• No results found

Renal-Cell Carcinoma

N/A
N/A
Protected

Academic year: 2021

Share "Renal-Cell Carcinoma"

Copied!
14
0
0

Loading.... (view fulltext now)

Full text

(1)

The n e w e n g l a n d j o u r n a l of m e d i c i n e

r e v i e w a r t i c l e

m e d i c a l p r o g r e s s

Renal-Cell Carcinoma

Herbert T. Cohen, M.D., and Francis J. McGovern, M.D.

From the Renal and Hematology–Oncolo-gy Sections, Departments of Medicine and Pathology, Boston University School of Medicine (H.T.C.); and the Department of Urology, Massachusetts General Hospital and Harvard Medical School (F.J.M.) — all in Boston. Address reprint requests to Dr. Cohen at the Department of Medicine, Renal Section, Boston University School of Medicine, Evans Biomedical Research Cen-ter, 650 Albany St., Rm. X-535, Boston, MA 02118, or at htcohen@bu.edu.

N Engl J Med 2005;353:2477-90.

Copyright © 2005 Massachusetts Medical Society. n the united states, renal cancer is the 7th leading malignant

condition among men and the 12th among women, accounting for 2.6 percent of all cancers.1 About 2 percent of cases of renal cancer are associated with inherited syndromes. In the United States, 36,160 new cases of renal cancer are predicted to oc-cur in 2005, many of which are being discovered earlier because of the widespread availability of radiographic testing. Nevertheless, 12,660 deaths from the disease are predicted to occur in 2005.1 Renal-cell carcinomas arise from the renal epithelium and account for about 85 percent of renal cancers. A quarter of the patients present with ad-vanced disease, including locally invasive or metastatic renal-cell carcinoma. Moover, a third of the patients who undergo resection of localized disease will have a re-currence. Median survival for patients with metastatic disease is about 13 months. Thus, there is a great need for more effective surgical and medical therapies.

The classic presentation of renal-cell carcinoma includes the triad of flank pain, hema-turia, and a palpable abdominal mass. Few patients now present in this manner. Roughly half the cases are now detected because a renal mass is incidentally identified on radiographic examination. Other common presenting features may be nonspecific, such as fatigue, weight loss, or anemia. Risk factors for renal-cell carcinoma include smoking, obesity, and hypertension,2 as well as acquired cystic kidney disease associ-ated with end-stage renal disease. A 1.6:1.0 male predominance exists,1 and the peak incidence is in the sixth and seventh decades. Gross or microscopic hematuria is an im-portant clinical clue to the diagnosis of renal-cell carcinoma; thus, hematuria should be evaluated promptly by a computed tomographic (CT) scan of the genitourinary tract and, in patients older than 40 years of age, by cystoscopy to rule out bladder cancer. Prognosis is closely related to the stage of disease (Fig. 1). The Heidelberg classifica-tion of renal tumors was introduced in 19976 as a means of more completely correlat-ing the histopathological features with the identified genetic defects (Table 1).

c o n v e n t i o n a l o r c l e a r - c e l l r e n a l - c e l l c a r c i n o m a

Von Hippel–Lindau disease (number 193300 in Mendelian Inheritance in Man [MIM]) is a rare, autosomal dominant, familial cancer syndrome consisting chiefly of retinal angiomas, hemangioblastomas of the central nervous system, pheochromocytomas, and renal-cell carcinoma of the clear-cell type (Fig. 2). The von Hippel–Lindau tumor-suppressor gene (VHL) was identified in 1993.7 In this disease, one VHL allele is inher-ited with a mutation. Associated focal lesions, such as renal-cell carcinoma, arise from the inactivation or silencing of the remaining normal (wild-type) VHL allele (Fig. 3). Re-markably, defects in the VHL gene also appear to be responsible for about 60 percent of

i

o v e r v i e w

(2)

The n e w e n g l a n d j o u r n a l of m e d i c i n e

the cases of sporadic clear-cell renal-cell carcino-ma,8 which represents a major portion of all cases of renal-cell carcinoma.

VHL protein, the product of the VHL gene, func-tions as a tumor suppressor, inhibiting growth when reintroduced into cultures of renal-cell carci-noma.9,10 Hypoxia-inducible genes are normally in-hibited by VHL protein,11 including several encod-ing proteins involved in angiogenesis (e.g., vascular endothelial growth factor [VEGF]), cell growth (e.g., transforming growth factor a [TGF-a]), glu-cose uptake (e.g., the GLUT-1 gluglu-cose transporter), and acid–base balance (e.g., carbonic anhydrase IX [CA9]). When VHL protein is lost, these proteins are overexpressed, creating a microenvironment favor-able for epithelial-cell proliferation (Fig. 4A). Thus, cells deficient in VHL protein behave as if they are hy-poxic, even in conditions of normoxia. VHL protein, with elongin proteins C and B, binds cul2 protein (a member of the cullin family of ubiquitin ligase pro-teins), indicating that some VHL protein serves as the receptor subunit of a ubiquitin ligase complex

that promotes the ubiquitination and destruction of proteins (Fig. 4B).12,13 VHL protein binds the transcriptional activators hypoxia-inducible factor 1a (HIF-1a) and 2a (HIF-2a) directly and destabiliz-es them.14 Furthermore, VHL protein promotes the ubiquitination and destruction of HIF-a.15-17 These VHL-regulated pathways are being studied as po-tential targets of therapies for clear-cell renal-cell carcinoma.

HIF is the key regulator of the hypoxic response in multicellular organisms. Thus, VHL protein has a central role in oxygen sensing. For HIF-a to bind VHL protein, a proline residue must undergo hy-droxylation, which is an unusual protein modifi-cation18,19 (Fig. 4B). A family of proline hydroxy-lases operates on HIF-a in a graded fashion, so that the extent of hydroxylation depends on oxygen ten-sion.20,21 Hydroxylation of an asparagine residue blocks the interaction of HIF-a with the transcrip-tional coactivator p300.22 Thus, multiple hydroxyl-ation steps cooperate to inhibit HIF-a activity.

To correlate the genotype with the disease

phe-Figure 1. Staging Overview and Five-Year Survival Rates for Renal Cancer.

Survival data3 are based on the 1997 tumor–node–metastasis (TNM) staging guidelines.4 More recent renal-cancer

(3)

m e d i c a l p r o g r e s s

notype, naturally occurring VHL mutations have been evaluated to determine their effect on HIF-a ubiquitination. An intriguing finding is that the

VHL mutations that disrupt HIF-a processing are the same as those associated with the vascular man-ifestations of von Hippel–Lindau disease, such as hemangioblastoma (Fig. 2).15,16,23,24 Since renal-cell carcinoma develops in only a subgroup of pa-tients with hemangioblastoma, the overexpression of HIF-a appears to be necessary for, but not suf-ficient to induce, renal tumorigenesis. Neverthe-less, HIF-a is vitally important to the pathogenesis of this disease. VHL-induced inhibition of HIF-a is sufficient to suppress the growth of clear-cell renal-cell carcinoma in preclinical models.25,26 The cell-matrix protein fibronectin,27 chaperonin TRiC/ CCT,28 microtubules,29 and transcription factor Jade-130-32 are all molecules that interact with VHL protein in a manner that is dependent on VHL mu-tation, suggesting that they may also contribute to disease pathogenesis.

Distinct from von Hippel–Lindau disease, famil-ial clear-cell renal cancer has been reported in pa-tients with translocations of chromosome 3p at a fragile site at 3p14.33 Loss of the translocated chro-mosome 3p probably implicates VHL protein in the development of these tumors. Additional trans-locations of chromosome 3 have been associated with clear-cell renal-cell carcinoma as well.

p a p i l l a r y r e n a l - c e l l c a r c i n o m a

Sporadic papillary renal-cell carcinoma has a five-year survival rate approaching 90 percent and a

strik-ing 5:1 male predominance. Localized papillary re-nal-cell carcinoma metastasizes less frequently than clear-cell renal-cell carcinoma.34 However, the sur-vival rate for metastatic papillary renal-cell carci-noma is probably worse than that for clear-cell renal-cell carcinoma.35 The risk of both types is particularly increased among patients with end-stage renal disease. Chromosome 7, which harbors the MET proto-oncogene, is duplicated in 75 percent of sporadic papillary cases. There are two subtypes of papillary renal-cell carcinoma.36 Type 1 tumors are papillary lesions covered by small cells with pale cytoplasm and small oval nuclei with indistinct nucleoli, and type 2 tumors are papillary lesions covered by large cells with abundant eosinophilic cytoplasm. Type 2 cells are typified by pseudostratifi-cation and large, spherical nuclei with distinct nu-cleoli. Type 2 tumors are genetically more heteroge-neous, have a poorer prognosis, and may arise from type 1 tumors.37

Papillary renal-cell carcinoma occurs in several familial syndromes (MIM number 605074). He-reditary papillary renal carcinoma is an autosomal dominant disorder associated with multifocal pap-illary renal-cell carcinoma38 with type 1 histologic features (Fig. 3).39 The causative gene, mutations in which are responsible for hereditary papillary re-nal carcinoma, has been identified at chromo-some 7 and encodes MET, a receptor tyrosine kinase that is normally activated by hepatocyte growth factor40 (Fig. 4C). In hereditary papillary renal car-cinoma, the MET receptor tyrosine kinase domain undergoes autoactivating

amino-acid–substitu-*VHL denotes von Hippel–Lindau, FCRC familial clear-cell renal cancer, SDHB succinate dehydrogenase B, HPRC hered-itary papillary renal carcinoma, HLRCC heredhered-itary leiomyomatosis and renal-cell cancer, and FH fumarate hydratase. † Additional rare syndromes or infrequent associations are not included.

Table 1. Sporadic and Hereditary Renal-Cell Carcinomas and Genetic Defects According to Histologic Appearance.*

Sporadic Renal-Cell Carcinomas Renal-Cell Carcinomas in an Inherited Syndrome

Histologic Appearance Incidence Gene and Frequency Rare Syndrome† Gene percent Conventional 75 VHL, 60 VHL disease FCRC Hereditary paraganglioma VHL Chromosome 3p translocation SDHB Papillary 12 MET, 13 TFE3, <1 HPRC HLRCC MET FH

Chromophobe 4 Birt–Hogg–Dubé syndrome BHD

Oncocytoma 4 Birt–Hogg–Dubé syndrome BHD

Collecting duct <1 Unclassified 3–5

(4)

The n e w e n g l a n d j o u r n a l of m e d i c i n e

tion mutations, which promote cellular transforma-tion.41 Subsequently, chromosome 7 harboring the

MET mutation is duplicated, increasing the gene dose.42,43 Only a small percentage of the cases of the sporadic papillary type have MET mutations.40 Thus, the pathogenesis of hereditary papillary re-nal carcinoma is usually different from that of spo-radic papillary renal-cell carcinoma.

Patients with the hereditary leiomyomatosis and renal-cell cancer syndrome (MIM number 605839) are at risk for cutaneous and uterine leiomyo-mas and solitary papillary renal-cell carcinoma with

type 2 histologic features.44 Occasionally, cases of collecting-duct or clear-cell renal-cell carcinoma occur. These cases of papillary renal-cell carcino-ma metastasize early and are the most aggressive of the familial types.45 Intriguingly, FH, the gene that causes this autosomal dominant syndrome, encodes fumarate hydratase, a Krebs-cycle enzyme.46 As with the loss of a tumor-suppressor gene, the wild-type

FH allele is lost in hereditary leiomyomatosis and in lesions of renal-cell carcinoma.47 Along simi-lar lines, cases of renal-cell carcinoma with solid histologic features or cases of the clear-cell form

Figure 2. Schematic Representation of the Clinical Spectrum of von Hippel–Lindau Disease and Potential Biologic Mechanisms.

The three major manifestations of von Hippel–Lindau (VHL) disease — central nervous system hemangioblastoma, pheochromocytoma, and clear-cell renal-cell carcinoma — are represented as Venn diagrams. Affected families may not have all three conditions, depending large-ly on the type of mutation in the VHL gene inherited. Circle sizes are not intended to reflect the numbers of patients in each group. Type 1 VHL disease (Panel A) is most common; it comprises clear-cell renal-cell carcinoma with central nervous system hemangioblastoma and is due to major disruptions in the VHL protein, such as those resulting from truncating mutations. Type 2 VHL disease (Panel A) includes pheochro-mocytoma with one, both, or neither of the other features. Type 2B disease predisposes patients to all three conditions, whereas type 2A in-cludes hemangioblastoma and pheochromocytoma and type 2C inin-cludes only pheochromocytoma. Type 2 is most often due to more subtle VHL gene mutations, such as missense mutations that result in the substitution of a single amino acid. Besides classic von Hippel–Lindau disease, sporadic cases of hemangioblastoma or pheochromocytoma have been shown to be due to germ-line VHL mutations. In addition, familial Chuvash polycythemia, which does not include any of the features of von Hippel–Lindau disease, is an autosomal recessive disorder caused by homozygosity for the specific VHL missense mutation R200W. In Panel B, the spectrum of manifestations of von Hippel–Lindau disease, along with biochemical studies, suggests the involvement of specific biochemical pathways. For example, the mutations in VHL pro-tein that disrupt the ubiquitination and destruction of hypoxia-inducible factor a (HIF-a) are the same as those that correlate with the lesions of hemangioblastoma. Interestingly, von Hippel–Lindau disease type 2A mutations disrupt HIF-a processing and do not cause renal cancer, suggesting that the HIF pathway plus additional, unknown VHL pathways (called X pathways) must be disrupted for clear-cell renal-cell car-cinoma to develop. The probable involvement of additional pathways is supported by the multistep nature of renal tumorigenesis. Alterna-tively, the unknown pathway or pathways may simply reflect a higher level of expression of HIF-a in this subgroup of patients. Pheochromocy-tomas can arise from mutations in the gene for VHL protein that do not affect HIF-a processing, suggesting that they arise through the disruption of other, unknown VHL pathways (called Y pathways). Thus, the overexpression of HIF-a correlates closely with the development of hemangioblastoma; appears to be necessary for, but not sufficient to induce, renal tumorigenesis; and is not necessary for the develop-ment of pheochromocytoma.

(5)

m e d i c a l p r o g r e s s

Figure 3. Steps in the Development of Renal-Cell Carcinoma.

In contrast to sporadic renal-cell carcinoma (Panels A and C), fewer steps are required for the development of renal-cell carcinoma in the in-herited forms of the disease (Panels B and D), because all of the patient’s cells have a mutation that predisposes the patient to the disease. As a result, the disease associated with the familial syndromes occurs earlier and is often multifocal. Each familial renal cancer syndrome is autosomal dominant. In von Hippel–Lindau disease, a cellular recessive mechanism is involved, since both copies of the VHL gene are inac-tivated (Panels A and B). VHL is a classic tumor-suppressor gene. In hereditary papillary renal carcinoma, one copy of the MET gene has an activating mutation, which is inherited (Panel D). Chromosome 7, which includes the defective MET allele, becomes duplicated, increasing the level of expression of the activated MET protein, which is a receptor tyrosine kinase for hepatocyte growth factor. Activated MET is a clas-sic oncogene. A plus sign represents the wild-type allele; a minus sign represents a null allele. A plus sign in red type represents a mutated, activated allele; two plus signs in red type represent duplication of that allele.

(6)

The n e w e n g l a n d j o u r n a l of m e d i c i n e

have been reported in patients with the hereditary paraganglioma syndrome (MIM number 115310). Certain forms of hereditary paraganglioma are as-sociated with germ-line defects in the succinate de-hydrogenase B gene.48 Succinate dehydrogenase B protein is another mitochondrial, Krebs-cycle en-zyme. Thus, an intriguing connection exists among cellular ATP production, the hypoxic response, and tumorigenesis in both neuronal and kidney tissue. A number of sporadic cases of papillary renal-cell carcinoma have chromosomal translocations in-volving the TFE3 gene at chromosome Xp11.2.49-51 Children and young adults are affected without predilection for sex, and the histologic features of such cases have been variably described as papil-lary renal-cell carcinoma, clear-cell renal-cell carci-noma, or a unique type of pathology. The TFE3 gene encodes a helix–loop–helix transcription factor related to the proto-oncogene product c-myc. Key TRE3 domains become fused with other gene prod-ucts, and renal-cell carcinoma is probably due to

TFE3 overexpression.

o n c o c y t o m a a n d c h r o m o p h o b e r e n a l - c e l l c a r c i n o m a

Oncocytomas, which are benign, account for about 4 percent of nephrectomies performed because re-nal-cell carcinoma is suspected. The chromophobe variant of renal-cell carcinoma also accounts for 4 percent of all cases of renal-cell carcinoma52 and may have a benign course after surgery, provided that the tumor stage and grade are favorable.53 On-cocytoma is thought to originate from type A in-tercalated cells of the collecting duct, whereas chromophobe renal-cell carcinoma is thought to originate from type B intercalated cells.

The Birt–Hogg–Dubé syndrome (MIM number 135150) is a rare autosomal dominant disorder characterized by hair-follicle hamartomas (fibro-folliculomas) of the face and neck.54-57 About 15 percent of affected patients have multiple renal tumors, most often chromophobe or mixed chro-mophobe–oncocytomas. Occasionally, papillary or clear-cell renal-cell carcinoma develops in patients with the Birt–Hogg–Dubé syndrome. BHD, the gene implicated in the syndrome, encodes the protein fol-liculin,58 a suspected tumor suppressor. BHD mu-tations occur only rarely in sporadic renal-cell car-cinoma.59,60 The Birt–Hogg–Dubé renal phenotype supports the existence of a close relationship be-tween oncocytoma and chromophobe renal-cell carcinoma.

c o l l e c t i n g - d u c t r e n a l - c e l l c a r c i n o m a

Collecting-duct renal-cell carcinoma accounts for less than 1 percent of all cases of renal-cell carcino-ma and is typically an aggressive tumor. Medullary carcinoma of the kidney, which may be a variant of the collecting-duct type, is associated with sickle

Figure 4 (facing page). Molecular Mechanisms of the Development of Renal-Cell Carcinoma.

Pathologic cooperativity between renal-cancer cells of the clear-cell type and adjacent vasculature is shown in Panel A. In clear-cell renal-cell carcinoma, hypoxia-inducible factor a (HIF-a) transcription factor accumu-lates, resulting in the overexpression of proteins that are normally inducible with hypoxia, such as transforming growth factor a and b (TGF-a and TGF-b, respectively), vascular endothelial growth factor (VEGF), and platelet-derived growth factor B chain (PDGF-B). The overex-pressed VEGF, PDGF-B, and TGF-b act on neighboring vascular cells to promote tumor angiogenesis. The aug-mented tumor vasculature provides additional nutrients and oxygen to promote the growth of tumor cells. TGF-a

acts in an autocrine manner on the tumor cells by signal-ing through the epidermal growth factor receptor, which promotes tumor-cell proliferation and survival. The role of von Hippel–Lindau (VHL) protein in clear-cell renal-cell carcinoma and in controlling the expression of the HIF-a transcription factors is shown in Panel B. Under normoxic conditions HIF-a is hydroxylated on two pro-line residues by a propro-line hydroxylase and on an aspar-agine residue by an asparaspar-agine hydroxylase.

Hydroxylation (OH) by proline hydroxylase permits bind-ing of HIF-a to VHL protein, which promotes the ubiquit-ination (Ub) and destruction of HIF-a by the proteasome pathway. Hydroxylation by asparagine hydroxylase blocks the interaction of HIF-a with transcriptional coactivator p300. VHL protein, with elongin proteins C and B, binds cul2 protein (a member of the cullin family of ubiquitin ligase proteins). RING-box protein Rbx1 serves as the ubiquitin transferase for the VHL skp-cullin-F-box pro-tein (SCF) complex. In the absence of wild-type VHL protein, hydroxylated HIF-a accumulates and is able to heterodimerize with HIF-b and activate transcription at hypoxia-response elements (HREs), which are found in genes such as VEGF. In hypoxic conditions, HIF-a is not hydroxylated and so cannot bind VHL protein. Panel C shows the role of MET in papillary renal carcinoma. MET is a receptor tyrosine kinase for hepatocyte growth factor. In the absence of ligand, MET normally exists in an autoinhibited state. MET homodimerizes in the pres-ence of ligand hepatocyte growth factor and undergoes reciprocal phosphorylation and activation. In hereditary papillary renal carcinoma and in occasional sporadic papillary renal-cell carcinoma, activating mutations of MET disinhibit the receptor, even in the absence of ligand. Furthermore, the mutated MET allele on chromo-some 7 becomes duplicated, increasing the expression of the activated MET protein.

(7)
(8)

The n e w e n g l a n d j o u r n a l of m e d i c i n e

cell trait or disease. The collecting-duct form may be most similar to transitional-cell carcinoma of the urothelium.

An enhancing renal mass on a CT scan obtained after the administration of contrast material is a strong clue that renal cancer is present. A staging workup should be performed before treatment is initiated. Multiple enhancing lesions, or a family his-tory of renal-cell carcinoma, particularly in persons younger than 50 years of age, suggests a hereditary predisposition to the disease. Von Hippel–Lindau disease, hereditary leiomyomatosis and renal-cell cancer, and the Birt–Hogg–Dubé syndrome all have extrarenal manifestations, whereas familial clear-cell renal cancer and hereditary papillary renal carci-noma do not. Thus, a careful physical examination including ophthalmologic, neurologic, and derma-tologic evaluation may be helpful. CT scanning or magnetic resonance imaging (MRI) of the abdomen and pelvis may reveal uterine tumors in patients with hereditary leiomyomatosis and renal-cell can-cer or renal cysts or pancreatic or adrenal involve-ment in patients with von Hippel–Lindau disease.

Patients with hereditary renal-cell carcinoma should be closely monitored. CT before and after the administration of contrast material is the best test for detection and assessment of renal masses, with gadolinium-enhanced MRI as an alternative. Such studies can be performed at intervals ranging from every three to six months to every two to three years, depending on the size of the lesions and the type of syndrome. Larger masses require more fre-quent evaluation. Because small masses are usually of low grade, they can be observed until they reach 3 cm, at which time they should be removed.61,62 However, tumors caused by hereditary leiomyoma-tosis and renal-cell cancer should be excised im-mediately because of their aggressive nature. Pa-tients with von Hippel–Lindau disease should undergo MRI studies of the brain and spinal cord to screen for hemangioblastoma. A family pedigree should be generated, and family members at risk should be encouraged to seek medical attention. Testing is available for the VHL, MET, FH, and BHD

genes. One goal of such testing is to free unaffect-ed family members from continuunaffect-ed cancer

screen-ing. Organizations such as the VHL Family Alli-ance (www.vhl.org) are a vital resource for patients, families, physicians, and researchers.

Defining the prognosis of renal-cell carcinoma is important for both therapeutic decision-making and counseling patients. For metastatic renal-cell carcinoma, poor prognostic factors include a low Karnofsky performance-status score (a standard way of measuring functional impairment in pa-tients with cancer), a high level of serum lactate de-hydrogenase, a low hemoglobin level, and a high corrected level of serum calcium.63,64 The Univer-sity of California, Los Angeles, Integrated Staging System was developed to evaluate the prognosis at diagnosis and in the presence of metastatic dis-ease; it includes tumor–node–metastasis (TNM) staging, the patient’s score on the Eastern Cooper-ative Oncology Group performance-status scale (an-other measure of functional impairment in patients with cancer), and the Fuhrman nuclear grade, which assesses histologic features of the tumor.65 This sys-tem has been used successfully in more than 4000 patients at eight international centers.66

r a d i c a l n e p h r e c t o m y

Surgical excision is the primary treatment for re-nal-cell carcinoma. Radical nephrectomy, which includes removal of the kidney en bloc with Gero-ta’s fascia, the ipsilateral adrenal gland, and re-gional lymph nodes, has been the standard thera-py, although more limited approaches are being explored. The surgical approach is determined by the size and location of the tumor within the kid-ney, the TNM stage, and any special anatomical considerations.

Staging and evaluation for the presence of me-tastases, including a careful history-taking and physical examination, should be completed before surgery. Routine laboratory studies should include measurement of the hematocrit and serum levels of creatinine, calcium, and alkaline phosphatase and a urinalysis for proteinuria. Imaging studies, such as radiographs of the chest, CT of the abdomen and pelvis, and in some cases, MRI evaluation of the re-nal vein and inferior vena cava, CT of the chest or head, or bone scanning may be needed. The fre-m a n a g e fre-m e n t o f s p o r a d i c

a n d h e r e d i t a r y r e n a l - c e l l c a r c i n o m a

p r o g n o s i s

(9)

m e d i c a l p r o g r e s s

quency of follow-up after surgery depends on the stage of the tumor.

s u r g e r y f o r m e t a s t a t i c d i s e a s e

Nephrectomy may be warranted, even in the pres-ence of metastatic disease. The combination of in-terferon alfa and nephrectomy is superior to inter-feron alfa alone, offering a survival advantage of 3 to 10 months.67,68 Surgical excision of a solitary metastasis in patients with advanced renal-cell car-cinoma is recommended in many cases, but this approach has not yet been proved to be effective in prolonging survival.

n e p h r o n - s p a r i n g p a r t i a l n e p h r e c t o m y

Nephron-sparing partial nephrectomy has gained acceptance for treating tumors less than 4 cm in di-ameter. Other indications for partial nephrectomy may include a solitary kidney, bilateral renal masses, or renal insufficiency, as well as the presence of hy-pertension, diabetes, or hereditary renal-cell carci-noma syndromes. Results achieved with nephron-sparing surgery are similar to those with radical nephrectomy, but a disadvantage is a rate of local recurrence of 3 to 6 percent.69

l a p a r o s c o p i c n e p h r e c t o m y

First reported in 1991,70 laparoscopic nephrecto-my has accelerated the evolution toward minimally invasive surgical management of renal-cell carci-noma. The benefits of the laparoscopic approach include decreased postoperative pain, a shorter hos-pitalization, and a quicker recovery. The laparo-scopic approach has been used for both radical nephrectomy and partial nephrectomy. The laparo-scopic partial nephrectomy, however, is a technically demanding procedure with the potential for in-creased perioperative complications.71

p e r c u t a n e o u s a b l a t i v e a p p r o a c h e s

The most recent evolution in the surgical manage-ment of small tumors has been percutaneous ther-mal ablative techniques that use radiofrequency heat ablation72 or cryoablation73 to destroy tumor cells. A needle probe is advanced through the skin and directed into the tumor under image guid-ance. Although early results of radiofrequency ab-lation and cryoabab-lation are encouraging, larger tri-als with long-term follow-up are needed. The rates of complications appear to be low, but reported ad-verse events include intraoperative and postopera-tive hemorrhage, urinary leakage, and injury to

adjacent structures. Because identification of the type of renal-cell carcinoma is important, a core bi-opsy of the renal mass should be performed as part of the procedure. Ideal candidates for minimally invasive percutaneous ablative therapy are patients with tumors less than 3 cm in diameter who have serious coexisting conditions and for whom stan-dard approaches would pose substantial risks. Pa-tients with multifocal tumors may also benefit from minimally invasive percutaneous procedures. High-frequency focused ultrasound applied externally to the body is being studied as another potential min-imally invasive therapy.

Medical therapies are generally offered for locally advanced or metastatic renal-cell carcinoma (Ta-ble 2), and much of the clinical experience with this approach is in patients with the clear-cell type. Because response rates are low, the need to identify new therapeutic agents is great.74

c h e m o t h e r a p y

Rates of response to chemotherapy alone are low (roughly 4 to 6 percent).75 Drug resistance may be related to the expression of the multidrug resis-tance transporter in proximal-tubule cells — the cells from which clear-cell and papillary renal-cell carcinoma may originate. Chemotherapy may be more efficacious for advanced non–clear-cell re-nal-cell carcinoma, particularly the collecting-duct type.76-78 A phase 2 trial of carboplatin and pacli-taxel for the collecting-duct form of the disease is under way.

i m m u n o m o d u l a t o r y t h e r a p i e s

The value of immunomodulatory therapy for clear-cell renal-clear-cell carcinoma is supported by reports of occasional spontaneous tumor regression, infre-quent complete regression of metastatic disease with cytokine therapies, and promising early results with allogeneic stem-cell transplantation and tumor vaccines. The goal of immunomodulatory therapy is to boost either tumor antigenicity or host surveil-lance. Unique tumor antigens may also be induc-ible in renal-cell carcinoma.79

Interferon Alfa

About 14 percent of cases of metastatic clear-cell renal carcinoma respond to interferon alfa alone. Various doses and routes have been used.80 The

(10)

The n e w e n g l a n d j o u r n a l of m e d i c i n e

dian duration of response is six months and rarely exceeds two years. Because the side effects of the drug are not onerous, it appears to be a good choice to use in combination with other agents in experi-mental approaches.

Interleukin-2

High-dose interleukin-2 is the standard therapy for advanced renal-cell carcinoma and is the only regi-men for this disease approved by the Food and Drug Administration. However, many patients with metastatic disease cannot take high-dose interleu-kin-2, because it causes a capillary leak syndrome or because it is not available in all treatment cen-ters. High-dose interleukin-2 induces responses in 21 percent of patients, as compared with only 13 percent of patients who receive low-dose interleu-kin-2.81 The median duration of response has been reported to be 54 months overall, and for those with a complete response, the median duration of a response is yet to be reached.82 Interleukin-2 has also been used in combination with other drugs, but it is unclear whether combined therapy achieves better results than interleukin-2 alone. Thus, inter-leukin-2 is a highly effective therapy for a subgroup

of patients with metastatic disease. Identifying fea-tures predictive of a response to interleukin-2 would represent a further advance, and efforts are being made to identify patients with clear-cell renal carci-noma who would be likely to have a response to in-terleukin-2 therapy on the basis of pathological characteristics and expression of CA9.

a d j u v a n t t h e r a p y

Given the high rate of recurrence of renal-cell car-cinoma after nephrectomy, a follow-up adjuvant approach would be desirable, especially for pa-tients with high-risk, locally advanced disease. However, conventional chemotherapy, interferon alfa, or even interleukin-283 has not proved effec-tive as an adjuvant therapy. Approaches currently being tested include tumor vaccines and a mono-clonal antibody directed against CA9.

s t e m - c e l l t r a n s p l a n t a t i o n

Allogeneic stem-cell transplantation performed af-ter the administration of a non–marrow ablative regimen elicits a potent graft-versus-tumor effect and appears promising for treating clear-cell renal-cell carcinoma.84 Protocols developed at the Nation-al Institutes of HeNation-alth have used myelosuppress-ive pretreatment, followed by an infusion of donor CD34+ cells and T cells from an HLA-identical sib-ling.84 A course of immunosuppressive agents, such as cyclosporine, is used to limit graft-versus-host disease and is rapidly tapered. Twenty of the first 45 patients with metastatic renal-cell carcino-ma who underwent stem-cell transplantation had a response (44 percent).85 However, results in some other centers have been less promising. The re-sponses have correlated well with the development of graft-versus-host disease and with the conver-sion of T-cell chimerism to full donor origin. One goal is to identify the tumor epitopes that are initi-ating the graft-versus-tumor response to improve treatment specificity. The two drawbacks to stem-cell transplantation have been severe graft-versus-host disease, which can be life-threatening, and the need for a haplotype-matched sibling donor. Prognosis is also an important guide to patient se-lection, since responses take several months. The next generation of strategies for stem-cell transplan-tation may include the use of tumor vaccines after transplantation as well as the use of cytokine thera-py to boost recipient or even donor immunity.

e v o l v i n g t h e r a p i e s

* FDA denotes Food and Drug Administration, VEGF vas-cular endothelial growth factor, and EGFR epidermal growth factor receptor. The trade name for interleukin-2 is Proleukin; for bevacizumab, Avastin; for sunitinib malate, Sutent; for sorafenib tosylate, Nexavar; for gefi-tinib, Iressa; for erlogefi-tinib, Tarceva.

Table 2. Medical Therapies for Advanced Renal Cancer.* FDA-approved regimen

High-dose interleukin-2 (aldesleukin, immunomodula-tory cytokine)

Commonly used agents

Low-dose interleukin-2

Interferon alfa (immunomodulatory cytokine)

Experimental therapies

Bevacizumab (humanized VEGF-neutralizing antibody) Sunitinib malate (VEGF receptor and multitargeted kinase

inhibitor)

Sorafenib tosylate (VEGF receptor and multitargeted ki-nase inhibitor)

Panitumumab (human EGFR-neutralizing antibody) Gefitinib (EGFR tyrosine kinase inhibitor)

Erlotinib (EGFR tyrosine kinase inhibitor)

Temsirolimus (inhibitor of the mammalian target of rapamycin)

Tumor vaccines

(11)

m e d i c a l p r o g r e s s

t u m o r v a c c i n e s

Tumor vaccines represent a potential means of en-hancing host immunity. A promising approach to the treatment of advanced clear-cell renal carcinoma uses autologous or donor dendritic cells, which initiate a primary immune response by presenting antigen in the context of costimulatory molecules. Dendritic cells can be pulsed with tumor protein,86 DNA, or RNA87; they can even be fused with tumor cells88,89 to present tumor antigens in a context fa-vorable for therapy. Such vaccines are generally well tolerated, but they will require further optimi-zation. Concomitant administration of cytokines may improve the response to vaccines.

t a r g e t a n t i g e n s

A goal of stem-cell or vaccine therapies is to char-acterize the tumor antigens involved in the immune response. One potential target is the G250 renal can-cer antigen, which has been identified as CA9. The

CA9 gene is a target of HIF and so is overexpressed in VHL-related clear-cell renal carcinoma, even in the earliest lesions of von Hippel–Lindau disease.90 Thus, in cases of renal-cell carcinoma, a high pro-portion of CA9-positive cells may be associated with a more favorable prognosis.91 As a transmembrane protein, CA9 may also be a therapeutically useful tu-mor antigen.92,93 It will be important to identify additional target antigens.

t h e r a p i e s t a r g e t i n g v e g f a n d t g f -a p a t h w a y s

Originally identified as regulated by VHL, VEGF and TGF-a are now promising therapeutic targets in clear-cell renal carcinoma. The manner in which these molecules interact with the cancer epitheli-um and surrounding vascular endotheliepitheli-um leads to tumor progression (Fig. 4A). A combination of ther-apies based on rational targets such as these may therefore be a powerful approach to advanced renal-cell carcinoma.

VEGF-Pathway Components as Molecular Targets VEGF is overexpressed throughout clear-cell renal-cell carcinoma tissue and may be the most impor-tant tumor angiogenic factor. A randomized phase 2 trial involving patients with metastatic renal-cell carcinoma investigated the efficacy of bevacizu-mab, a humanized VEGF-neutralizing antibody.94 This agent extended the interval before tumor pro-gression to 4.8 months, as compared with 2.5 months for placebo. Bevacizumab therefore

pro-vided a key “proof of principle” of the efficacy of anti-angiogenic therapy and may offer additional benefit when given in combination with other drugs. Inhibitors of VEGF receptor tyrosine kinase are being developed and tested. Indeed, the multi-targeted kinase inhibitors sunitinib and sorafenib have shown great promise in phase 2 and phase 3 trials, with at least stabilization of disease in as many as 70 percent of patients with cytokine-refractory disease.

TGF-a–Pathway Components as Molecular Targets

TGF-a is a potent growth factor for epithelial cells that acts through the epidermal growth factor re-ceptor (EGFR), which is a rere-ceptor tyrosine kinase. TGF-a is overexpressed in the epithelium in clear-cell renal carcinoma and is a VHL target.95 Over-expression of TGF-a is an early event in the patho-genesis of this disease.96 Furthermore, growth of renal cancer cells in culture is dependent on TGF-a.97 Thus, the TGF-a pathway is a logical choice for therapeutic intervention.

Antibodies against EGFR are thought to bind EGFR and promote its down-regulation from the cell surface. A fully human monoclonal anti-body against human EGFR, called panitumumab (ABX-EGF), has been evaluated in a phase 2 trial involving 88 patients with metastatic renal-cell car-cinoma.98 Only one patient had a complete re-sponse, and two had partial responses — a disap-pointing result. Small-molecule inhibitors of the EGFR tyrosine kinase are also being developed.99 The quinazolines gefitinib and erlotinib are now in phase 2 trials. In a phase 1 trial of erlotinib, just one patient with metastatic disease had a com-plete response.100 Erlotinib is also being tested in combination with bevacizumab, although encour-aging initial results could not be confirmed in a randomized phase 2 trial.

Other Approaches

Temsirolimus (CCI-779), a selective inhibitor of the mammalian target of rapamycin, has shown efficacy in a phase 2 trial of metastatic renal-cell carcinoma.101 Temsirolimus may inhibit HIF as well. Partial responses were noted in 7 percent of patients, and minor responses in 26 percent. The median survival rate was 15 months. The notable activity of the drug in patients with poor prognos-tic features prompted a phase 3 trial. Other op-tions are being pursued, including agents target-ing HIF.

(12)

The n e w e n g l a n d j o u r n a l of m e d i c i n e

The poor prognosis of advanced renal-cell carci-noma demands an aggressive search for new ther-apeutic agents and strategies. Leads will probably come from both a careful elucidation of the biology of each type of the disease and broader approach-es, such as gene-expression–array and proteome analyses. Much has been accomplished since the identification of the VHL gene in 1993. Already, VHL protein pathways, such as those involving VEGF and TGF-a, are being exploited therapeuti-cally, and agents affecting these pathways might be more effective when used in combination. Identi-fication of the MET gene was another key advance.

The mutated, activated hepatocyte growth factor receptor MET could be targeted in the papillary form of the disease.102,103 The immune respon-siveness of renal-cell carcinoma provides an op-portunity for the development and optimization of vaccines and other immune therapies. Preserva-tion of as much renal funcPreserva-tion as possible and re-duced rates of complications are two goals of new minimally invasive approaches to renal-cell carci-noma; other goals are to identify early markers of disease, prognosis, or responsiveness to therapy.

Supported in part by grants (RO1 CA79830 and RO1 DK67569, to Dr. Cohen) from the National Institutes of Health.

We are indebted to Drs. Michael B. Atkins, William G. Kaelin, Jr., Matthew R. Smith, Walter M. Stadler, and Berton Zbar for careful re-view of the manuscript.

s u m m a r y a n d p r o s p e c t s f o r t h e f u t u r e

r e f e r e n c e s

1. Jemal A, Murray T, Ward E, et al. Cancer statistics, 2005. CA Cancer J Clin 2005;55: 10-30.

2. Chow WH, Gridley G, Fraumeni JF Jr, Jarvholm B. Obesity, hypertension, and the risk of kidney cancer in men. N Engl J Med 2000;343:1305-11.

3. Javidan J, Stricker HJ, Tamboli P, et al. Prognostic significance of the 1997 TNM classification of renal cell carcinoma. J Urol 1999;162:1277-81.

4. Kidney. In: Fleming ID, Cooper JS, Hen-son DE, et al., eds. AJCC cancer staging hand-book. Philadelphia: Lippincott-Raven, 1998: 215-7.

5. Kidney. In: Greene FL, Page DL, Fleming ID, et al., eds. AJCC cancer staging manual. New York: Springer-Verlag, 2002:323-5.

6. Kovacs G, Akhtar M, Beckwith BJ, et al. The Heidelberg classification of renal cell tumours. J Pathol 1997;183:131-3.

7. Latif F, Tory K, Gnarra J, et al. Identifica-tion of the von Hippel-Lindau disease tumor suppressor gene. Science 1993;260:1317-20.

8. Kim WY, Kaelin WG. Role of VHL gene mutation in human cancer. J Clin Oncol 2004;22:4991-5004.

9. Iliopoulos O, Kibel A, Gray S, Kaelin WG Jr. Tumour suppression by the human von Hippel-Lindau gene product. Nat Med 1995;1:822-6.

10.Chen F, Kishida T, Duh FM, et al. Sup-pression of growth of renal carcinoma cells by the von Hippel-Lindau tumor suppressor gene. Cancer Res 1995;55:4804-7.

11.Iliopoulos O, Levy AP, Jiang C, Kaelin WG Jr, Goldberg MA. Negative regulation of hypoxia-inducible genes by the von Hippel-Lindau protein. Proc Natl Acad Sci U S A 1996;93:10595-9.

12.Pause A, Lee S, Worrell RA, et al. The von Hippel-Lindau tumor-suppressor gene product forms a stable complex with human CUL-2, a member of the Cdc53 family of

proteins. Proc Natl Acad Sci U S A 1997;94: 2156-61.

13.Lonergan KM, Iliopoulos O, Ohh M, et al. Regulation of hypoxia-inducible mRNAs by the von Hippel-Lindau tumor suppressor protein requires binding to complexes con-taining elongins B/C and Cul2. Mol Cell Biol 1998;18:732-41.

14.Maxwell PH, Wiesener MS, Chang GW, et al. The tumour suppressor protein VHL targets hypoxia-inducible factors for oxy-gen-dependent proteolysis. Nature 1999; 399:271-5.

15.Cockman ME, Masson N, Mole DR, et al. Hypoxia inducible factor-alpha binding and ubiquitylation by the von Hippel-Lindau tumor suppressor protein. J Biol Chem 2000; 275:25733-41.

16.Ohh M, Park CW, Ivan M, et al. Ubiquit-ination of hypoxia-inducible factor requires direct binding to the beta-domain of the von Hippel-Lindau protein. Nat Cell Biol 2000; 2:423-7.

17.Tanimoto K, Makino Y, Pereira T, Poel-linger L. Mechanism of regulation of the hypoxia-inducible factor-1 alpha by the von Hippel-Lindau tumor suppressor protein. EMBO J 2000;19:4298-309.

18.Ivan M, Kondo K, Yang H, et al. HIF-alpha targeted for VHL-mediated destruc-tion by proline hydroxyladestruc-tion: implicadestruc-tions for O2 sensing. Science 2001;292:464-8.

19.Jaakkola P, Mole DR, Tian YM, et al. Tar-geting of HIF-alpha to the von Hippel-Lindau ubiquitylation complex by O2-regulated

pro-lyl hydroxylation. Science 2001;292:468-72.

20.Epstein AC, Gleadle JM, McNeill LA, et al. C. elegans EGL-9 and mammalian ho-mologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell 2001;107:43-54.

21.Bruick RK, McKnight SL. A conserved family of prolyl-4-hydroxylases that modify HIF. Science 2001;294:1337-40.

22.Lando D, Peet DJ, Whelan DA, Gorman JJ, Whitelaw ML. Asparagine hydroxylation of the HIF transactivation domain a hypoxic switch. Science 2002;295:858-61.

23.Clifford SC, Cockman ME, Smallwood AC, et al. Contrasting effects on HIF-1alpha regulation by disease-causing pVHL muta-tions correlate with patterns of tumourigen-esis in von Hippel-Lindau disease. Hum Mol Genet 2001;10:1029-38.

24.Hoffman MA, Ohh M, Yang H, Klco JM, Ivan M, Kaelin WG Jr. von Hippel-Lindau protein mutants linked to type 2C VHL dis-ease preserve the ability to downregulate HIF. Hum Mol Genet 2001;10:1019-27.

25.Kondo K, Kim WY, Lechpammer M, Kaelin WG Jr. Inhibition of HIF2alpha is sufficient to suppress pVHL-defective tumor growth. PLoS Biol 2003;1:E83.

26.Zimmer M, Doucette D, Siddiqui N, Il-iopoulos O. Inhibition of hypoxia-inducible factor is sufficient for growth suppression of VHL¡/¡ tumors. Mol Cancer Res 2004;2: 89-95.

27.Ohh M, Yauch RL, Lonergan KM, et al. The von Hippel-Lindau tumor suppressor protein is required for proper assembly of an extracellular fibronectin matrix. Mol Cell 1998;1:959-68.

28.Feldman DE, Spiess C, Howard DE, Frydman J. Tumorigenic mutations in VHL disrupt folding in vivo by interfering with chaperonin binding. Mol Cell 2003;12: 1213-24.

29.Hergovich A, Lisztwan J, Barry R, Ballschmieter P, Krek W. Regulation of mi-crotubule stability by the von Hippel-Lindau tumour suppressor protein pVHL. Nat Cell Biol 2003;5:64-70.

30.Zhou MI, Wang H, Ross JJ, Kuzmin I, Xu C, Cohen HT. The von Hippel-Lindau tumor suppressor stabilizes novel plant homeo-domain protein Jade-1. J Biol Chem 2002; 277:39887-98.

(13)

m e d i c a l p r o g r e s s

31.Zhou MI, Wang H, Foy RL, Ross JJ, Co-hen HT. Tumor suppressor von Hippel-Lindau (VHL) stabilization of Jade-1 protein occurs through plant homeodomains and is VHL mutation dependent. Cancer Res 2004; 64:1278-86.

32.Zhou MI, Foy RL, Chitalia VC, et al. Jade-1, a candidate renal tumor suppressor that promotes apoptosis. Proc Natl Acad Sci U S A 2005;102:11035-40.

33.Cohen AJ, Li FP, Berg S, et al. Hereditary renal-cell carcinoma associated with a chro-mosomal translocation. N Engl J Med 1979; 301:592-5.

34.Beck SD, Patel MI, Snyder ME, et al. Ef-fect of papillary and chromophobe cell type on disease-free survival after nephrectomy for renal cell carcinoma. Ann Surg Oncol 2004;11:71-7.

35.Motzer RJ, Bacik J, Mariani T, Russo P, Mazumdar M, Reuter V. Treatment outcome and survival associated with metastatic re-nal cell carcinoma of non-clear-cell histolo-gy. J Clin Oncol 2002;20:2376-81.

36.Delahunt B, Eble JN. Papillary renal cell carcinoma: a clinicopathologic and immu-nohistochemical study of 105 tumors. Mod Pathol 1997;10:537-44.

37.Gunawan B, von Heydebreck A, Fritsch T, et al. Cytogenetic and morphologic typing of 58 papillary renal cell carcinomas: evidence for a cytogenetic evolution of type 2 from type 1 tumors. Cancer Res 2003;63:6200-5.

38.Zbar B, Glenn G, Lubensky I, et al. He-reditary papillary renal cell carcinoma: clini-cal studies in 10 families. J Urol 1995;153: 907-12.

39.Lubensky IA, Schmidt L, Zhuang Z, et al. Hereditary and sporadic papillary renal car-cinomas with c-met mutations share a dis-tinct morphological phenotype. Am J Pathol 1999;155:517-26.

40.Schmidt L, Duh FM, Chen F, et al. Germline and somatic mutations in the ty-rosine kinase domain of the MET proto-oncogene in papillary renal carcinomas. Nat Genet 1997;16:68-73.

41.Jeffers M, Schmidt L, Nakaigawa N, et al. Activating mutations for the met tyrosine kinase receptor in human cancer. Proc Natl Acad Sci U S A 1997;94:11445-50.

42.Zhuang Z, Park WS, Pack S, et al. Tri-somy 7-harbouring non-random duplica-tion of the mutant MET allele in hereditary papillary renal carcinomas. Nat Genet 1998; 20:66-9.

43.Fischer J, Palmedo G, von Knobloch R, et al. Duplication and overexpression of the mutant allele of the MET proto-oncogene in multiple hereditary papillary renal cell tu-mours. Oncogene 1998;17:733-9.

44.Launonen V, Vierimaa O, Kiuru M, et al. Inherited susceptibility to uterine leiomyo-mas and renal cell cancer. Proc Natl Acad Sci U S A 2001;98:3387-92.

45.Zbar B, Klausner R, Linehan WM. Study-ing cancer families to identify kidney cancer genes. Annu Rev Med 2003;54:217-33.

46.Tomlinson IP, Alam NA, Rowan AJ, et

al. Germline mutations in FH predispose to dominantly inherited uterine fibroids, skin leiomyomata and papillary renal cell cancer. Nat Genet 2002;30:406-10.

47.Kiuru M, Launonen V, Hietala M, et al. Familial cutaneous leiomyomatosis is a two-hit condition associated with renal cell can-cer of characteristic histopathology. Am J Pathol 2001;159:825-9.

48.Vanharanta S, Buchta M, McWhinney SR, et al. Early-onset renal cell carcinoma as a novel extraparaganglial component of SDHB-associated heritable paraganglioma. Am J Hum Genet 2004;74:153-9.

49.Meloni AM, Dobbs RM, Pontes JE, Sand-berg AA. Translocation (X;1) in papillary re-nal cell carcinoma: a new cytogenetic sub-type. Cancer Genet Cytogenet 1993;65:1-6.

50.Sidhar SK, Clark J, Gill S, et al. The t(X;1)(p11.2;q21.2) translocation in papil-lary renal cell carcinoma fuses a novel gene PRCC to the TFE3 transcription factor gene. Hum Mol Genet 1996;5:1333-8.

51.Weterman MA, Wilbrink M, Geurts van Kessel A. Fusion of the transcription fac-tor TFE3 gene to a novel gene, PRCC, in t(X;1)(p11;q21)-positive papillary renal cell carcinomas. Proc Natl Acad Sci U S A 1996; 93:15294-8.

52.Cheville JC, Lohse CM, Zincke H, Weav-er AL, Blute ML. Comparisons of outcome and prognostic features among histologic subtypes of renal cell carcinoma. Am J Surg Pathol 2003;27:612-24.

53.Peyromaure M, Misrai V, Thiounn N, et al. Chromophobe renal cell carcinoma: analysis of 61 cases. Cancer 2004;100:1406-10.

54.Birt AR, Hogg GR, Dube WJ. Hereditary multiple fibrofolliculomas with trichodisco-mas and acrochordons. Arch Dermatol 1977; 113:1674-7.

55.Weirich G, Glenn G, Junker K, et al. Fa-milial renal oncocytoma: clinicopathologi-cal study of 5 families. J Urol 1998;160:335-40.

56.Toro JR, Glenn G, Duray P, et al. Birt-Hogg-Dube syndrome: a novel marker of kidney neoplasia. Arch Dermatol 1999;135: 1195-202.

57.Zbar B, Alvord WG, Glenn G, et al. Risk of renal and colonic neoplasms and sponta-neous pneumothorax in the Birt-Hogg-Dube syndrome. Cancer Epidemiol Bio-markers Prev 2002;11:393-400.

58.Nickerson ML, Warren MB, Toro JR, et al. Mutations in a novel gene lead to kidney tumors, lung wall defects, and benign tumors of the hair follicle in patients with the Birt-Hogg-Dube syndrome. Cancer Cell 2002;2: 157-64.

59.Khoo SK, Kahnoski K, Sugimura J, et al. Inactivation of BHD in sporadic renal tu-mors. Cancer Res 2003;63:4583-7.

60.da Silva NF, Gentle D, Hesson LB, Mor-ton DG, Latif F, Maher ER. Analysis of the Birt-Hogg-Dube (BHD) tumour suppressor gene in sporadic renal cell carcinoma and co-lorectal cancer. J Med Genet 2003;40:820-4.

61.Walther MM, Choyke PL, Glenn G, et al. Renal cancer in families with hereditary re-nal cancer: prospective are-nalysis of a tumor size threshold for renal parenchymal spar-ing surgery. J Urol 1999;161:1475-9.

62.Herring JC, Enquist EG, Chernoff A, Linehan WM, Choyke PL, Walther MM. Pa-renchymal sparing surgery in patients with hereditary renal cell carcinoma: 10-year ex-perience. J Urol 2001;165:777-81.

63.Motzer RJ, Mazumdar M, Bacik J, Berg W, Amsterdam A, Ferrara J. Survival and prog-nostic stratification of 670 patients with ad-vanced renal cell carcinoma. J Clin Oncol 1999;17:2530-40.

64.Motzer RJ, Bacik J, Schwartz LH, et al. Prognostic factors for survival in previously treated patients with metastatic renal cell carcinoma. J Clin Oncol 2004;22:454-63.

65.Zisman A, Pantuck AJ, Dorey F, et al. Im-proved prognostication of renal cell carcino-ma using an integrated staging system. J Clin Oncol 2001;19:1649-57.

66.Patard JJ, Kim HL, Lam JS, et al. Use of the University of California Los Angeles in-tegrated staging system to predict survival in renal cell carcinoma: an international multi-center study. J Clin Oncol 2004;22:3316-22.

67.Flanigan RC, Salmon SE, Blumenstein BA, et al. Nephrectomy followed by interfer-on alfa-2b compared with interferinterfer-on alfa-2b alone for metastatic renal-cell cancer. N Engl J Med 2001;345:1655-9.

68.Mickisch GH, Garin A, van Poppel H, de Prijck L, Sylvester R. Radical nephrecto-my plus interferon-alfa-based immunother-apy compared with interferon alfa alone in metastatic renal-cell carcinoma: a random-ised trial. Lancet 2001;358:966-70.

69.Novick AC. Nephron-sparing surgery for renal cell carcinoma. Annu Rev Med 2002; 53:393-407.

70.Clayman RV, Kavoussi LR, Soper NJ, et al. Laparoscopic nephrectomy: initial case report. J Urol 1991;146:278-82.

71.Gill IS, Matin SF, Desai MM, et al. Com-parative analysis of laparoscopic versus open partial nephrectomy for renal tumors in 200 patients. J Urol 2003;170:64-8.

72.Zlotta AR, Wildschutz T, Raviv G, et al. Radiofrequency interstitial tumor ablation (RITA) is a possible new modality for treat-ment of renal cancer: ex vivo and in vivo ex-perience. J Endourol 1997;11:251-8.

73.Uchida M, Imaide Y, Sugimoto K, Ueha-ra H, Watanabe H. Percutaneous cryosur-gery for renal tumours. Br J Urol 1995;75: 132-6.

74.Motzer RJ. Renal cell carcinoma: a pri-ority malignancy for development and study of novel therapies. J Clin Oncol 2003;21: 1193-4.

75.Yagoda A, Abi-Rached B, Petrylak D. Chemotherapy for advanced renal-cell car-cinoma: 1983-1993. Semin Oncol 1995;22: 42-60.

76.Gollob JA, Upton MP, DeWolf WC, At-kins MB. Long-term remission in a patient with metastatic collecting duct carcinoma

(14)

m e d i c a l p r o g r e s s

treated with taxol/carboplatin and surgery. Urology 2001;58:1058.

77.Peyromaure M, Thiounn N, Scotte F, Vieillefond A, Debre B, Oudard S. Collecting duct carcinoma of the kidney: a clinicopath-ological study of 9 cases. J Urol 2003;170: 1138-40.

78.Milowsky MI, Rosmarin A, Tickoo SK, Papanicolaou N, Nanus DM. Active chemo-therapy for collecting duct carcinoma of the kidney: a case report and review of the litera-ture. Cancer 2002;94:111-6.

79.Hanada K, Yewdell JW, Yang JC. Im-mune recognition of a human renal cancer antigen through post-translational protein splicing. Nature 2004;427:252-6.

80.Small EJ, Motzer RJ. Interferon for renal cell carcinoma. In: Belldegrun A, Ritchie AWS, Figlin RA, Oliver RTD, Vaughan ED, eds. Renal and adrenal tumors. New York: Oxford University Press, 2003:381-7.

81.Yang JC, Sherry RM, Steinberg SM, et al. Randomized study of high-dose and low-dose interleukin-2 in patients with meta-static renal cancer. J Clin Oncol 2003;21: 3127-32.

82.Fisher RI, Rosenberg SA, Fyfe G. Long-term survival update for high-dose recombi-nant interleukin-2 in patients with renal cell carcinoma. Cancer J Sci Am 2000;6:Suppl 1: S55-S57.

83.Clark JI, Atkins MB, Urba WJ, et al. Ad-juvant high-dose bolus interleukin-2 for pa-tients with high-risk renal cell carcinoma: a cytokine working group randomized trial. J Clin Oncol 2003;21:3133-40.

84.Childs R, Chernoff A, Contentin N, et al. Regression of metastatic renal-cell carcino-ma after nonmyeloablative allogeneic periph-eral-blood stem-cell transplantation. N Engl J Med 2000;343:750-8.

85.Srinivasan R, Barrett J, Childs R. Alloge-neic stem cell transplantation as immuno-therapy for nonhematological cancers. Sem-in Oncol 2004;31:47-55.

86.Holtl L, Zelle-Rieser C, Gander H, et al. Immunotherapy of metastatic renal cell

car-cinoma with tumor lysate-pulsed autolo-gous dendritic cells. Clin Cancer Res 2002; 8:3369-76.

87.Su Z, Dannull J, Heiser A, et al. Immu-nological and clinical responses in meta-static renal cancer patients vaccinated with tumor RNA-transfected dendritic cells. Can-cer Res 2003;63:2127-33.

88.Marten A, Renoth S, Heinicke T, et al. Allogeneic dendritic cells fused with tumor cells: preclinical results and outcome of a clinical phase I/II trial in patients with meta-static renal cell carcinoma. Hum Gene Ther 2003;14:483-94.

89.Avigan D, Vasir B, Gong J, et al. Fusion cell vaccination of patients with metastatic breast and renal cancer induces immuno-logical and clinical responses. Clin Cancer Res 2004;10:4699-708.

90.Mandriota SJ, Turner KJ, Davies DR, et al. HIF activation identifies early lesions in VHL kidneys: evidence for site-specific tumor suppressor function in the nephron. Cancer Cell 2002;1:459-68.

91.Bui MH, Seligson D, Han KR, et al. Car-bonic anhydrase IX is an independent pre-dictor of survival in advanced renal clear cell carcinoma: implications for prognosis and therapy. Clin Cancer Res 2003;9:802-11.

92.Luiten RM, Coney LR, Fleuren GJ, War-naar SO, Litvinov SV. Generation of chimeric bispecific G250/CD3 monoclonal anti-body, a tool to combat renal cell carcinoma. Br J Cancer 1996;74:735-44.

93.Steffens MG, Boerman OC, de Mulder PH, et al. Phase I radioimmunotherapy of metastatic renal cell carcinoma with 131I-labeled chimeric monoclonal antibody G250. Clin Cancer Res 1999;5:3268s-3274s.

94.Yang JC, Haworth L, Sherry RM, et al. A randomized trial of bevacizumab, an anti-vascular endothelial growth factor antibody, for metastatic renal cancer. N Engl J Med 2003;349:427-34.

95.Knebelmann B, Ananth S, Cohen HT, Sukhatme VP. Transforming growth factor alpha is a target for the von Hippel-Lindau

tumor suppressor. Cancer Res 1998;58:226-31.

96.Everitt JI, Walker CL, Goldsworthy TW, Wolf DC. Altered expression of transform-ing growth factor-alpha: an early event in re-nal cell carcinoma development. Mol Car-cinog 1997;19:213-9.

97.de Paulsen N, Brychzy A, Fournier MC, et al. Role of transforming growth factor-al-pha in von Hippel–Lindau (VHL)(¡/¡) clear cell renal carcinoma cell proliferation: a pos-sible mechanism coupling VHL tumor sup-pressor inactivation and tumorigenesis. Proc Natl Acad Sci U S A 2001;98:1387-92.

98.Rowinsky EK, Schwartz GH, Gollob JA, et al. Safety, pharmacokinetics, and activity of ABX-EGF, a fully human anti-epidermal growth factor receptor monoclonal antibody in patients with metastatic renal cell cancer. J Clin Oncol 2004;22:3003-15.

99.Arteaga CL. Overview of epidermal growth factor receptor biology and its role as a therapeutic target in human neoplasia. Semin Oncol 2002;29:Suppl:3-9.

100. Hidalgo M, Siu LL, Nemunaitis J, et al. Phase I and pharmacologic study of OSI-774, an epidermal growth factor receptor ty-rosine kinase inhibitor, in patients with ad-vanced solid malignancies. J Clin Oncol 2001; 19:3267-79.

101. Atkins MB, Hidalgo M, Stadler WM, et al. Randomized phase II study of multiple dose levels of CCI-779, a novel mammalian target of rapamycin kinase inhibitor, in pa-tients with advanced refractory renal cell car-cinoma. J Clin Oncol 2004;22:909-18.

102. Lynch TJ, Bell DW, Sordella R, et al. Ac-tivating mutations in the epidermal growth factor receptor underlying responsiveness of non–small-cell lung cancer to gefitinib. N Engl J Med 2004;350:2129-39.

103. Paez JG, Janne PA, Lee JC, et al. EGFR mutations in lung cancer: correlation with clinical response to gefitinib therapy. Sci-ence 2004;304:1497-500.

Copyright © 2005 Massachusetts Medical Society.

apply for jobs electronically at the nejm careercenter

Physicians registered at the NEJM CareerCenter can apply for jobs electronically using their own cover letters and CVs. You can keep track of your job-application history with a personal account that is created when you register with the CareerCenter and apply for jobs seen online at our Web site. Visit www.nejmjobs.org for more information.

References

Related documents

A lot of research work has already been done on switching behaviours around the globe, but seemingly, no research has been carried out specifically on employee perceptions on

notes: Psychiatric acute beds include those in psychiatric units in general acute care hospitals (including city and county hospitals), acute psychiatric hospitals, and psychiatric

Good public sector service performance means that the public sector organization is effective and efficient in providing goods and services to its

Recommendations emanating from the literature review are presented to enhance service delivery for the Nelson Mandela Bay Municipality (NMBM) through Performance

 Acknowledged as Gold level sponsor in 2013 OPFA newsletter. SILVER SPONSOR

Evaluation of the livestock sector’s contribution to the EU greenhouse gas emissions (GGELS) – final report. European Commission, Joint Research Centre. Errors in conventional

Conditional on documentation level and FICO score, there is more variation in the percentage fees of smaller loans than of larger ones, and broker fees reveal more about

By undertaking a thorough review of the sentencing remarks of 40 insider trading cases (37 criminal and 3 civil cases), this paper presented an overview of the