemistry and Biophysics 1991; 284: 232–7.CrossRefGoogle ScholarPubMed
Savera, AT, Gown, AM, Zarbo, RJ. Immunolocalization of three novel smooth muscle-specific proteins in salivary gland pleomorphic adenoma: assessment of the morphogenetic role of myoepithelium. Modern Pathology 1997; 10: 1093–100.Google ScholarPubMed
Wang, NP, Wan, BC, Skelly, M, et al. Antibodies to novel myoepithelium-associated proteins distinguish benign lesions and carcinoma in situ from invasive carcinoma of the breast. Applied Immunohistochemistry 1997; 5: 141–51.CrossRefGoogle Scholar
White, S. Martin, AG, Periasamy, M. Identification of a novel smooth muscle myosin heavy chain cDNA: isoform diversity in the SI head region. American Journal of Physiology 1993; 264: 1252–8.CrossRefGoogle Scholar
Yaziji, H, Gown, AM, Sneige, N. Detection of stromal invasion in breast cancer: The myoepithelial markers. Advances in Anatomic Pathology 2000; 7: 100–9.CrossRefGoogle ScholarPubMed

Selected reference

Wehrli, BM, Huang, W, De Crombrugghe, B, et al. Sox9, a master regulator of chondrogenesis, distinguishes mesenchymal chondrosarcoma from other small blue round cell tumors. Human Pathology 2003; 34: 3263–9.CrossRefGoogle ScholarPubMed

Selected references

Kwon, AY, Heo, I, Lee, HJ, et al. Sox10 expression in ovarian epithelial tumors is associated with poor overall survival. Virchows Archiv 2016: 468: 597–60.CrossRefGoogle ScholarPubMed
Nonaka, D, Chiriboga, L, Rubin, BP. Sox10: a pan-schwannian and melanocytic marker. American Journal of Surgical Pathology 2008; 32: 1291–8.CrossRefGoogle ScholarPubMed

Selected reference

Bennett, V. The spectrinactin junction of erythrocyte membrane skeletons. Biochemica et Biophysica Acta 1989; 988: 107–22.Google ScholarPubMed

Selected references

Braidotti, P, Cigala, C, Graziani, D, et al. Surfactant protein A expression in human normal and neoplastic breast epithelium. American Journal of Clinical Pathology 2001; 116: 721–8.CrossRefGoogle ScholarPubMed
Yousem, SA, Wick, MR, singh, G, et al. So-called sclerosing hemangioma of lung. An immunohistochemical study supporting a respiratory epithelial origin. American Journal of Surgical Pathology 1988; 12: 582–90.Google Scholar

Selected references

Chejfec, G, Falkmer, S, Grimelius, L, et al. Synaptophysin: a new marker for pancreatic neuroendocrine tumors. American Journal of Surgical Pathology 1987; 11: 241–7.CrossRefGoogle ScholarPubMed
Gould, VE, Lee, I, Wiedenmann, B, et al. Synaptophysin: a novel marker for neurons, certain neuroendocrine cells, and their neoplasms. Human Pathology 1986; 17: 979–83.CrossRefGoogle ScholarPubMed

Selected references

Cork, LC, Sternberger, NH, Sternberger, LA, et al. Phosphorylated neurofilament antigens in neurofibrillary tangles in Alzheimer's disease. Journal of Neuropathology and Experimental Neurology 1986; 45: 5664.CrossRefGoogle ScholarPubMed
Feany, MB, Dickson, DW. Neurodegenerative disorders with extensive tau pathology: a comparative study and review. Annals of Neurology 1996; 40: 139–48.CrossRefGoogle Scholar
Joachim, CL, Morris, JH, Kosik, KS, Selkoe, DJ. Tau antisera recognize neurofibrillary tangles in a range of neurodegenerative disorders. Annals of Neurology 1987; 22: 514–20.CrossRefGoogle Scholar

Selected references

Ozsan, N, Feldman, AF, Caron, BL, et al. A new immunohistochemistry method to detect T-cell receptor gamma-chain expression in paraffin-embedded biopsies identifies a unique set of peripheral T-cell lymphomas co-expressing T-cell receptor beta and gamma chains. Modern Pathology 2011; 24: 313a.Google Scholar
Rodriguez-Pinilla, SA, Ortiz-Romero, PL, Monsalvez, V, et al. TCR-γ expression in primary cutaneous T cell lymphomas. American Journal of Surgical Pathology 2013; 37: 375384.CrossRefGoogle ScholarPubMed

Selected references

Erickson, HP, Bourdon, MA. Tenascin: an extracellular matrix protein prominent in specialized embryonic tissues and tumors. Annual Review of Cell Biology 1989; 5: 7192.CrossRefGoogle ScholarPubMed
Koukoulis, GK, Gould, VE, Bhattacharyya, A, et al. Tenascin in normal, reactive, hyperplastic and neoplastic tissues: biologic and pathologic implications. Human Pathology 1991; 22: 636–43.CrossRefGoogle ScholarPubMed
Sedele, M, Karaveli, S, Pestereli, HE, et al. Tenascin expression in normal, hyperplastic and neoplastic endometrium. International Journal of Gynecological Pathology 2002; 12: 161–6.Google Scholar

Selected references

Chilosi, M, Pizzolo, G. Review of terminal deoxynucleotidy transferase. Biological aspects, methods of detection, and selected diagnostic applications. Applied Immunohistochemistry 1995; 3: 209–21.Google Scholar
Onciu, M, Lorsbach, RB, Henry, EC, Behm, FG. Terminal deoxynucleotidyl transferase-positive lymphoid cells in reactive lymph nodes from children with malignant tumor: incidence, distribution pattern, and immunophenotype in 26 patients. American Journal of Clinical Pathology 2002; 118: 248–54.CrossRefGoogle ScholarPubMed
Orazi, A, Cattoretti, G, Joh, K, Neiman, RS. Terminal deoxynucleotidyl transferase staining of malignant lymphomas in paraffin sections. Modern Pathology 1994; 7: 582–6.Google ScholarPubMed

Selected references

Argani, P, Aulmann, S, Illei, PB, et al. A distinctive subset of PEComas harbors TFE3 gene fusions. American Journal of Surgical Pathology 2010; 34: 1395–406.CrossRefGoogle ScholarPubMed
Argani, P, Lal, P, Hutchinson, B, et al. Aberrant nuclear immunoreactivity for TFE3 in neoplasms with TFE3 gene fusions: a sensitive and specific immunohistochemical assay. American Journal of Surgical Pathology 2003; 27: 750–61.CrossRefGoogle ScholarPubMed
Argani, P, Olgac, S, Tickoo, SK, et al. Xp11 translocation renal cell carcinoma in adults: expanded clinical, pathologic, and genetic spectrum. American Journal of Surgical Pathology 2007; 31: 1149–60.CrossRefGoogle ScholarPubMed
Dickson, BC, Brooks, JS, Pasha, TL, Zhang, PJ. TFE3 expression in tumors of the microphthalmia-associated transcription factor (MiTF) family. International Journal of Surgical Pathology 2011; 19: 2630.CrossRefGoogle ScholarPubMed
Inamura, K, Fujiwara, M, Togashi, Y, et al. Diverse fusion patterns and heterogeneous clinicopathologic features of renal cell carcinoma with t(6; 11) translocation. American Journal of Surgical Pathology 2012; 36: 3542.CrossRefGoogle Scholar
Lazar, AJ, Lahat, G, Myers, SE, et al. Validation of potential therapeutic targets in alveolar soft part sarcoma: an immunohistochemical study utilizing tissue microarray. Histopathology 2009 55: 750–5.CrossRefGoogle ScholarPubMed
Reis, H, Hager, T, Wohlschlaeger, J, et al. Mammalian target of rapamycin pathway activity in alveolar soft part sarcoma. Human Pathology 2013; 44: 2266–74.CrossRefGoogle ScholarPubMed

Selected references

Appleton, MAC, Attanoos, RL, Jasani, B. Thrombomodulin as a marker of vascular and lymphatic tumors. Histopathology 1996; 29: 153–7.CrossRefGoogle Scholar
Attanoos, RL, Goddard, H, Gibbs, AR. Mesothelioma-binding antibodies: thrombomodulin, OV632 and HBME-1 and their use in the diagnosis of malignant mesothelioma. Histopathology 1996; 29: 209–15.CrossRefGoogle Scholar
Collins, CL, Ordonez, NG, Schaefer, R, et al. Thrombomodulin expression in malignant pleural mesothelioma and pulmonary adenocarcinoma. American Journal of Pathology 1992; 141: 827–33.Google ScholarPubMed
Ordonez, NG. Value of thrombomodulin immunostaining in the diagnosis of mesothelioma. Histopathology 1997; 31: 2530.CrossRefGoogle ScholarPubMed

Selected references

De Micco, C, Ruf, J, Carayon, P, et al. Immunohistochemical study of thyroglobulin in thyroid carcinomas with monoclonal antibodies. Cancer 1987; 59: 471–6.3.0.CO;2-S>CrossRefGoogle ScholarPubMed
Wilson, NW, Pambakian, H, Richardson, TC, et al. Epithelial markers in thyroid carcinoma: an immunoperoxidase study. Histopathology 1986; 10: 815–29.CrossRefGoogle ScholarPubMed

Selected references

Chen, G, Courey, AJ. Groucho/TLE family proteins and transcriptional repression. Gene 2000; 249: 116.CrossRefGoogle ScholarPubMed
Foo, WC, Cruise, MW, Wick, MR, Hornick, JL. Immunohistochemical staining for TLE1 distinguishes synovial sarcoma from histologic mimics. American Journal of Clinical Pathology 2011; 135: 839–44.CrossRefGoogle ScholarPubMed
Jagdis, A, Rubin, BP, Tubbs, RR, Pacheco, M, Nielsen, TO. Prospective evaluation of TLE1 as a diagnostic immunohistochemical marker in synovial sarcoma. American Journal of Surgical Pathology 2009; 33: 1743–51.CrossRefGoogle ScholarPubMed
Kosemehmetoglu, K, Vrana, JA, Folpe, AL. TLE1 expression is not specific for synovial sarcoma: a whole section study of 163 soft tissue and bone neoplasms. Modern Pathology 2009; 22: 872–8.CrossRefGoogle Scholar
Terry, J, Saito, T, Subramanian, S, et al. TLE1 as a diagnostic immunohistochemical marker for synovial sarcoma emerging from gene expression profiling studies. American Journal of Surgical Pathology 2007; 31: 240–6.CrossRefGoogle ScholarPubMed

Selected references

Di Leo, A, Larsimont, D, Gancberg, D, et al. HER-2 and topo-isomerase IIα as predictive markers in a population of node-positive breast cancer patients randomly treated with adjuvant CMF or epirubicin plus cyclophosphamide. Annals of Oncology 2001; 12: 1081–9.CrossRefGoogle ScholarPubMed
Gotleib, WH, Goldberg, I, Weisz, B, et al. Topoisomerase II immunostaining as a prognostic marker for survival in ovarian cancer. Gynecologic Oncology 2001; 82: 99104.CrossRefGoogle Scholar

Selected references

Hasselblatt, M, Böhm, C, Tatenhorst, L, et al. Identification of novel diagnostic markers for choroid plexus tumors: a microarray-based approach. American Journal of Surgical Pathology 2006; 30: 6674.CrossRefGoogle ScholarPubMed
Megerian, CA, Pilch, BZ, Bhan, AK, McKenna, MJ. Differential expression of transthyretin in papillary tumors of the endolymphatic sac and choroid plexus. Laryngoscope 1997; 107: 216–21.CrossRefGoogle ScholarPubMed

Selected references

Hofbauer, GF, Kamarashev, J, Geertsen, R, et al. Tyrosinase immunoreactivity in formalin-fixed, paraffin-embedded primary and metastatic melanoma: frequency and distribution. Journal of Cutaneous Pathology 1998; 25: 204–9.CrossRefGoogle ScholarPubMed
Jungbluth, AA, Iversen, K, Copian, K, et al. T311: an anti-tyrosinase monoclonal antibody for the detection of melanocytic lesions in paraffin embedded tissues. Pathology, Research and Practice 2000; 196: 235–42.CrossRefGoogle ScholarPubMed
Kaufmann, O, Koch, S, Burghardt, J, et al. Tyrosinase, Melan-A, and KBA62 as markers for the immunohistochemical identification of metastatic amelanotic melanomas on paraffin sections. Modern Pathology 1998; 11: 740–6.Google ScholarPubMed

Selected references

Ceccamea, A, Carlei, F, Dominici, C, et al. Correlation between tyrosine hydroxylase immunoreactive cells in tumors and urinary catecholamine output in neuroblastoma patients. Tumori 1986; 72: 451–7.CrossRefGoogle ScholarPubMed
Iwase, K, Nagasaka, A, Nagatsu, I, et al. Tyrosine hydroxylase indicates cell differentiation of catecholamine biosynthesis in neuroendocrine tumors. Journal of Endocrinological Investigation 1994; 17: 235–9.CrossRefGoogle ScholarPubMed
Meijer, WG, Copray, SC, Hollema, H, et al. Catecholamine-synthesizing enzymes in carcinoid tumors and pheochromocytomas. Clinical Chemistry 2003; 49: 586–93.CrossRefGoogle ScholarPubMed
Takahashi, H, Wakabayashi, K, Ikuta, F, Tanimura, K. Esthesioneuroblastoma: a nasal catecholamine-producing tumor of neural crest origin. Demonstration of tyrosine hydroxylase-immunoreactive tumor cells. Acta Neuropathologica 1988; 76: 522–7.CrossRefGoogle ScholarPubMed

Selected reference

Ciechanover, A, Schwartz, AL. The ubiquitin-mediated proteolytic pathway: mechanisms of recognition of the proteolytic substrate and involvement in the degradation of native cellular proteins. FASEB Journal 1994; 8: 182–91.CrossRefGoogle ScholarPubMed

Selected references

Holthofer, H, Virtanen, I, Kariniemi, AL, et al. Ulex europaeus I lectin as a marker for vascular endothelium in human tissues. Laboratory Investigation 1982; 47: 60–6.Google ScholarPubMed
Meittinen, M, Holthofer, H, Lehto, VP, Miettinen, A, Virtanen, I. Ulex europaeus I lectin as a marker for tumors derived from endothelial cells. American Journal of Clinical Pathology 1983; 79: 32–6.Google Scholar
Ordonez, NG, Batsakis, JG. Comparison of Ulex europaeus I lectin and factor VII-related antigen in vascular lesions. Archives of Pathology and Laboratory Medicine 1984; 108: 129–32.Google Scholar

Selected references

Mattern, J, Koomagi, R, Volm, M. Association of vascular endothelial growth factor expression with intratumoral microvessel density and tumour cell proliferation in human epidermoid lung carcinoma. British Journal of Cancer 1996; 72: 931–4.Google Scholar
Salven, P, Keikkila, P, Anttonen, A, et al. Vascular endothelial growth factor in squamous cell head and neck carcinoma: expression and prognostic significance. Modern Pathology 1997; 10: 1128–33.Google ScholarPubMed
Toi, M, Inada, K, Suzuki, H, et al. Tumor angiogenesis in breast cancer: its importance as a prognostic indicator and the association with vascular endothelial growth factor expression. Breast Cancer Research and Treatment 1995; 36: 193204.CrossRefGoogle ScholarPubMed
Zymed Laboratories. Polyclonal rabbit anti-VEGF (data sheet).Google Scholar

Selected reference

Bacchi, CE, Gown, AM. Distribution and pattern of expression of villin, a gastrointestinal-associated cytoskeletal protein, in human carcinomas: a study employing paraffin-embedded tissue. Laboratory Investigation 1991; 64: 418–24.Google ScholarPubMed

Selected reference

Azumi, N, Battifora, H. The distribution of vimentin and keratin in epithelial and non-epithelial neoplasms. American Journal of Clinical Pathology 1987; 88: 286–97.CrossRefGoogle Scholar

Selected references

Banham, AH, Turley, H, Pulford, K, et al. The plasma cell associated antigen detectable by antibody VS38 is the p63 rough endoplasmic reticulum protein. Journal of Clinical Pathology 1997; 50: 485–9.CrossRefGoogle ScholarPubMed
Turley, H, Jones, M, Erber, W, et al. VS38: a new monoclonal antibody for detecting plasma cell differentiation in routine sections. Journal of Clinical Pathology 1994; 47: 418–22.CrossRefGoogle ScholarPubMed

Selected references

Barnoud, R, Sabourin, JC, Pasquier, D, et al. Immunohistochemical expression of WT1 by desmoplastic small round cell tumor: a comparative study with other small round cell tumors. American Journal of Surgical Pathology 2000; 24: 830–6.CrossRefGoogle ScholarPubMed
Charles, AK, Moore, IE, Berry, PJ. Immunohistochemical detection of the Wilms’ tumour gene WT1 in desmoplastic small round cell tumour. Histopathology 1997; 30: 312–14.CrossRefGoogle ScholarPubMed
Foster, MR, Johnson, JE, Olson, SJ, Allred, DC. Immunohistochemical analysis of nuclear versus cytoplasmic staining of WT1 in malignant mesotheliomas and primary pulmonary adenocarcinomas. Archives of Pathology and Laboratory Medicine 2001; 125: 1316–20.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×