Bladder Cancer

Back

Practice Essentials

Bladder cancer is a common urologic cancer that has the highest recurrence rate of any malignancy. In North America, South America, Europe, and Asia, the most common type is transitional cell carcinoma. Other types include squamous cell carcinoma (see the image below) and adenocarcinomas.



View Image

Cross-section through the bladder, uterus, and vagina with squamous cell carcinoma of the bladder infiltrating through the bladder wall into the vagin....

Signs and symptoms

Clinical manifestations of bladder cancer are as follows:

See Presentation for more detail.

Diagnosis

Urine studies include the following:

Urinary cytology

Cystoscopy

Upper urinary tract imaging

The diagnostic strategy for patients with negative cystoscopy is as follows:

No blood tests are specific for bladder cancer, but a general evaluation is necessary prior to initiating therapy with intravesical bacillus Calmette-Guérin (BCG) vaccine. Laboratory tests include the following:

See Workup for more detail.

Management

The treatment of non–muscle-invasive bladder cancer (Ta, T1, carcinoma in situ [CIS]) begins with transurethral resection of bladder tumor (TURBT). Subsequent treatment is as follows:

The treatment of muscle-invasive bladder cancer is as follows:

Alternatively, a bladder-sparing approach of TURBT followed by concurrent radiation therapy and systemic chemotherapy (trimodality therapy) may be used.

Chemotherapeutic regimens for metastatic bladder cancer include the following:

Inhibitors of programmed cell death 1 (PD-1) protein and its ligands PD-L1 and PD-L2, are first-line agents in patients with metastatic urothelial carcinoma who are not candidates for cisplatin chemotherapy, and second-line agents for those with disease progression despite cisplatin-based chemotherapy. Agents in this category include the following[1] :

Erdafitinib, a fibroblast growth factor receptor inhibitor, is approved for locally advanced or metastatic urothelial carcinoma that has FGFR2 or FGFR3 genetic alterations and has progressed during or following at least 1 line of prior platinum-containing chemotherapy.

See Treatment and Medication for more detail.

Background

Bladder cancer is a common urologic cancer. Almost all bladder cancers originate in the urothelium, which is a 3- to 7-cell mucosal layer within the muscular bladder.

Transitional (urothelial) cell carcinoma

In North America, South America, Europe, and Asia, the most common type of urothelial tumor diagnosed is transitional (urothelial) cell carcinoma (TCC); it constitutes more than 90% of bladder cancers in those regions. TCC can arise anywhere in the urinary tract, including the renal pelvis, ureter, bladder, and urethra, but it is usually found in the urinary bladder. Carcinoma in situ (CIS) is frequently found in association with high-grade or extensive TCC. (See the image below.)



View Image

The classic appearance of carcinoma in situ as a flat, velvety patch. However, using special staining techniques such as 5-aminolevulinic acid, it has....

Squamous cell carcinoma

Squamous cell carcinoma (SCC) is the second most common cell type associated with bladder cancer in industrialized countries. In the United States, around 5% of bladder cancers are SCCs.[2] Worldwide, however, SCC is the most common form of bladder cancer, accounting for 75% of cases in developing nations (see Epidemiology).

In the United States, the development of SCC is associated with persistent inflammation from long-term indwelling Foley catheters and bladder stones, as well as, possibly, infections. In developing nations, SCC is often associated with bladder infection by Schistosoma haematobium (see Etiology).

Other types of bladder cancer

Approximately 2% of bladder cancers are adenocarcinomas. Nonurothelial primary bladder tumors are extremely rare and may include small cell carcinoma, carcinosarcoma, primary lymphoma, and sarcoma (see Pathophysiology). Small cell carcinoma of the urinary bladder accounts for only 0.3-0.7% of all bladder tumors. High-grade urothelial carcinomas can also show divergent histologic differentiation, such as squamous, glandular, neuroendocrine, and sarcomatous features.

Phenotypes

Clinical and pathologic data indicate that at least 3 different phenotypes, as follows, exist in urothelial carcinoma[2, 3] :

Clinical course

The clinical course of bladder cancer is marked by a broad spectrum of aggressiveness and risk. Low-grade, superficial bladder cancers have minimal risk of progression to death; however, high-grade non–muscle-invasive cancers frequently progress and muscle-invasive cancers are often lethal (see Prognosis).

The classic presentation of bladder cancer is painless gross hematuria, which is seen in approximately 80-90% of patients. Physical examination results are often unremarkable (see Presentation). Cystoscopy, cytology, and biopsy when necessary are the principal diagnostic tests (see Workup).

Upon presentation, 55-60% of patients have low-grade, noninvasive disease, which is usually treated conservatively with transurethral resection of bladder tumor (TURBT) and periodic cystoscopy. Intravesical agents may also be given selectively to decrease the frequency of recurrences. The remaining patients have high-grade disease, of which 50% is muscle invasive and is typically treated with radical cystectomy or with trimodality therapy (ie, TURBT followed by concurrent radiation therapy and systemic chemotherapy; see Treatment).

Carcinoma in situ (CIS) is managed by TURBT and instillation of chemotherapeutic or immunotherapeutic agents—most commonly, immunotherapy with bacillus Calmette-Guérin (BCG) vaccine—into the bladder via catheter. These intravesical treatments are not effective in the 20% of patients in whom cancer has invaded the bladder wall muscle; those cases require cystectomy or a combination of radiation therapy and chemotherapy (see Treatment).

Bladder cancer has the highest recurrence rate of any malignancy. Although most patients with bladder cancer can be treated with organ-sparing therapy, most experience either recurrence or progression, creating a great need for accurate and diligent surveillance (see Treatment).

For more information on bladder cancer, see the following:

Anatomy

The bladder is an extraperitoneal muscular urine reservoir that lies behind the pubis symphysis in the pelvis. At the dome of the bladder lies the median umbilical ligament, a fibrous cord that is anchored to the umbilicus and that represents the obliterated urachus (allantois). The ureters, which transport urine from kidney to bladder, approach the bladder obliquely and posterosuperiorly, entering at the trigone (the area between the interureteric ridge and the bladder neck). The intravesical ureteral orifices are roughly 2-3 cm apart and form the superolateral borders of the trigone. The bladder neck serves as an internal sphincter, which is sacrificed during a radical cystectomy.

In males, the seminal vesicles, vas deferens, ureters, and rectum border the inferoposterior aspect of the bladder. Anterior to the bladder is the space of Retzius, which is composed of fibroadipose tissue and the prevesical fascia. The dome and posterior surface of the bladder are covered by parietal peritoneum, which reflects superiorly to the seminal vesicles and is continuous with the anterior rectal peritoneum. In females, the posterior peritoneal reflection is continuous with the uterus and vagina.

The vascular supply to the bladder arrives primarily via the internal iliac (hypogastric) arteries, branching into the superior, middle, and inferior vesical arteries, which are often recognizable as lateral and posterior pedicles. The arterial supply also arrives via the obturator and inferior gluteal artery and, in females, via the uterine and vaginal arteries. Bladder venous drainage is a rich network that often parallels the named arterial vessels, most of which ultimately drain into the internal iliac vein.

Initial lymphatic drainage from the bladder is primarily into the external iliac, obturator, internal iliac (hypogastric), and common iliac nodes. Following the drainage to these sentinel pelvic regions, spread may continue to the presacral, paracaval, interaortocaval, and para-aortic lymph node chains.

Almost all bladder cancers originate in the urothelium, which is a 3- to 7-cell mucosal layer within the muscular bladder. Squamous cell carcinoma of the bladder can involve multiple sites; however, the lateral wall and trigone are more commonly involved by this tumor. All small cell carcinomas of the urinary system identified so far have been located in the urinary bladder, most commonly in the dome and vesical lateral wall.[4]

See Bladder Anatomy.

Pathophysiology

Bladder cancer is often described as a polyclonal field change defect with frequent recurrences due to a heightened potential for malignant transformation. However, bladder cancer has also been described as resulting from implantation of malignant cells that have migrated from a previously affected site. The latter occurs less often and may account for only a small percentage of cases.

Use of the common term superficial bladder cancer should be discouraged. The term implies a harmless nature, which is misleading in many instances. Because it was used to describe the disparate disorders of low-grade papillary bladder cancer and the markedly more aggressive form, carcinoma in situ (CIS), the World Health Organization (WHO) has recommended it be abandoned.

In its place, the term non–muscle-invasive bladder cancer should be used and qualified with the appropriate American Joint Committee on Cancer stage (ie, Ta, T1, Tis). Stage T1 cancer invades lamina propria but not the muscle of the bladder. High-grade T1 tumor associated with CIS carries a relatively high risk for disease recurrence and progression (approximately 60%).

The WHO classifies bladder cancers as low grade (grades 1 and 2) or high grade (grade 3). Tumors are also classified by growth patterns: papillary (70%), sessile or mixed (20%), and nodular (10%). (See the images below.)



View Image

Papillary bladder tumors such as this one are typically of low stage and grade (Ta-G1). Courtesy of Abbott and Vysis Inc.



View Image

Sessile lesions as shown usually invade muscle, although occasionally a tumor is detected at the T1-G3 stage prior to muscle invasion. Courtesy of Abb....

Transitional cell carcinoma

Transitional cell carcinoma (TCC) arises from stem cells that are adjacent to the basement membrane of the epithelial surface. Depending on the genetic alterations that occur, these cells may follow different pathways in the expression of their phenotype.

The most common molecular biologic pathway for TCCs involves the development of a papillary tumor that projects into the bladder lumen and, if untreated, eventually penetrates the basement membrane, invades the lamina propria, and then continues into the bladder muscle, where it can metastasize. Nearly 90% of transitional cell bladder tumors exhibit this type of behavior.

This progression occurs with high-grade cancers only. Low-grade cancers rarely, if ever, progress and are thought to have a distinct molecular pathway, different from the high-grade cancers and CIS.

The remaining 10% of TCCs follow a different molecular pathway and are called CIS. This is a flat, noninvasive, high-grade urothelial carcinoma tumor that spreads along the surface of the bladder and, over time, may progress to an invasive form of cancer that behaves the same as invasive TCC.

Many urothelial tumors are primarily composed of TCC but contain small areas of squamous differentiation, squamous cell carcinoma (SCC), or adenocarcinoma.

Squamous cell carcinoma

SCC of the urinary bladder is a malignant neoplasm that is derived from bladder urothelium and has a pure squamous phenotype.[5, 6, 7] SCC of the bladder is essentially similar to squamous cell tumors arising in other organs. Because many urothelial carcinomas contain a minor squamous cell component, a diagnosis of SCC of the bladder should be rendered only when the tumor is solely composed of squamous cell components, with no conventional urothelial carcinoma component.

Reportedly, SCC has less of a tendency for nodal and vascular distant metastases than does urothelial carcinoma.[8, 9]

Rare forms of bladder cancer

Adenocarcinomas account for less than 2% of primary bladder tumors. These lesions are observed most commonly in exstrophic bladders and are often associated with malignant degeneration of a persistent urachal remnant.

Other rare forms of bladder cancer include leiomyosarcoma, rhabdosarcoma, carcinosarcoma, lymphoma, and small cell carcinoma. Leiomyosarcoma is the most common sarcoma of the bladder. Rhabdomyosarcomas most commonly occur in children. Carcinosarcomas are highly malignant tumors that contain a combination of mesenchymal and epithelial elements. Primary bladder lymphomas arise in the submucosa of the bladder. Except for lymphomas, all these rare bladder cancers carry a poor prognosis.

Small cell carcinoma of the urinary bladder is a poorly differentiated, malignant neoplasm that originates from urothelial stem cells and has variable expression of neuroendocrine markers. Morphologically, it shares features of small cell carcinoma of other organs, including the lung.

Genetic factors in pathogenesis

Divergent, yet interconnected and overlapping, molecular pathways are likely responsible for the development of noninvasive and invasive bladder tumors. Somatic mutations in fibroblast growth receptor3 (FGFR-3) and tumor protein p53 (TP53) in tumor cells appear to be important early molecular events in the noninvasive and invasive pathways, respectively.

FGFR-3, Ras, and PIK3CA mutations occur with high frequency in noninvasive tumors, leading to upregulation of Akt and mitogen-activated protein kinase (MAPK).[10, 11] Loss of heterozygosity (LOH) on chromosome 9 is among the most frequent genetic alterations in bladder tumors and is considered an early event.[12]

Large numbers of genomic changes have been detected using karyotyping and comparative genomic hybridization (CGH) analysis in urothelial carcinoma. Numerically common are losses of 2q, 5q, 8p, 9p, 10q, 18q, and Y. Gains of 1q, 5p, 8q, and 17q are frequently present, and high-level amplifications can be found; however, the target genes in the regions of amplifications have not been conclusively identified.[13]

Alterations in the TP53 gene are noted in approximately 60% of invasive bladder cancers. Progression-free survival is significantly shorter in patients with TP53 mutations and is an independent predictor of death among patients with muscle-invasive bladder cancer.[14]

Additionally, alterations in retinoblastoma (Rb), PTEN, and p16 are common in high-grade invasive cancers.[15] Overexpression of JUN, MAP2K6, STAT3, and ICAM1 and molecules associated with survival (Bcl-2, caspase-3, p53, survivin), as well as insensitivity to antigrowth signals (p53, p21, p16, pRB), has been demonstrated.[16]

In advanced disease, multiple mechanisms may lead to tumor progression. These include those that promote proliferation, survival, invasion, and metastasis, as well as those that involve deficiencies in DNA damage repair and the finding of stemlike cells.

In addition to tumor cell alterations, the microenvironment may promote tumor growth by paracrine influences, including vascular endothelial growth factor (VEGF) production and aberrant E-cadherin expression. Finally, a growing body of research over the last decade indicates that epigenetic alterations may silence tumor suppressor genes and that they represent important events in tumor progression.[17, 18, 19]

Etiology

Up to 80% of bladder cancer cases are associated with environmental exposure. Tobacco use is by far the most common cause of bladder cancer in the United States and is increasing in importance in some developing countries. Smoking duration and intensity are directly related to increased risk.[20, 21, 2]

The risk of developing bladder carcinoma is 2-6 times greater in smokers than in nonsmokers. This risk appears to be similar between men and women.[22] Nitrosamine, 2-naphthylamine, and 4-aminobiphenyl are possible carcinogenic agents found in cigarette smoke.

A number of occupations involve exposure to substances that may increase risk for bladder cancer. Of occupationally related bladder cancer cases, the incidence rate is highest in workers exposed to aromatic amines, while mortality is greatest in those exposed to polycyclic aromatic hydrocarbons and heavy metals.[23]

Numerous occupations associated with diesel exhaust, petroleum products, and solvents (eg, auto work, truck driving, plumbing, leather and apparel work, rubber and metal work) have also been associated with an increased risk of bladder cancer. In addition, increased bladder carcinoma risk has been reported in persons, including the following, who work with organic chemicals and dyes:

People living in urban areas are also more likely to develop bladder cancer. The etiology in these cases is thought to be multifactorial, potentially involving exposure to numerous carcinogens.

Arsenic exposure may be a factor in the development of bladder cancer. Results of a population-based case-control study in Maine, New Hampshire, and Vermont support an association between low-to-moderate levels of arsenic in drinking water and bladder cancer risk in those states, where incidence rates of bladder cancer have long been about 20% higher than in the United States overall.[24] A likely source of the arsenic is residue of arsenic-based pesticides, which were used extensively on crops such as blueberries, apples, and potatoes in that region from the 1920s through to the 1950s.[25]

Several medical risk factors are associated with an increased risk of bladder cancer, including the following:

Although certain common genetic polymorphisms appear to increase susceptibility in persons with occupational exposure associated with increased bladder cancer risk,[27] no convincing evidence exists for a hereditary factor in the development of bladder cancer. Nevertheless, familial clusters of bladder cancer have been reported.

Schistosomiasis

In many developing countries, particularly in the Middle East, Schistosoma haematobium infection causes most cases of bladder SCC. In a study from Egypt, 82% of patients with bladder carcinoma harbored S haematobium eggs in the bladder wall. In egg-positive patients, the tumor tended to develop at a younger age (with SCC predominant) than it did in egg-negative persons. A higher degree of adenocarcinoma has also been reported in schistosomal-associated bladder carcinomas.[28]

Along with S haematobium, the species S mansoni and S japonicum are responsible for schistosomiasis in humans. The eggs reside in the pelvic and mesenteric venous plexus. In the bladder, a severe inflammatory response and fibrosis secondary to the deposition of Schistosoma eggs is common. (See the image below.)



View Image

Histopathology of bladder shows eggs of Schistosoma haematobium surrounded by intense infiltrates of eosinophils and other inflammatory cells.

The eggs are found embedded in the lamina propria and muscularis propria of the bladder wall. Many of the eggs are destroyed by host reaction and become calcified, resulting in a lesion commonly known as a sandy patch, which appears as a granular, yellow-tan surface lesion.

In normal epithelial cells, S haematobium total antigen reportedly induces increased proliferation, migration, and invasion and decreases apoptosis.[29] Keratinous squamous metaplasia has been associated with the increased risk of developing SCC, with approximately one half of the cases arising subsequent to the metaplasia.[30, 31]

The majority of schistosomiasis-related cases of SCC will arise in the setting of chronic cystitis.[32] Chronic irritation secondary to lithiasis,[5, 6] urinary retention, and indwelling catheters has also been linked to the development of SCC.[6]

Other squamous cell carcinoma risk factors

Having bladder diverticula may increase an individual’s chance of developing SCC.[33] Rarely, bacillus Calmette-Guerin (BCG) treatment for CIS has been reported to lead to development of SCC.[34] Development of bladder cancer at a younger age has been associated with bladder exstrophy.[35, 36, 37, 38] SCC has also been described in urachal remnants.[39, 40, 41, 42, 43]

Coffee consumption does not increase the risk of developing bladder cancer. Early studies of rodents and a minority of human studies suggested a weak connection between artificial sweeteners (eg, saccharin, cyclamate) and bladder cancer; however, most recent studies show no significant correlation.

Epidemiology

Occurrence in the United States

The American Cancer Society estimates that 80,470 new cases of bladder cancer will be diagnosed in the United States in 2019 and that 17,670 people will die of the disease.[44] The incidence of bladder cancer increases with age, with the median age at diagnosis being 72 years; bladder cancer is rarely diagnosed before age 40 years.[45]

Bladder cancer is about 3 times more common in men than in women. Over the past 2 decades, however, the rate of bladder cancer has been stable in men but has increased in women by 0.2% annually.[46] The male predominance in bladder cancer in the United States reflects the prevalence of transitional cell carcinoma (TCC). With small cell carcinoma (SCC)—in contrast to TCC—the male-to-female incidence ratio is 1:2.

Bladder cancer is the fourth most common cancer in men in the United States, after prostate, lung, and colorectal cancer, but it is not among the top 10 cancers in women. Accordingly, more men than women are expected to die of bladder cancer in 2016, with 11,820 deaths in men versus 4570 in women.[44] Nevertheless, women generally have a worse prognosis than men.

The incidence of bladder cancer is twice as high in white men as in black men in the United States. However, blacks have a worse prognosis than whites.[46, 47]

Limited data indicate that SCC of the urinary bladder probably has the same epidemiologic characteristics as urothelial carcinoma. Patients are more likely to be male and older than 50 years.[48, 49]

International occurrence

Worldwide, bladder cancer is diagnosed in approximately 275,000 people each year, and about 108,000 die of this disease. In industrialized countries, 90% of bladder cancers are TCC. In developing countries—particularly in the Middle East and Africa—the majority of bladder cancers are SCCs, and most of these cancers are secondary to Schistosoma haematobium infection. Urothelial carcinoma is reported to be the most common urologic cancer in China.

In Africa, the highest incidence of SCC has been seen in schistosomal-endemic areas, notably Sudan and Egypt, where SCC ranges from two thirds to three quarters of all malignant tumors of the bladder. In recent years, a few studies from Egypt have shown a reversal of this trend due to the better control of schistosomiasis in the region, whereas in other parts of Africa the association is unchanged.[9, 50, 51] Increased smoking incidence is believed to have contributed to the shift in Egypt toward TCC, which has a stronger smoking association.

Prognosis

Untreated bladder cancer produces significant morbidity, including the following:

The recurrence rate for superficial TCC of the bladder is high. As many as 80% of patients have at least one recurrence.

The most significant prognostic factors for bladder cancer are grade, depth of invasion, and the presence of CIS. In patients undergoing radical cystectomy for muscle-invasive bladder cancer, the presence of nodal involvement is the most important prognostic factor. To date, there is no convincing evidence of genetic factors affecting outcome.[52]

Non–muscle invasive bladder cancer has a good prognosis, with 5-year survival rates of 82-100%. The 5-year survival rate decreases with increasing stage, as follows:

Prognosis for patients with metastatic urothelial cancer is poor, with only 5-10% of patients living 2 years after diagnosis.

The risk of progression, defined as an increased tumor grade or stage, depends primarily on the tumor grade, as follows:

Prognosis in carcinoma in situ

CIS in association with T1 papillary tumor carries a poorer prognosis. It has a recurrence rate of 63-92% and a rate of progression to muscle invasion of 50-75% despite intravesical BCG. Diffuse CIS is an especially ominous finding; in one study, 78% of cases progressed to muscle-invasive disease.[53]

Prognosis in squamous cell carcinoma

Tumor stage, lymph node involvement, and tumor grade have been shown to be of independent prognostic value in SCC.[54, 55] However, pathologic stage is the most important prognostic factor. In one relatively large series of 154 cases, the overall 5-year survival rate was 56% for pT1 and 68% for pT2 tumors. However, the 5-year survival rate for pT3 and pT4 tumors was only 19%.[52]

Several studies have demonstrated grading to be a significant morphologic parameter in SCC.[52] In one series, 5-year survival rates for grade 1, 2, and 3 squamous cell carcinoma was 62%, 52%, and 35%, respectively.[52] In the same study of patients undergoing cystectomy, the investigators suggested that a higher number of newly formed blood vessels predicts unfavorable disease outcome.

In SCC, the survival rate appears to be better with radical surgery than with radiation therapy and/or chemotherapy. In locally advanced tumors, however, neoadjuvant radiation improves the outcome.[56] Sex and age have not been prognostically significant in SCC.[52]

Prognosis in small cell carcinoma

Patients with small cell carcinoma of the bladder usually have disease in an advanced stage at diagnosis, and they have a poor prognosis.[57, 58, 59] Overall median survival is only 1.7 years. The 5-year survival rates for stage II, III, and IV disease are 64%, 15%, and 11%, respectively.[60]

Recurrent bladder cancer

Bladder cancer has the highest recurrence rate of any malignancy (ie, 70% within 5 y). Although most patients with bladder cancer can be treated initially with organ-sparing therapy, most experience either recurrence or progression. The underlying genetic changes that result in a bladder tumor occur in the entire urothelium, making the whole lining of the urinary system susceptible to tumor recurrence.

Risk factors for recurrence and progression include the following[61, 62] :

The time interval to recurrence is also significant. Patients with tumor recurrences within 2 years, and especially with recurrences within 3-6 months, have an aggressive tumor and an increased risk of disease progression.

History

Approximately 80-90% of patients with bladder cancer present with painless gross hematuria. All patients with this classic presentation should be considered to have bladder cancer until proof to the contrary is found.

Irritative bladder symptoms such as dysuria, urgency, or frequency of urination occur in 20-30% of patients with bladder cancer. Although irritative symptoms may be related to more advanced muscle-invasive disease, carcinoma in situ (CIS) is the more likely cause. Therefore, patients presenting with unexplained or refractory irritative symptoms should be considered for cystoscopy and urine cytology. The threshold for doing so should be especially low in persons who smoke or have other risk factors.

Patients with advanced disease can present with pelvic or bony pain, lower-extremity edema from iliac vessel compression, or flank pain from ureteral obstruction.

Physical Examination

Non ̶ muscle-invasive bladder cancer is typically not found during a physical examination. In rare cases, a mass is palpable during abdominal, pelvic, rectal, or bimanual examination. A bimanual examination may be considered part of the staging of such lesions. In women, a bimanual pelvic examination is done; in men, a bimanual examination is performed with one hand per rectum and the other on the lower abdominal wall.

Attention to the anterior vaginal wall in women and the prostate in men may reveal findings that suggest local extension of bladder cancer. Assessment of fixation of the bladder to the surrounding pelvic sidewall is also important when planning definitive management for locally advanced tumors that may not be surgically resectable.

Approach Considerations

Urine studies include the following:

Urinalysis is performed to detect hematuria or infection. Microscopic hematuria from bladder cancer may be intermittent; therefore, a repeat negative result on urinalysis does not exclude the diagnosis.

Cytology

Urinary cytology is extremely valuable and is often the test used for diagnosis; suggestive urine cytology findings should encourage the urologist to perform a cystoscopy and possibly bladder biopsy. All patients with gross hematuria and many with significant microscopic hematuria should undergo a cystoscopy and urinary cytology (preferably barbotage urine for cytology).

Urinary cytology is most helpful in diagnosing high-grade tumors and carcinoma in situ (CIS). Low-grade, noninvasive tumors may be missed by routine cytologic analysis.

Endoscopic biopsies are used to establish the diagnosis and determine the extent of the cancer. However, a study by Cha et al found that immunocytology outperforms urine cytology and increases the accuracy of predictive models by a statistically and clinically significant margin for patients with painless hematuria.[63]

Strittmatter et al found that the quality of urinary cytology is impacted by the individual learning curve. Specificity of cytology and sensitivity for low-grade tumors significantly changed when performed by a local cytologist at the beginning of the learning period. This suggests that in the diagnosis of bladder cancer, the cytologist’s level of experience has an important impact on the clinical value of urinary cytology.[64]

Findings

Because cytology is the most reliable urine test for detecting bladder cancer, a positive cytology finding should be treated as indicating cancer until proven otherwise. If cystoscopy findings are negative in the setting of positive cytology findings, further evaluation of the urinary tract is required. The upper urinary tract should be evaluated with contrast imaging. Cystoscopy with bilateral retrograde pyelography and bilateral ureteral washings should be performed.

Most patients with CIS have coexisting papillary cancer. In general, the papillary tumor is diagnosed first, and CIS is discovered during the evaluation and treatment of the papillary tumor. Only 10% of patients with bladder cancer have a pure CIS. The combination of CIS and papillary transitional cell carcinoma (TCC) is associated with a higher risk of recurrence and progression.

In cases of pure CIS, urinary cytology may lead to the diagnosis. CIS exfoliates cells that have an unusual appearance and are easy to identify via cytologic examination, prompting further evaluation. Unfortunately, even findings from urine cytology may be normal in some patients; in these cases the diagnosis is made only when the urologist maintains a high level of suspicion for CIS and obtains random bladder biopsy specimens from patients with worrisome symptoms. However, if the urinary cytology is performed properly, this should happen rarely.

Cystoscopy

Cystoscopy is the criterion standard for detecting bladder cancer, but it is invasive and relatively expensive.[65] Moreover, visibility can be reduced by bleeding, and flat urothelial lesions such as CIS may be difficult to distinguish from normal bladder tissue. Use of adjunctive endoscopic techniques, such as blue light cystoscopy with 5-aminolevulinic acid, may improve the accuracy of cystoscopy. Cytologic analysis of voided urine is frequently used as an adjunctive test to aid in identifying occult cancers.

Virtual cystoscopy can help detect many bladder tumors, but it is more expensive than cystoscopy and has lower sensitivity and specificity. Therefore, it does not play a role in surveillance at this time.

Imaging studies

Imaging studies of the upper urinary tract are an integral part of the hematuria workup. Computed tomography (CT) scans of the abdomen and pelvis with contrast are recommended. Two commonly used alternative techniques are magnetic resonance imaging (MRI) and renal ultrasonography. Few US centers still perform intravenous pyelography (IVP) for upper tract imaging.

The bladder urothelium is not well visualized with routine imaging studies, including CT scanning and MRI. Small tumors are easily missed on images produced by these modalities. Irregular areas on images, which may appear to represent mucosal abnormalities, are often artifacts of incomplete bladder filling; delayed images following contrast administration can better visualize actual filling defects. CIS is not visible on images from any current radiographic study.

Other tests

Markers

Newer molecular and genetic markers, including detection of mutations in genes such as RAS, FGFR3,PIK3CA, and TP53, and methylation pathways in urinary sediment,[66] may help in the early detection and prediction of urothelial carcinoma. At this time, however, no urinary assay has been shown to effectively replace urine cytology and cystoscopy, with or without biopsy, for the diagnosis of bladder cancer. Nevertheless, marker assays may be useful adjuncts to urine cytology and cystoscopy.

Blood tests

No blood tests are specific for bladder cancer. In patients with CIS, however, a general evaluation is necessary prior to initiating therapy with intravesical bacillus Calmette-Guérin (BCG) vaccine.

Complete Blood Count and Chemistry Panel

On the complete blood count (CBC), the presence of anemia or an elevated white blood cell (WBC) count warrants further investigation for an explanation.

The chemistry panel should include liver function studies. One of the intravesical agents used to treat CIS is BCG vaccine, but systemic absorption of this agent can produce acute hepatitis. Performing baseline liver function tests before initiating therapy and repeating these tests during the course of therapy help to prevent serious adverse events and to determine when therapy should be stopped. Liver function tests, as well as measurement of the bony fraction of alkaline phosphatase, are also indicated in patients with possible metastasis to liver or bone.

Kidney function should be evaluated prior to the initiation of therapy because patients with marginal or abnormal renal function may have an obstruction or some type of renal disease that may worsen with intravesical therapy. Renal function can be evaluated with serum creatinine measurements or technetium scans of the kidneys.

Urinalysis and Urine Culture

Urinalysis is used routinely to evaluate for the presence of red blood cells (RBCs), WBCs, and protein and to assess for urinary tract infection. The presence of RBCs in the urine mandates either a repeat study or an evaluation by a urologist to investigate for any serious disease.

Gross hematuria always requires a careful assessment with imaging studies of the entire urinary tract (CT urography) and cystoscopy. Prior to performing an endoscopic examination or initiating any therapy, a urine culture should be performed to confirm that the urine is free of evidence of infection.

Men do not usually have RBCs in their urine, and any number should lead to further urologic testing. In women, RBCs can frequently be found in a voided specimen, but persistent microhematuria warrants further testing.

Less than 20% of bladder cancer cases are identified on the basis of microscopic hematuria. Routine care visits or employment screens may be the setting for such a finding.

According to the American Urological Association Guidelines Committee, "the recommended definition of microscopic hematuria is 3 or more red blood cells per high-power microscopic field in urinary sediment from two of three properly collected urinalysis specimens. This definition accounts for some degree of hematuria in normal patients, as well as the intermittent nature of hematuria in patients with urologic malignancies."[67]

Urine Cytology

Voided urine cytology is the standard noninvasive method for diagnosis in the detection of bladder carcinoma. Cytology is used to assess morphologic changes in intact cells. Exfoliated urothelial cells are viewed using microscopy. In some urothelial cancers, cellular clumping, a high nuclear-to-cytoplasmic ratio, nucleoli, and atypia are seen.

As with any type of cytologic examination, the experience and skill of the cytopathologist is extremely important. Many hospital laboratories lack the personnel and technology necessary to accurately perform this type of study. Good reference laboratories are available if local facilities cannot provide this service.

At least 100 mL of a freshly voided specimen is usually sufficient for urine cytology. The first morning sample should not be used, because cells sitting in the urine overnight tend to become distorted and are difficult to analyze. If the urine is very dilute, the number of cells may be insufficient, necessitating a larger urine volume.

Bladder washings can be obtained by placing a catheter into the bladder and vigorously irrigating with saline (ie, barbotage). Bladder wash cytology yields more tumor cells in the sample and is more sensitive in identifying cancer, especially for high-grade tumors, but it also yields a higher false-positive rate than voided urine cytology.[68]

Sensitivity and specificity

Unfortunately, the sensitivity of cytology is low, with various studies reporting values between 11% and 76%.[69] Sensitivity depends largely on the degree of tumor differentiation. High-grade tumors with marked pleomorphism and distinctly abnormal nuclear features are identified more accurately.

Small and/or well-differentiated tumors are less likely to exfoliate cells because intercellular attachments are better preserved and the degree of morphologic departure from normal is smaller, complicating cytologic recognition.[70] This results in poor sensitivity in low-grade and early stage cancers.

Several other factors affect the sensitivity of cytology, including specimen quality, number of exfoliated cells, and pathologist expertise. However, the overall low sensitivity of cytology is due to its low sensitivity in detecting low-grade bladder tumors.[71] Urine cytology is associated with a significant false-negative rate, especially for low-grade carcinoma (10-50% accuracy rate). The false-positive rate is 1-12%, although cytology has a 95% accuracy rate for diagnosing high-grade carcinoma and CIS.

Urine cytology is often the test used for diagnosis of CIS. Suggestive urine cytology findings encourage the urologist to perform a bladder biopsy. With a properly collected urine sample that is promptly placed into fixative, CIS is detected in 70-75% of patients.

Instrumentation may cause reactive cellular changes, contributing to variability in interpretation. False-positive reports of malignant cells are uncommon, but ambiguous reports of atypical cells are frequent.

Improving accuracy

Perform urine cytology at the same time as cystoscopy, although its routine use for screening is controversial. If the cystoscopic examination yields normal findings but the urine cytology result is positive, further evaluation should include an upper tract study and random biopsies of the bladder. Obtain biopsy samples of the prostatic urethra in men.

Fluorescence in situ hybridization (FISH) may improve the sensitivity and specificity of routine cytology. The US Food and Drug Administration (FDA) has approved a FISH assay for the detection of recurrent bladder cancer in voided urine specimens from patients with a history of bladder cancer, as well as for the detection of bladder cancer in voided urine specimens from patients with gross or microscopic hematuria but no previous history of bladder cancer.[72]

Cytoimmunologic techniques have been developed using cytokeratin 20 as a target molecule. This assay may be more sensitive than conventional cytology, although the ability to detect low-grade tumors tends to be poor in all cytologic examinations. In contrast, the positive predictive value in patients with CIS tends to be around 75%.

Urine Tumor Markers

Noninvasive urine markers can offer an alternative to the standard means of detecting bladder cancer or can be used as an adjunct to cystoscopy.[73] Over 30 urinary biomarkers have been reported for use in bladder cancer diagnosis, but only a few are commercially available.[74]

The European Association of Urology (EAU) guidelines on non–muscle-invasive bladder cancer  that most of these tests are more sensitive than cytology but are less specific, and none have been accepted for diagnosis or follow-up in routine urologic practice or in guidelines.[75]  The National Comprehensive Cancer Network's (NCCN) Clinical Practice Guidelines in Oncology: Bladder Cancer generally agree with the EAU guidelines but note that urinary biomarker tests approved by the U.S. Food and Drug Administration may be considered for use in monitoring for recurrence; however, this is a category 2B recommendation (ie, based on lower-level evidence).[1]

See Urine Tumor Markers in Bladder Cancer Diagnosis for more information on this topic.

Cystoscopy

Cystoscopy is the primary modality for the diagnosis of bladder carcinoma because of its low risk and because biopsy specimens can be taken and papillary tumors resected during a single procedure. The EAU guidelines on non–muscle-invasive bladder cancer state that cystoscopy should be performed in all patients with symptoms of possible bladder cancer and that no noninvasive test can take its place.[75]

However, cystoscopy may be an embarrassing procedure for the patient because of exposure and handling of the genitalia. The procedure must therefore be performed with respect, and the patient should remain exposed only as long as necessary to complete the evaluation.

Men are most easily evaluated with a flexible cystoscope. In women, cystoscopy can be performed as described for men, using a flexible cystoscope, although, because the female urethra is relatively straight, a rigid cystoscope may be used instead.

See Cystoscopy for more information on this topic.

Diagnostic Strategy

In the setting of findings that are negative for cystoscopy and urinary cytology but positive for FISH, 2 possible scenarios arise. One is that the FISH result is falsely positive. The other is that it is an anticipatory positive result; in such cases, the patient has a 30% chance of developing a bladder tumor over 2 years. Patients in this category should undergo surveillance with increased frequency (see Table 1, below).

Table 1. Clinical Findings and Recommended Action in Patients with Negative Cystoscopy



View Table

See Table

Histologic Findings

In North America, South America, Europe, and Asia, more than 90% of bladder cancers are TCCs. Approximately 5% are squamous cell carcinomas (SCCs), and less than 2% are adenocarcinomas.

Carcinoma in situ

The typical visual appearance of CIS is that of a flat carcinoma extending along the surface of the bladder. This is in contrast to a papillary tumor, which extends on a stalk into the lumen of the bladder. CIS, by definition, does not invade through the basement membrane into the lamina propria. When it does, the cancer is considered to behave as an aggressive TCC and is managed accordingly.

The histologic pattern of CIS is characterized by bizarre, abnormal cells in the epithelial layer. The cells appear to be those of high-grade cancer; thus, they are readily detected in cytology specimens.

The pathologist may have difficulty distinguishing between cellular atypia and CIS. A consultant should review the slides if the pathologist is uncertain or diagnoses atypia. Upon further review, these cases usually prove to be CIS. The distinction is important because CIS requires therapy while atypia can be managed with observation. Finally, some pathologists attempt to grade CIS; however, CIS is not graded. An associated papillary tumor would be graded as low or high grade.

For more information, see Pathology of Urinary Bladder Squamous Cell Carcinoma and Pathologic Findings in Small Cell Bladder Carcinoma.

Computed Tomography Scanning

Upper tract imaging is necessary for the hematuria workup. The imaging modality chosen should be able to visualize the kidneys and the urothelium.

American Urologic Association Best Practice Policy recommends CT scanning of the abdomen and pelvis with contrast, with preinfusion and postinfusion phases.[76] This evaluation is ideally performed with CT urography, using multidetector CT, or it can be performed with a single-detector CT-scan study followed by an excretory radiographic study of the kidneys, ureters, and bladder (KUB) to obtain images similar to those produced with IVP.

Conduct retrograde pyelography in patients in whom contrast CT scanning cannot be performed because of azotemia or a severe allergy to intravenous contrast.

Intravenous Pyelography

IVP is the traditional standard for upper tract urothelium imaging; however, it is a poor modality for evaluating the renal parenchyma. Few centers in the United States perform IVP today, although the test is still included in the NCCN guidelines as an acceptable modality for imaging the upper tract collecting system.[1] The EAU guideline recommends CT urography as more informative than IVP for upper urinary tract tumors.[75]

Renal Ultrasonography

Ultrasonography is also commonly used in the diagnosis of bladder cancer. However, urothelial tumors of the upper tract and small stones are easily missed. The EAU guideline states that ultrasonography is useful for identifying obstruction in patients with hematuria. It can detect renal masses, hydronephrosis, and bladder intraluminal masses but cannot rule out all potential causes of hematuria. It cannot reliably exclude the presence of upper tract urothelial carcinomas and cannot replace CT urography.[75]

Staging

The International Union Against Cancer and the American Joint Committee on Cancer Staging developed the tumor, node, and metastases (TNM) staging system, which is used to stage bladder cancer (see below).[77] Ta and T1 tumors and CIS were once considered superficial bladder tumors. T2, T3, and T4 tumors were traditionally described as invasive bladder cancer. However, urologic oncologists now recommend avoiding the term superficial bladder cancer to describe Ta, T1, and CIS tumors because it is a misnomer and tends to group together patients who may require different treatments and who may have different prognoses.

Urothelial carcinoma is histologically graded as low grade (formerly graded 1-2) or high grade (formerly graded 3). CIS is characterized by full mucosal thickness and high-grade dysplasia of the bladder epithelium and is associated with a poorer prognosis.

The following is the TNM staging system for bladder cancer:

See Bladder Cancer Staging for more information on this topic.

More than 70% of all newly diagnosed bladder cancers are non–muscle invasive, approximately 50-70% are Ta, 20-30% are T1, and 10% are CIS. Approximately 5% of patients present with metastatic disease, which commonly involves the lymph nodes, lung, liver, bone, and central nervous system. Approximately 25% of affected patients have muscle-invasive disease at diagnosis.

Clinically stage a patient who has muscle-invasive disease with CT scanning of the abdomen and pelvis,[78] chest radiography, and serum chemistries. If the patient is asymptomatic with normal calcium and alkaline phosphatase levels, a bone scan is unnecessary.

Metastasis

As many as 50% of patients with muscle-invasive bladder cancer may have occult metastases that become clinically apparent within 5 years of initial diagnosis. Most patients with overt metastatic disease die within 2 years despite chemotherapy. Approximately 25-30% of patients with only limited regional lymph node metastasis discovered during cystectomy and pelvic lymph node dissection may survive beyond 5 years.

Grade

Stage and grade are critical to the likelihood of cancer recurrence and progression in persons with bladder cancer who are treated with local therapy. Using the American Joint Committee on Cancer staging system combined with grade, tumors may be classified using a T-G system of labeling. For example, a Ta tumor that is grade 2 (intermediate differentiation) is described as Ta-G2. The extremes are Ta-G1 (low stage, low grade) to T1-G3 (invading the lamina propria, high grade), with correspondingly favorable or unfavorable prognoses.

In 2004, the International Society of Urologic Pathologists and World Health Organization eliminated grade 2 and adopted low-grade or high-grade designations. Papillary urothelial neoplasia of low malignant potential (PUNLMP) has also been added as a designation.

CIS, which is defined as a flat, high-grade, noninvasive cancer, is an exception to the above concept. Although some are tempted to consider CIS a premalignant condition, in reality it is an aggressive form of cancer that is detected prior to invasion. Therefore, aggressive management and surveillance are warranted. Likewise, the opportunity to affect CIS-associated mortality is significant because this type of cancer may respond to conservative therapy. If left untreated, however, CIS eventually becomes invasive and progresses.

In addition, a move is developing toward classifying such cancers as either high grade or low grade instead of using the multiple-level classification that has been employed in the past. Regardless, the correlation between stage and grade is significant.

Designation of CIS

By definition, CIS is confined to the epithelial surface of the urinary tract and has no other official stage definition. It grows along the inner surface of the bladder and is therefore considered to be a form of non–muscle-invasive bladder cancer, but this specific designation is reserved for papillary or solid TCCs.

In another commonly used staging system, tumors are Ta or T0 if they are confined to the epithelial layer and T1 if present in the lamina propria. However, CIS is just CIS. If there is evidence of tumor cell invasion into the lamina propria or deeper into the muscle, the cancer is considered to be a TCC and the designation of CIS is no longer used.

Approach Considerations

The treatment of non–muscle-invasive (Ta, T1, carcinoma in situ [CIS]) and muscle-invasive bladder cancer should be differentiated. Treatments within each category include surgical and medical approaches. European Association of Urology (EAU) and National Comprehensive Cancer Network (NCCN) guidelines for non–muscle-invasive cancer strongly recommend stratifying risk of recurrence and progression and using risk tables to determine appropriate treatment.[1, 75]

The two principal treatment choices in muscle-invasive bladder cancer are radical cystectomy and transurethral resection of bladder tumor (TURBT) followed by concurrent radiation therapy and systemic chemotherapy (trimodality therapy). Each choice has its advocates.

Chedgy and Black propose that radical cystectomy should be considered the gold-standard treatment for muscle-invasive bladder cancer. They cite recommendations from European and United States guidelines, as well as published literature showing a 75% 5-year cancer-specific survival for all stages of bladder cancer treated with cystectomy, while noting that the published literature on trimodality therapy shows evidence of inferior survival and frequent treatment failure, with almost one-third of patients eventually requiring a salvage cystectomy.[79]

Nevertheless, Chedgy and Black consider cystectomy and trimodality therapy to be complementary. They observe that many patients considered ineligible for radical cystectomy may be candidates for trimodality therapy, especially radiosensitization is performed with 5-fluorouracil and mitomycin.[79]

In contrast, Mitin recommends considering trimodality therapy as the first option for patients with muscle-invasive bladder cancer, with cystectomy reserved for patients who are unable or unwilling to undergo bladder preservation or for salvage in the case of local recurrence. Mitin cites literature demonstrating similar or better outcomes compared with cystectomy and excellent quality of life, with low rates of radiation-induced adverse effects.[80]

Trimodality therapy carries a significant local recurrence rate, however, which necessitates thorough and frequent cystoscopic follow-up. Local recurrence requiring salvage cystectomy is usually identified within the first 3 years. Morbidity and mortality rates with salvage cystectomy are remarkably similar to those with first-line radical cystectomy, but reconstructive options with salvage cystectomy may be limited by the presence of irradiated bowel, which may be unacceptable for a continent reservoir or a neobladder creation.[80]

In the future, treatment selection for muscle-invasive bladder cancer is likely to be based on testing of tumors for biomarkers that indicate treatment sensitivity. Patients with resistant tumors would be offered upfront cystectomy, while those with chemoradiation-sensitive tumors would be offered bladder-preserving therapy, with regimens selected on the basis of genetic analysis.[80] For example, presence of the MRE11A single-nucleotide polymorphism rs1805363 has been associated with worse cancer-specific survival after radiation therapy, with a gene-dosage effect observed, but not after cystectomy.[81]

Immunotherapy and chemotherapy

Patients with low-grade, low-stage disease may receive expectant treatment or may benefit from a single instillation of intravesical chemotherapy. Both guidelines also recommend 1 immediate instillation of chemotherapy as the entire adjuvant treatment for patients at low risk of recurrence and progression.[1, 75]

Bacillus Calmette-Guérin (BCG) immunotherapy or other intravesical chemotherapies may be used for patients with recurrent disease or those at intermediate risk, although they are not necessary for all high-risk patients. Patients with T1-high grade or CIS are advised to undergo intravesical BCG immunotherapy because of the substantial risk of disease recurrence and progression.[82, 83, 84, 85, 86, 87]

The EAU guidelines recommend that patients with intermediate- and high-risk tumours receive intravesical bacillus Calmette-Guérin (BCG) after TURBT to reduce the risk of tumor recurrence. For optimal efficacy, BCG must be given in a maintenance schedule and a three-year maintenance is more effective than one year to prevent recurrence in patients with high-risk tumours, but not in patients with intermediate-risk tumours.[75]

In May 2016, the FDA granted accelerated approval of atezolizumab, the first cancer immunotherapy that acts as an inhibitor of  programmed cell death ligand 1 (PD-L1) for the treatment of urothelial carcinoma. Nivolumab, another PD-L1 inhibitor, also was approved by the FDA in February 2017. For more information, see Chemotherapeutic Regimens for Metastatic Bladder Cancer, below.

Surgery and radiation therapy

Endoscopic treatment with transurethral resection of bladder tumor (TURBT) is the first-line treatment to diagnose, stage, and treat visible tumors. TURBT is not effective for CIS, because the disease is often so diffuse and difficult to visualize that complete surgical removal may not be feasible. It is critically important to surgically remove all non–muscle-invasive disease prior to beginning intravesical therapy. When a combination of papillary tumor and CIS is present, the papillary tumor is removed before treatment of the CIS is initiated.

The EAU guidelines recommend the use of fluorescence-guided resection, as it is more sensitive than conventional white-light cystoscopy for detection of tumors.[88, 89, 90] The added detection rate with fluorescence-guided cystoscopy is 20% for all tumors and 23% for CIS.[91] The FDA has approved the use of blue-light cystoscopy with 5-aminolevulinic acid (Cysview) in patients suspected or known to have non–muscle-invasive bladder cancer on the basis of prior cystoscopy.[92]

As many as 20% of patients initially diagnosed with CIS may have unrecognized invasion beyond the lamina propria. Thus, they may not respond to intravesical therapy. These patients are candidates for radical cystectomy or radiation therapy and/or chemotherapy. Radiation therapy with or without chemotherapy is of limited benefit in patients with pure CIS but can be useful in some patients with muscle-invasive transitional cell carcinoma (TCC).

The criterion standard for the treatment of patients with stage T2-T4 disease is radical cystoprostatectomy for men and anterior pelvic exenteration for women. Additionally, all patients should undergo bilateral pelvic lymphadenectomy.

Level 1 evidence supports the use of preoperative (neoadjuvant) chemotherapy in patients with muscle-invasive bladder cancer. In addition, a series of studies have shown substantial benefit for adjuvant chemotherapy in these patients. booth and Tannock have noted that few patients in North America receive neoadjuvant therapy, however, and suggest that neoadjuvant or adjuvant therapy should be provided to all patients with muscle-invasive bladder cancer who are sufficiently fit to receive it.[93]

Lymph node dissection

Much controversy exists regarding the optimal extent of the lymph node dissection that should accompany cystectomy. It is clear that at minimum, a meticulous standard dissection should be performed.

While a number of retrospective studies demonstrate no difference in overall survival with standard versus extended lymph node dissection, a growing body of evidence suggests that more extended node dissection may improve survival in lymph node–positive and lymph node–negative disease. The first prospective, randomized trial to address this question, sponsored by the Southwest Oncology Group (SWOG), is currently enrolling patients.[94]

Small cell carcinoma

The treatment of localized small cell carcinoma is neoadjuvant chemotherapy followed by radical cystectomy or external beam radiation therapy. Chemotherapy using a platinum-based protocol is applied to metastatic disease. In addition, adjuvant therapy may be used in cases of stage III and IV disease that were treated with radical cystectomy.[49, 60, 95]

Adenocarcinoma

Adenocarcinomas respond poorly to radiation and chemotherapy. Radical cystectomy is the treatment of choice. Lymphomas may be effectively treated with chemotherapy or radiation.

Squamous cell carcinoma

For patients with squamous cell carcinoma (SCC) and those with squamous differentiation, a study by Ehdaie et al found no difference between the 2 groups regarding cancer-specific and overall survival following treatment with radical cystectomy and pelvic lymph node dissection.[96]

See Bladder Cancer Treatment Protocols for more information on this topic.

Lifestyle measures

Smoking cessation decreases the risk of tumor recurrence and progression and improves overall health. Increased water intake has been advocated because it may help to dilute carcinogens and decrease the exposure of the urothelium to them, but conclusive benefit has not been shown. Multivitamin or vitamin A supplementation has also been advocated, but data do not fully support this practice.[97]

Treatment of Non–Muscle-Invasive Disease (Ta, T1, CIS)

Immunotherapy and chemotherapy

Intravesical instillation of BCG is used in the treatment of high-risk Ta, T1, and CIS urothelial carcinoma of the bladder. Immunotherapy with BCG is the most effective intravesical therapy for CIS and T1 tumors. It is less effective in reducing the 5-year recurrence rate for low-grade and low-stage urothelial carcinoma (see Table 2, below).

Table 2. Recurrence and Progression Rates at 5 Years for Ta, T1, and CIS TCC of the Bladder Treated With BCG



View Table

See Table

The intravesical instillation of either BCG vaccine or chemotherapy is initiated approximately 2-4 weeks following endoscopic resection of any visible papillary tumors or bladder biopsies. By that time, the bladder has usually healed enough to avoid systemic distribution of the vaccine organism.

See Bacillus Calmette-Guérin Immunotherapy for Bladder Cancer for more information on this topic.

Interferon alfa or gamma has been used in the treatment of stages Ta and T1 and CIS urothelial carcinoma, either as single-agent therapy or in combination with BCG.[98] Its role has primarily been in treatment following BCG failure. Early results in nonrandomized, retrospective series have reported a 42% response with tolerable adverse effects after BCG failure. However, no evidence has indicated that retreatment with BCG with interferon is superior to retreatment with BCG alone.

Patients who have a recurrence within 12 months after 2 courses of BCG (preferably 6+3 treatments) do not benefit from treatment with BCG plus interferon. In addition, in a randomized study of BCG versus BCG plus interferon in BCG-naive high-risk patients, the addition of interferon was equivalent to BCG alone.[99]

Kamat et al found that the results of fluorescence in situ hybridization (FISH) assays can identify patients at risk for tumor recurrence and progression who are undergoing BCG immunotherapy. This information could be useful in counseling patients about alternative treatment strategies.[100]

Patients with BCG-refractory CIS may also be treated with intravesical valrubicin (Valstar), which is currently the only FDA-approved agent for this particular indication. However, any patient who has persistent or recurrent disease after BCG should be considered for radical cystectomy, given the high rate of disease progression.

Intravesical docetaxel appears to be a promising agent for BCG-refractory non–muscle-invasive bladder cancer; adding maintenance treatments of docetaxel may increase the duration of recurrence-free survival. Barlow et al reported that 32 of 54 patients with BCG-refractory bladder cancer showed a complete response to 6 weekly treatments of intravesical docetaxel.[101] Median time to recurrence was 39.3 months in responders treated with maintenance docetaxel, compared with 19 months in those who did not receive maintenance therapy.

Surgery

TURBT

Endoscopic TURBT is the first-line means of diagnosing, staging, and treating visible tumors. Electrocautery or laser fulguration of the bladder tumor is sufficient for low-grade, small-volume, papillary tumors. However, the EAU guidelines recommend resection of small tumors (< 1 cm) in a single piece that includes part of the underlying bladder wall.[75]

The EAU and NCCN guidelines offer similar recommendations for surgical treatment.[1, 75] Patients with bulky, high-grade, or multifocal tumors should undergo a second procedure to ensure complete resection and accurate staging 2-6 weeks after the initial TURBT.

Both guidelines state that a second resection should be performed at this time if these or other factors, such as an absence of muscle tissue in the initial specimen, indicate that the initial TURBT was incomplete. Resection of large tumors (>1 cm diameter) should be performed in fractions, including muscle tissue.[1, 75] Approximately 30% of stage T1 tumors are upgraded to muscle-invasive disease.

Fluorescence-guided resection

The EAU guidelines recommend fluorescence-guided resection, as it is more sensitive than white-light cystoscopy alone for detection of tumors, particularly CIS.[102, 103, 104] The FDA has approved blue-light cystoscopy with hexaminolevulinate (Cysview) as an adjunct to white-light cystoscopy in patients suspected or known to have non–muscle-invasive papillary cancer of the bladder on the basis of a prior cystoscopy. This technique is not a replacement for random bladder biopsies or other procedures used in the detection of bladder cancer and is not for repetitive use.

Blue-light cystoscopy with hexaminolevulinate detects more Ta/T1 bladder cancer lesions than does white-light cystoscopy alone.[102, 103, 104, 105, 106] (See the image below.) Stenzl et al reported that in patients with Ta or T1 tumors, at least one of the tumors was seen only with fluorescent cystoscopy in 16% of patients.[107] Improved detection leads to improved tumor resection, as every tumor detected is resected in the same TURBT.[108]



View Image

(A) When infused into the bladder, the optical imaging agent hexaminolevulinate (Cysview) accumulates preferentially in malignant cells. (B) On blue-l....

No further metastatic workup is needed for obviously superficial tumors. Because bladder cancer is a polyclonal field change defect, continued surveillance is mandatory.

See Transurethral Resection of Bladder Tumors for more information on this topic.

Radical cystectomy

Although radical cystectomy is typically reserved for muscle-invasive disease, it is also appropriately used to treat some patients with high-risk, non–muscle-invasive bladder cancer, including CIS. Indications in non ̶ muscle-invasive disease include the following:

Eliminating visible lesions with resection is preferable prior to intravesical BCG, but some CIS lesions may not be readily visible. Blue-light cystoscopy may improve the detection of CIS.[107] Patients who do not respond to BCG instillations often find cystectomy difficult to accept and, instead, want to continue trying various intravesical instillations.

The difficulty of accurately staging CIS preoperatively was demonstrated by Tilki and a group of international investigators.[109] These researchers reported that of 243 patients who were considered to have only CIS before cystectomy, only 117 (48.1%) were found to actually have CIS; 20 patients (8.2%) had no cancer (pT0), and 19 patients (7.8%) had urothelial cancer only. The disease was up-staged in 36% of the patients. The overall 5-year recurrence-free and cancer-specific survival was 74% and 85%, respectively.

From 35-50% of patients who undergo cystectomy for Ta, T1, or CIS are discovered to have muscle-invasive disease, with 10-15% demonstrating microscopic lymph node metastasis. According to the NCCN guidelines, cystectomy should involve at least bilateral node dissection, including iliac and obturator nodes.[1]

Patients with T1 high-grade cancer in association with diffuse CIS are at especially high risk of progression, and they may be treated with early cystectomy based on a decision made by the physician and patient. The EAU guidelines recommend that immediate cystectomy be considered for such patients.[75]

CIS progresses to muscle-invasive disease in upwards of 80% of affected patients, with 20% of patients found to have muscle-invasive disease at the time of cystectomy. High-grade T1 tumors that recur despite BCG have a 50% likelihood of progressing to muscle-invasive disease. Cystectomy performed prior to progression yields a 90% 5-year survival rate. The 5-year survival rate drops to 30-50% in muscle-invasive disease. The EAU guidelines strongly advocate cystectomy in patients with early BCG failure.[75]

Treatment of Muscle-Invasive Disease (T2 and Greater)

Radical cystectomy

The criterion standard for the treatment of patients with stage T2-T4 disease is radical cystoprostatectomy for men and anterior pelvic exenteration for women.

Cystoprostatectomy involves removal of the bladder, peritoneal covering, perivesical fat, distal ureters, prostate, seminal vesicles, vas deferentia, and, sometimes, the membranous or entire urethra. Anterior pelvic exenteration consists of cystectomy, urethrectomy, hysterectomy, salpingo-oophorectomy, and partial anterior vaginectomy. Both procedures also include regional lymph node dissection.

In experienced hands, robot-assisted radical cystectomy may offer the advantages of reduced blood loss, opiate requirement, and hospital stay. As this is a relatively new procedure, surgeons performing it need to provide detailed informed consent and a full description of potential complications and outcomes.[110]

Emerging retrospective data from multiple institutions suggest that a longer interval from the time of diagnosis to radical cystectomy can adversely affect pathologic stage and survival.[111] For example, at the University of Pennsylvania, patients who underwent radical cystectomy within 12 weeks of the diagnosis had a lower incidence of advanced pathologic stage (42% vs 84% with extravesical disease), lower incidence of positive lymph nodes, and an increased 3-year survival rate (62% vs 35% with extravesical disease).[112]

Pelvic lymphadenectomy

Approximately 25% of patients undergoing radical cystectomy have lymph node metastases at the time of surgery. Bilateral pelvic lymphadenectomy (PLND) should be performed in conjunction with radical cystoprostatectomy and anterior pelvic exenteration. PLND adds prognostic information by appropriately staging the patient and may confer a therapeutic benefit.

PLND can be performed in a standard or an extended version. The boundaries of a standard PLND include the bifurcation of the common iliac artery and vein superiorly, the genitofemoral nerve laterally, the obturator fossa posteriorly, and the circumflex iliac vein (or node of Cloquet) inferiorly.

Extended PLND includes the lymph nodes in the presacral region and those surrounding the common iliac vessels to the level of the aortic bifurcation. For a supra-extended PLND, dissection can be continued to the level of the inferior mesenteric artery. Patients with lymph node metastases rarely have “skip lesions” (ie, positive nodes in the extended dissection with negative pelvic lymph nodes).

The additional benefit of an extended PLND is controversial. On the basis of several retrospective studies, some experts believe that an extended dissection provides additional staging information and offers a survival benefit. However, no randomized trials to date have proved that an extended PLND is more beneficial than the standard procedure. A prospective, randomized trial comparing standard with extended pelvic lymphadenectomy is currently recruiting participants.[94]

Urinary diversion

After cystectomy is performed, a urinary diversion must be created from an intestinal segment. Diversions can be incontinent or continent. Contraindications to performing continent urinary diversions are as follows:

Incontinent urinary diversion

Conduits can be constructed from either ileum or colon. The most common incontinent diversion is the ileal conduit (see the image below), which has been used for more than 40 years with excellent reliability and minimal morbidity.



View Image

In an ileal conduit, a small segment of ileum is taken out of continuity with the gastrointestinal tract but is maintained on its mesentery. Ureters a....

In this procedure, a small segment of ileum (at least 15 cm proximal to the ileocecal valve) is taken out of gastrointestinal continuity but maintained on its mesentery, with care to preserve its blood supply. The gastrointestinal tract is restored with a small-bowel anastomosis. The ureters are anastomosed to an end or side of this intestinal segment and the other end is brought out as a stoma to the abdominal wall. Urine continuously collects in an external collection device worn over the stoma.

Continent urinary diversion

The most commonly used continent cutaneous urinary diversion is the Indiana pouch (see the image below). Introduced in 1987, the Indiana pouch is a urinary reservoir created from a detubularized right colon and an efferent limb of terminal ileum. The terminal ileum is plicated and brought to the abdominal wall. The ileocecal valve acts as a continence mechanism. The Indiana pouch is emptied with a clean, intermittent catheterization 4-6 times per day.



View Image

In an Indiana pouch, a urinary reservoir is created from detubularized right colon and an efferent limb of terminal ileum. Terminal ileum is plicated ....

The orthotopic neobladder is another form of continent urinary diversion. In neobladder diversions (see the image below), various segments of intestine, including the ileum, ileum and colon, and sigmoid colon, can be used to construct a reservoir. The ureters are implanted to the reservoir, and the reservoir is anastomosed to the urethra.



View Image

In an orthotopic neobladder, a segment of ileum is used to construct a neobladder, which is connected to the urethra. Orthotopic neobladder most close....

Neobladder diversions have been performed successfully in men for more than 20 years and, more recently, in women. The orthotopic neobladder most closely restores the natural storage and voiding function of the native bladder. Patients have volitional control of urination and void by Valsalva.

A variety of other continent urinary reservoirs have been developed. These vary primarily in the continence mechanisms utilized.

Neoadjuvant chemotherapy in muscle-invasive disease

Cisplatin-based neoadjuvant chemotherapy has become the standard of care in muscle-invasive bladder cancer.[66] Giving chemotherapy prior to radical cystectomy may improve cancer-specific survival, presumably by treating micrometastatic disease and pathologic downstaging.[113] A meta-analysis of 11 trials showed an overall survival rate benefit of 5% in patients who received neoadjuvant chemotherapy.[114]

If locally advanced TCC is suspected, based on clinical staging, the rationale for neoadjuvant chemotherapy prior to cystectomy may be even stronger. In a multicenter, randomized, prospective study by the Southwestern Oncology Group (SWOG) of neoadjuvant therapy with a combination of methotrexate, vinblastine, doxorubicin, and cisplatin, the investigators concluded that neoadjuvant therapy conferred a treatment benefit compared with surgery alone for locally advanced bladder cancer.[115]

However, several criticisms of this study exist. The study was underpowered because of slow recruitment (317 patients over 11 y), because 20% of the patients who were to undergo cystectomy alone never had the surgery, and because there was no comparison to neoadjuvant therapy alone or adjuvant therapy. In addition, a study that reevaluated the SWOG data found that surgical factors significantly affected outcomes.[116]

Patients with P3-P4 or N+ urothelial carcinoma in the United States are typically advised to receive adjuvant chemotherapy. In one small series, the T4 tumors of 45% of affected patients responded to chemotherapy, making potentially curative cystectomy possible.

A phase III trial that assessed 976 patients with muscle-invasive bladder cancer using neoadjuvant cisplatin, methotrexate, and vinblastine (CMV) chemotherapy found that risk of death was decreased by 16%. Chemotherapy was followed by cystectomy and/or radiotherapy.[117]

Advanced urothelial carcinoma that develops resistance to platinum-based chemotherapy has often also developed resistance to inhibitors of the epithelial growth factor receptor (EGFR) family of receptor tyrosine kinases. However, a review by Mooso et al notes that EGFR family inhibitors such as erlotinib may be of use in patients with no prior chemotherapy in whom EGFR or ERBB2 is over expressed.[118]

Adjuvant chemotherapy in muscle-invasive bladder cancer

Although the evidence supporting adjuvant chemotherapy is less compelling than that for neoadjuvant chemotherapy, some patients may benefıt from adjuvant chemotherapy, such as those who underwent up-front radical cystectomy and have extensive tumor invasion of the bladder wall or lymph node involvement.[66]

A phase III trial in 284 patients with pT3 to T4 or node-positive bladder cancer found insignificant improvement in overall survival with adjuvant cisplatin-based chemotherapy given within 90 days after cystectomy, compared with deferral of chemotherapy until relapse: after a median follow-up of 7 years, 66 of 141 patients (47%%) in the immediate chemotherapy arm had died, compared with 82 out of 143 (57%) in the deferred chemotherapy arm (P=0.13). However, patients who received iimmediate adjuvant chemotherapy had significantly longer progression-free survival (PFS): 5-year PFS was 47.6% in immediate chemotherapy arm versus 31.8% in the deferred treatment arm (P< 0.0001).[119]

This study did not meet its accrual goal of 644 patients and was terminated early. Nevertheless, it remains the largest randomized trialof adjuvant chemotherapy to date.[66]

Systemic Regimens for Metastatic Bladder Cancer

First-line, platinum-based combinations are active in locally advanced and metastatic urothelial carcinoma. However, long-term outcomes, including disease-specific and overall survival, remain suboptimal.

Methotrexate, vinblastine, doxorubicin (Adriamycin), and cisplatin (MVAC) is a standard combination regimen for treatment of metastatic bladder cancer. MVAC has an objective response rate of 57-70% and a complete response rate of 15-20%. Median overall patient survival with this regimen is typically 13- 15 months, and the 2-year survival rate is 15-20%.[120, 121, 122]

Gemcitabine and cisplatin (GC) is a newer regimen that has been shown to be as effective as MVAC but with less toxicity.[123] In particular, less nephrotoxicity is seen with GC than with MVAC.[124] GC is now considered a first-line treatment for bladder cancer. Unfortunately, about 40-50% of patients with advanced urothelial carcinoma have coexisting medical issues that preclude the use of cisplatin-based therapy.

Programmed cell death inhibitors

Agents that inhibit programmed cell death 1 (PD-1) protein and its ligands PD-L1 and PD-L2, which are part of immune checkpoint pathways that regulate T-cell activation to escape antitumor immunity, have entered clinical practice as second-line therapy for metastatic urothelial carcinoma, and are beginning to establish a role as first-line agents in patients who are not candidates for cisplatin chemotherapy. Agents in this category include atezolizumab, nivolumab, durvalumab, avelumab, and pembrolizumab.

Previously, second-line therapy had represented a significant unmet medical need. While vinflunine, a vinca alkaloid, was approved in Europe based on the results of a phase III trial versus best supportive care in the second-line setting, its efficacy is marginal.[125] The median survival for patients treated with second-line therapy such as vinflunine or other agents, including the taxanes and pemetrexed, is only 6-9 months.

Atezolizumab

Atezolizumab (Tecentriq) is the first PD-L1 inhibitor approved for the treatment of urothelial carcinoma. In May 2016, the US Food and Drug Administration (FDA) granted accelerated approval of atezolizumab for locally advanced or metastatic urothelial carcinoma in patients who have disease progression during or following platinum-containing chemotherapy, or disease progression within 12 months of neoadjuvant or adjuvant treatment with platinum-containing chemotherapy.[126] In April 2017, atezolizumab’s indication was expanded to include use as initial treatment of locally advanced or metastatic urothelial carcinoma in patients who are not eligible for cisplatin chemotherapy.

Approval of atezolizumab was based on the phase 2 IMvigor 210 trial (n=310), an open-label, multicenter, single-arm study with two cohorts. In the previously treated patients (cohort 2), the objective response rate (ORR) was 26% for the subgroup with the highest positivity for PD-L1, 18% for the subgroup with lower positivity, and 15% for all patients. Median overall survival was 7.9 months for all patients, 11.4 months for the highest-positivity subgroup, and 6.7 months for the lowest-positivity subgroup. Twelve-month overall survival was 36% for all patients, 48% for the high group, and 30% for the low group.[126]

In the treatment-naive patients (cohort 1), the ORR was 23.5%. Complete responses were seen in 6.7%, and partial responses were seen in 16.8%.[127]

The FDA has also approved a complementary diagnostic, the Ventana PD-L1 (SP142) assay, which can detect PD-L1 protein expression levels on tumor-infiltrating immune cells and help physicians determine which patients may derive the most benefit from treatment with atezolizumab.

In IMvigor 211, a confirmatory phase 3 study that compared atezolizumab with chemotherapy in patients whose bladder cancer had progressed on at least one prior platinum-containing regimen, overall survival was numerically, but not significantly, longer in the atezolizumab group than in the chemotherapy group (median 15.9 vs 8.3 months, respectively). However, the safety profile of atezolizumab was superior to that of chemotherapy, with lower rates of  grade 3-4 treatment-related adverse events (20% vs 43%), and of adverse events leading to treatment discontinuation (7% vs 18%).[128]

Nivolumab

Nivolumab also gained accelerated approval for advanced or metastatic urothelial carcinoma from the FDA. Approval was based on the CheckMate-275, single-arm, phase 2 clinical trial (n=270). ORR, confirmed by an independent radiographic review committee using Response Evaluation Criteria in Solid Tumors 1.1, was 19.6% (52/265; 95% confidence index [CI]: 15.0-24.9). Seven patients had complete responses and 46 had partial responses. Estimated median response duration was 10.3 months with responses ongoing at data cutoff.[129]

Durvalumab

Durvalumab (Imfinzi) was granted accelerated approval by the FDA in April 2017. Approval was based on a single-arm trial of 182 patients with locally advanced or metastatic urothelial carcinoma whose disease progressed after prior platinum-containing chemotherapy. ORR was 17%. ORR was also analyzed by PD-L1 expression status, as measured by the approved assay. Among the 182 patients, the confirmed ORR was 26.3% in 95 patients with a high PD-L1 score and 4.1% in 73 patients with a low or negative PD-L1 score. The median response duration was not reached at the time of review (range, 0.9+ to 19.9+ months).[130]

Avelumab

Avelumab (Bavencio) was also granted accelerated approval for advanced urothelial carcinoma in patients progressing during or following platinum-containing chemotherapy in April 2017. Approval was based on the JAVELIN solid tumor trial, which comprised 2 cohorts with 44 patients and 200 patients, respectively. The data from both cohorts were pulled together and showed that 161 patients with a follow-up of >6 months had an overall response rate of 16.1%.[131]

Pembrolizumab

Pembrolizumab (Keytruda) gained approval from the FDA for urothelial carcinoma in May 2017. It is indicated for locally advanced or metastatic urothelial carcinoma (UC) in patients who are not eligible for cisplatin-containing chemotherapy. It is also indicated for patients with disease progression during or following platinum-containing chemotherapy or within 12 months of neoadjuvant or adjuvant treatment with platinum-containing chemotherapy.

Approval was based on the KEYNOTE-052 trial. As first-line treatment for patients who were ineligible to receive platinum-based chemotherapy, the ORR was 24% at the time of initial phase 2 data evaluation.[132]

The FDA approval for second-line therapy for advanced UC was based on an international, phase 3 trial (n=542). ORR was 10.3 months in the pembrolizumab recipients compared with 7.4 months in the chemotherapy group.[133]

Fibroblast Growth Factor Receptor Inhibitors

Erdafitnib

Erdafitinib inhibits fibroblast growth factor receptor (FGFR) phosphorylation and signaling, and thereby decreases cell viability in cell lines expressing FGFR genetic alterations, including point mutations, amplifications, and fusions. FGFRs regulate important biological processes including cell proliferation and differentiation, which are part of a complex signaling pathway in tumorigenesis. 

The FDA granted erdafitinib accelerated approval in April 2019 for locally advanced or metastatic urothelial carcinoma that has FGFR2 or FGFR3 genetic alterations and progressed during or following at least 1 line of prior platinum-containing chemotherapy, including within 12 months of neoadjuvant or adjuvant platinum-containing chemotherapy. Approval was based on a multicenter, open-label, single-arm study (n=87) of patients with urothelial carcinoma with FGFR3 or FGFR2 genetic alterations that had progressed on or after at least 1 prior chemotherapy. The overall response rate was 32.2%, with 2.3% having a complete response and almost 30% having a partial response.[134]

Anti-Nectin-4 Monoclonal Antibodies

Enfortumab vedotin

Enfortumab vedotin is an antibody-drug conjugate (ADC) composed of an anti–nectin-4 monoclonal antibody attached to the cell-killing agent, monomethylauristatin E (MMAE). Once the antibody attaches to nectin-4 that is expressed on the tumor, the complex is internalized in the lysosome, which releases MMAE. 

The FDA approved enfortumab vedotin in December 2019 for locally advanced or metastatic urothelial cancer in patients who have received a PD-1/L1 inhibitor and platinum-containing chemotherapy in the neoadjuvant/adjuvant, locally advanced, or metastatic setting. Approval was based on the single-arm, phase 2EV-201 global trial, which enrolled 125 patients with locally advanced or metastatic urothelial cancer who received prior treatment with a PD-1 or PD-L1 inhibitor and a platinum-based chemotherapy. The primary endpoint of confirmed ORR was 44% per blinded independent central review (55/125; 95% CI, 35.1-53.2). Among patients treated with the single agent enfortumab vedotin, 12% (15/125) experienced a complete response and 32% (40/125) experienced a partial response.[135]

Radiation Therapy

External beam radiation therapy has been shown to be inferior to radical cystectomy for the treatment of bladder cancer. The overall 5-year survival rate after treatment with external beam radiation is 20-40%, compared with a 90% 5-year survival after cystectomy for organ-confined disease. Nevertheless, external beam radiation therapy is used in various countries other than the United States for T2-T3 urothelial carcinoma of the bladder.

Neoadjuvant external beam radiation therapy has been attempted for muscle-invasive bladder cancer. However, no improvement in survival rate has been demonstrated.

In certain centers, a bladder-preserving strategy for T2-T3 urothelial carcinoma is applied using a combination of external beam radiation, chemotherapy, and endoscopic resection. Survival rates associated with this approach are comparable with those of cystectomy in selected patients. This combination has a widespread application that is limited by the complexity of the protocol, its toxicity, and a high peritreatment mortality rate of 4-5%, mostly due to nadir sepsis from the chemotherapy. In comparison, the mortality rate for most modern cystectomy series is 2-3%.

Overall, approximately 30% of patients treated with bladder-preserving therapy experience local recurrence of bladder cancer.[80] A significant number of patients ultimately require salvage cystectomy, which may be associated with  decreased options for urinary diversions.

Treatment Complications

Radical cystectomy

Radical cystectomy has a 2-3% perioperative mortality rate. However, the 6-month mortality may be as high as 7-8% in elderly patients. The two most common late complications are small-bowel obstruction and ureteroenteric stricture (see Table 3 below).

Table 3. Most Common Complications of Radical Cystectomy



View Table

See Table

The reported overall complication rate (including early and late complications) for radical cystectomy is approximately 30-40%. However, this may be an underestimation of the true complication rate because of a lack of standardized reporting in published studies.

Many patients who undergo a radical cystectomy have multiple comorbid health risk factors (eg, advanced age, cardiovascular disease, pulmonary disease). Despite these difficulties, this procedure may be performed safely even in patients older than 80 years.

After a radical cystectomy, all men are impotent if the parasympathetic nerves from the pelvic plexus (S2-S4) to the corpora cavernosum are not spared at the time of surgery. A nerve-sparing approach may be associated with potency rates of approximately 50-70%.

Urinary diversion

Complications of urinary diversion include the following:

Orthotopic neobladder complications

Owing to advances in surgical technique, this procedure is commonly used in tertiary centers. Complications include daytime and nighttime urinary incontinence at rates of approximately 5% and 20%, respectively. Urinary incontinence may develop from multiple factors, including injury to the external urethral sphincter, increased urine production from solute absorption, and relaxation of the external sphincter, which is greater at night.

Long-Term Monitoring

The high rate of disease recurrence and progression in non–muscle-invasive bladder cancer underscores the need for careful follow-up studies. According to the US National Cancer Institute, bladder cancer affects approximately 500,000 people in America. Because most still have an intact bladder, the number of patients under surveillance approaches this figure.

The EAU guidelines include schedules for follow-up cystoscopy, urinary cytology, and imaging in patients with TaT1 tumors, depending on risk of recurrence and progression. For follow-up in patients with no visible tumor in the bladder but positive cytology, the guidelines recommend biopsies and investigation of extravesical locations.[75]

For follow-up after a radical cystectomy, NCCN recommendations are as follows[1] :

For follow-up after a segmental (partial) cystectomy or bladder preservation, the NCCN recommends the same follow-up as after radical cystectomy, plus cystoscopy and urine cytology with or without selected mapping biopsy every 3 to 6 mo for 2 y, then at increasing intervals as appropriate.[1]

See Surveillance for Recurrent Bladder Cancer for more information on this topic.

Guidelines Summary

Organizations and Recommendation Approaches

Guidelines for the diagnosis and management of bladder cancer have been issusued by the following organizations:

Each organization uses a different methodology to determine levels of evidence and grades for recommendations, as follows:

Screening

In 2019, the US Preventive Services Task Force (USPSTF) concluded that the evidence was insufficient to determine the balance of benefits and harms of screening for bladder cancer in asymptomatic adults. Although adults with mild lower urinary tract symptoms (eg, urinary frequency, hesitancy, urgency, dysuria, nocturia) are not strictly asymptomatic, these symptoms are common and are not believed to be associated with an increased risk of bladder cancer. The USPSTF considered it reasonable to include these persons in the population under consideration for screening. Adults with gross hematuria or acute changes in lower urinary tract symptoms were not included in this population.[139]

No major organization recommends screening for bladder cancer in asymptomatic adults. In 2011, the American Academy of Family Physicians endorsed the USPSTF recommendation.[140]  The American Cancer Society states that prompt attention to bladder symptoms is the best approach for finding bladder cancer in its earliest, most treatable stages in persons with no known risk factors.[141]

Risk Reduction

Cigarette smoking is the most established risk factor for bladder cancer, increasing relative risk four- to five-fold relative to never smoking for current smokers and two- to three-fold for former smokers. It is the cause of 50% of cases in both men and women [22]

The European Association of Urology (EAU) guidelines estimate a 40% reduction in the risk of developing bladder cancer within 1-4 years of quitting smoking, increasing to a 60% reduction after 25 years of smoking cessation.[78]

In addition, the EAU guidelines address the second most important risk factor, occupational exposure to carcinogens, with a recommendation that workers are informed of the risk and protective measures are taken.[78]

Diagnosis

The following studies are used in the diagnosis of bladder cancer:

Novel molecular and genetic markers have been studied for the diagnosis of urothelial carcinoma (see Urine Tumor Markers in Bladder Cancer Diagnosis) but have not been shown to effectively replace urine cytology and cystoscopy. The National Comprehensive Cancer Network guidelines advise that urinary biomarker testing with FDA-approved fluorescence in situ hybridization or nuclear matrix protein 22 (NMP-22) assays may be used in the monitoring or surveillance for bladder cancer recurrence.[1]

Urinary Cytology

According to European Association of Urology (EAU) guidelines, urinary cytology is most helpful in diagnosing high-grade tumors and carcinoma in situ (CIS). Low-grade, noninvasive tumors may be missed by routine cytologic analysis. The guidelines caution that cystoscopy is still required and cannot be replaced by cytology or any other noninvasive test.[75] National Comprehensive Cancer Network (NCCN) guidelines state that urine cytology may be obtained around the time of cystoscopy.[1]

Cystoscopy

Joint guidelines issued in 2016 by the American Urological Association and Society of Urological Oncology recommend that at the time of resection of suspected bladder cancer, a clinician should perform a thorough cystoscopic examination of a patient’s entire urethra and bladder that evaluates and documents tumor size, location, configuration, number, and mucosal abnormalities. In a patient with non–muscle invasive bladder cancer (NMIBC), blue light cystoscopy, if available, should be offered at the time of TURBT to increase detection and decrease recurrence. In a patient with normal cystoscopy and positive cytology and a history of NMIBC, blue light cystoscopy (when available), prostatic urethral biopsies and upper tract imaging, as well as ureteroscopy, or random bladder biopsies should be considerd.[136]

The EAU allows for the omission of cystoscopy if the tumor was visualized with an imaging study.[75]

The NCCN also recommends cystoscopy for all patients with clinical suspicion of bladder cancer. When cystoscopy findings are negative in the setting of positive cytology findings, the NCCN guidelines recommend further evaluation of the upper urinary tract with radiographic imaging and/or ureteroscopy,[1] as well as evaluation of the prostatic urethra in men and the urethra in women.

In general, it may be preferable to obtain upper tract imaging with intravenous contrast for patients with hematuria prior to performing cystoscopy. If the imaging findings are positive, one may forgo office-based cystoscopy and perform the cystoscopy in the operating room. In addition, photodynamic diagnosis with hexaminolevulinate blue-light cystoscopy may be associated with fewer false negatives than white-light cystoscopies.[136, 107]

If cystoscopy indicates noninvasive disease, the NCCN guidelines recommend the following further studies[1] :

For imaging of the upper tract, CT urography is generally the preferred approach. Other options are an intravenous pyelogram (IVP), renal ultrasound with retrograde pyelogram, ureteroscopy, or MRI urogram. If the tumor has a purely papillary appearance or only the mucosa appears to be abnormal, suggesting carcinoma in situ (CIS), upper tract imaging can be deferred until after surgery.

If cystoscopy indicates muscle-invasive disease, the NCCN guidelines recommend the following[1] :

In patients with confirmed muscle-invasive bladder cancer, the EAU recommends CT of the chest, abdomen, and pelvis as optimal for staging. The imaging studies should include excretory-phase CT urography for complete examination of the upper urinary tract.[137]

Transurethral Resection of the Bladder Tumor

The guidelines (AUA, EAU, ESMO, NCCN) are in agreement that the final diagnosis of bladder cancer is based on cystoscopic examination and TURBT histology.[78, 136, 75, 138, 1] The guidelines further agree that all visible lesions be resected during TURBT with bimanual examination under anesthesia (EUA) and that adequate sampling is required for proper tumor identification and staging.[78, 136, 75, 138, 1]

EUA guidelines recommend performing a second TURBT 2-6 weeks after the initial resection in any of the following situations[75] :

 

Staging

Bladder cancer is staged using the International Union Against Cancer and the American Joint Committee on Cancer Staging’s tumor, node, and metastases (TNM) system.[77] See Bladder Cancer Staging.

Ta and T1 tumors and carcinoma in situ (CIS) were historically considered superficial bladder tumors. T2, T3, and T4 tumors were traditionally described as invasive bladder cancer. However, use of the term superficial bladder cancer is recommended against by the AUA/SUO, NCCN and EAU guidelines because it groups together patients who may require different treatments and who may have different prognoses. Instead, all guidelines now categorizes these tumors as non–muscle invasive bladder cancer (NMIBC).[136, 75, 138, 1]

Urothelial carcinoma is histologically graded as low grade (formerly grades 1-2) or high grade (formerly some grades 2 and all grades 3). CIS is characterized by full mucosal thickness and high-grade dysplasia or marked atypia of the bladder epithelium and is associated with a poorer prognosis. By definition, CIS is confined to the epithelial surface of the urinary tract and has no other official stage definition.

The EAU uses three levels of risk stratification for non–muscle-invasive tumors, as follows[75] :

 

 

Treatment

Non─Muscle-Invasive Disease

National Comprehensive Cancer Network (NCCN) recommendations for treatment of low-grade Ta tumors are as follows[1] :

NCCN and European Association of Urology (EAU) recommendations for treatment of high-grade Ta tumors are as follows[75, 1] :

NCCN recommendations for treatment of T1 tumors (low- and high-grade) are as follows[1] :

The European Association of Urology (EAU) recommendations for adjuvant treatment of Ta and T1 tumors are as follows[75] :

The AUA/SUO recommendations for repeat TURBT are as follows[136] :

The AUA/SUO recommendations for intravesical therapy are as follows[136] :

Carcinoma in situ

The NCCN recommendations are as follows[1] :

Follow-up and Surveillance

The European Association of Urology (EAU) guidelines recommend regular cystoscopy for follow-up of patients with Ta, T1 tumors and carcinoma in situ (CIS) who have undergone transurethral resection of the bladder tumor. Schedules are as follows[75] :

The NCCN guidelines specify that cystoscopy and urinary cytology should be performed every 3-6 months for 2 years and then at increasing intervals as appropriate.[1]

The AUA/SUO guidelines recommended surveillance schedules are as follows[136] :

Muscle-Invasive Disease

NCCN recommendations for treatment of muscle-invasive disease are as follows [1] :

For patients with advanced or metastatic disease, first-line chemotherapy regimens are as follows:

For patients who cannot receive cisplatin-based chemotherapy due to renal impairment or other comorbidities, the NCCN lists the following regimens as preferred:

For subsequent systemic therapy for locally advanced or metastatic disease stage IV), the NCCN recommends participation in clinical trials of new agents. Preferred agents are as follows:

The EAU recommends radical cystectomy as the curative treatment of choice for muscle-invasive disease.[78]  The EAU recommends pelvic lymph node dissection (LND) as a means of improving survival after radical cystectomy; however, it should be noted that no data exist to support this recommendation. Ongoing studies are assessing whether survival is influenced by the use of standard versus extended LND—ie, whether the cephalad margin should extend to the iliac arteries (standard) or to the aortic bifurcation (extended).

Further EAU recommendations for muscle-invasive disease are as follows[78] :

For patients with advanced or metastatic disease, first-line chemotherapy consists of the following cisplatin-containing combination regimens:

Second-line treatment is as follows:

European Society for Medical Oncology (ESMO) recommendations for treatment of metastatic disease are as follows[138] :

For patients who cannot receive cisplatin-based chemotherapy, palliation with carboplatin-based regimen or single-agent taxane or gemcitabine

There is insufficient evidence for the routine use of adjuvant chemotherapy; high-risk patients that did not receive neoadjuvant chemotherapy may benefit form adjuvant treatment.

When the patient is unfit for cystectomy, radiotherapy can also be offered for palliation (bleeding, pain)

Locally-advanced and metastatic disease

The updated 2019 NCCN Guidelines for bladder cancer include a new category of patients with locally advanced/metastatic bladder cancer: those who are platinum-ineligible with no PD-L1 expression. Preferred first-line regimens for these patients are atezolizumab and pembrolizumab; other options include gemcitabine and gemcitabine/paclitaxel. Patients who are cisplatin-ineligible but carboplatin-eligible should receive carboplatin over immune checkpoint inhibitors in first-line treatment.[1, 142]

 

Upper Urinary Tract Urothelial Cell Carcinoma

Diagnosis

The NCCN and EAU provide similar recommendations for the diagnosis of upper tract urothelial carcinoma (UTUC), which include cystology, cystoscopy (to rule out a concomitant bladder tumor), and CT (computed tomography) urography.[1, 137] . The EAU further recommends retrograde ureteropyelography and diagnostic ureteroscopy and biopsy.[137]

Treatment

The NCCN provides treatment recommendations based on the location and disease extent, as follows[1] :

The EUA provides conservative management recommendations, as well as recommendations for radical nephroureterectomy (RNU). The indications for RNU are as follows:[137]

Recommendations regarding RNU include the following:

The indications for conservative management are as follows:

Recommendations regarding conservative treatment include the following:

Follow-up and surveillance

The EAU recommends follow-up for at least 5 years after treatment of UTUC; however, the surveillance schedules specified are all grade C recommendations. Surveillance schedules are as follows[137] :

Medication Summary

The combination of methotrexate, vinblastine, doxorubicin (Adriamycin), and cisplatin (MVAC) is the standard treatment for metastatic bladder cancer. No proven role exists for adjuvant chemotherapy. MVAC has substantial toxicity, which must be weighed against the expected benefit. The major dose-limiting toxicity is myelosuppression.

The new combination regimens show response rates and median survival comparable to those for MVAC but with less toxicity. Gemcitabine plus cisplatin is now considered a first-line treatment for bladder cancer. Therapy with PDL1 inhibitors (eg, atezolizumab, nivolumab, durvalumab, avelumab, pembrolizumab) is now approved by the FDA for advanced urothelial carcinoma. Erdafitinib is the first fibroblast growth factor receptor (FGFR) inhibitor approved by the FDA for urothelial carcinoma in April 2019. The first anti-nectin-4 monoclonal antibody, enfortumab vedotin, was approved for urothelial carcinoma in December 2019.

Fluorouracil (Adrucil)

Clinical Context:  Fluorouracil is a pyrimidine antimetabolite. Several mechanisms of action have been proposed, including inhibition of thymidylate synthase and inhibition of RNA synthesis. This agent is also a potent radiosensitizer. Although not approved by the FDA for this indication, it is often used as a treatment for bladder cancer.

Methotrexate (Trexall, Rasuvo, Otrexup, Xatmep)

Clinical Context:  Methotrexate inhibits dihydrofolate reductase (DHFR), causing a block in the reduction of dihydrofolate to tetrahydrofolate. This inhibits the formation of thymidylate and purines and arrests DNA, RNA, and protein synthesis. It is often used as a treatment for bladder cancer, although that is not an FDA-approved indication.

Gemcitabine (Gemzar)

Clinical Context:  Gemcitabine is a pyrimidine analog. After intracellular metabolism to its active nucleotide, it inhibits ribonucleotide reductase and competes with deoxycytidine triphosphate for incorporation into DNA. Although it does not have FDA approval for this indication, it is often used as a treatment for bladder cancer. Gemcitabine is used in combination with cisplatin for the treatment of advanced or metastatic bladder cancer.

Pemetrexed (Alimta)

Clinical Context:  Pemetrexed disrupts the folate-dependant metabolic processes important for cell replication, inhibits the enzymes involved in folate metabolism and DNA synthesis, and inhibits protein synthesis. Although it does not have FDA approval for this indication, it is often used for the treatment of metastatic bladder cancer. Folic acid and vitamin B12 are typically given prior to initiation of treatment. Dexamethasone is also given with pemetrexed, to minimize cutaneous reactions.

Class Summary

These agents inhibit cell growth and proliferation. They interfere with DNA synthesis by blocking the methylation of deoxyuridylic acid.

Vinblastine

Clinical Context:  A vinca alkaloid with a cytotoxic effect (as a result of causing mitotic arrest), vinblastine binds to a specific site on tubulin, prevents polymerization of tubulin dimers, and inhibits microtubule formation. Although not FDA approved for this indication, vinblastine is often used as a treatment for bladder cancer in combination with a chemotherapy regimen.

Vinblastine is approved for intravenous use only; the FDA has issued a black box warning regarding possible death with intrathecal administration. Vinblastine is a moderate vesicant and extravasation should be avoided.

Class Summary

Vinca alkaloids act on the M and S phases of mitosis, inhibiting microtubule formation and inhibiting DNA/RNA synthesis.

Doxorubicin (Adriamycin)

Clinical Context:  Doxorubicin is an anthracycline antineoplastic that causes DNA strand breakage through effects on topoisomerase II and direct intercalation into DNA, which causes DNA polymerase inhibition. It has a labeled indication for the treatment of bladder cancer. This drug has several black box warnings, including bone marrow suppression, myocardial toxicity, and secondary malignancy.

Valrubicin (Valstar)

Clinical Context:  Valrubicin is a semisynthetic analog of doxorubicin that inhibits incorporation of nucleosides into nucleic acids. It is indicated for intravesicular treatment of bladder carcinoma in situ (CIS) that is refractory to treatment with bacillus Calmette-Guérin (BCG).

Class Summary

Anthracycline antineoplastics inhibit DNA and RNA synthesis by steric obstruction. They intercalate between DNA base pairs and trigger DNA cleavage by topoisomerase II.

Cisplatin

Clinical Context:  Cisplatin is a platinum-containing compound that exerts an antineoplastic effect by covalently binding to DNA, with preferential binding to the N-7 position of guanine and adenosine. It can react with 2 different sites on DNA to produce cross-links. The platinum complex also can bind to nuclear and cytoplasmic protein. Cisplatin has black box warnings, including anaphylactic-like reactions, ototoxicity, and renal toxicity.

Carboplatin

Clinical Context:  Carboplatin is a platinum alkylating agent that interferes with the function of DNA by producing interstrand DNA cross-links. It can be used in combination with paclitaxel for the treatment of bladder cancer, which is an off-label indication. Carboplatin has black box warnings, including bone marrow suppression, anaphylactic reactions, and vomiting.

Ifosfamide (Ifex)

Clinical Context:  Ifosfamide is a nitrogen mustard alkylating agent that inhibits DNA and protein synthesis. Although not FDA approved for this indication, ifosfamide is often used as a treatment for metastatic bladder cancer.

Thiotepa (Tepadina)

Clinical Context:  Thiotepa is an alkylating agent that inhibits DNA, RNA, and protein synthesis by producing cross-links between DNA strands. It is available as a powder for reconstitution and administration by injection. Thiotepa is indicated for the treatment of superficial papillary bladder cancer.

Class Summary

These agents inhibit cell growth and proliferation. They inhibit DNA synthesis by the formation of DNA cross-links. Alkylating agents can have serious adverse effects, including bone marrow suppression, anaphylactic-like reactions, ototoxicity, renal toxicity, and vomiting.

Docetaxel (Taxotere, Docefrez)

Clinical Context:  Docetaxel inhibits the depolymerization of tubulin, which inhibits DNA, RNA, and protein synthesis. It can be used for the treatment of bladder cancer, which is an off-label indication. It has several black box warnings, including bone marrow suppression, fluid retention, and hypersensitivity reactions. Its use is not recommended in certain patients with hepatic impairment. Patients receiving docetaxel treatment should be premedicated with corticosteroids the day before administration to help reduce fluid retention and hypersensitivity reactions.

Class Summary

These agents prevent cell growth and proliferation. They work by enhancing tubulin dimers, as well as by stabilizing existing microtubules and inhibiting their disassembly.

Atezolizumab (Tecentriq)

Clinical Context:  Monoclonal antibody to programmed cell death ligand-1 protein (PDL1). It blocks the interaction between PDL-1 and its ligands. It is indicated for locally advanced or metastatic urothelial carcinoma in patients who are not eligible for cisplatin-containing chemotherapy, or have disease progression during or following platinum-containing chemotherapy, or disease progression within 12 months of neoadjuvant or adjuvant treatment with platinum-containing chemotherapy.

Nivolumab (Opdivo)

Clinical Context:  Monoclonal antibody to programmed cell death ligand-1 protein (PDL1). It blocks the interaction between PDL-1 and its ligands. It is indicated for locally advanced or metastatic urothelial carcinoma in patients who have disease progression during or following platinum-containing chemotherapy, or disease progression within 12 months of neoadjuvant or adjuvant treatment with platinum-containing chemotherapy.

Durvalumab (Imfinzi)

Clinical Context:  Human IgG1 kappa monoclonal antibody that blocks PD-L1 binding to PD-1 and CD80, and therefore overcoming/preventing PD-L1-mediated inhibition/suppression of T-cell activation. It is indicated for locally advanced or metastatic urothelial carcinoma in patients who have disease progression during or following platinum-containing chemotherapy, or disease progression within 12 months of neoadjuvant or adjuvant treatment with platinum-containing chemotherapy.

Avelumab (Bavencio)

Clinical Context:  Avelumab is an anti-PD-L1 IgG1 monoclonal antibody.  It is indicated for locally advanced or metastatic urothelial carcinoma (UC) in patients who have disease progression during or following platinum-containing chemotherapy or disease progression within 12 months of neoadjuvant or adjuvant treatment with platinum-containing chemotherapy.

Pembrolizumab (Keytruda)

Clinical Context:  Monoclonal antibody to programmed cell death-1 protein (PD-1); blocks the interaction between PD-1 and its ligands, PD-L1 and PD-L2. Pembrolizumab is indicated as first-line treatment for locally advanced or metastatic urothelial carcinoma (UC) in patients who are not eligible for cisplatin-containing chemotherapy. It is also indicated for patients with disease progression during or following platinum-containing chemotherapy or within 12 months of neoadjuvant or adjuvant treatment with platinum-containing chemotherapy.

Class Summary

PDL1 is expressed on the surface of activated T cells under normal conditions. PDL1 interaction inhibits immune activation and reduces T-cell cytotoxic activity when bound. This negative feedback loop is essential for maintaining normal immune responses and limits T-cell activity to protect normal cells during chronic inflammation. Tumor cells may circumvent T-cell–mediated cytotoxicity by expressing PDL1 on the tumor itself or on tumor-infiltrating immune cells, resulting in the inhibition of immune-mediated killing of tumor cells.

Erdafitinib (Balversa)

Clinical Context:  Erdafitinib inhibits FGFR phosphorylation and signaling, and thereby, decreases cell viability in cell lines expressing FGFR genetic alterations, including point mutations, amplifications, and fusions. It is indicated for locally advanced or metastatic urothelial carcinoma that has FGFR2 or FGFR3 genetic alterations and progressed during or following at least 1 line of prior platinum-containing chemotherapy, including within 12 months of neoadjuvant or adjuvant platinum-containing chemotherapy.

Class Summary

Fibroblast growth factor receptor (FGFR) regulate important biological processes including cell proliferation and differentiation, which are part of a complex signaling pathway in tumorigenesis.

Enfortumab vedotin (Padcev, enfortumab vedotin-ejfv)

Clinical Context:  Enfortumab vedotin is an antibody-drug conjugate (ADC) composed of an anti-nectin-4 monoclonal antibody attached to the cell-killing agent, monomethylauristatin E (MMAE). Once the antibody attaches to nectin-4 that is expressed on the tumor, the complex is internalized in the lysosome, which releases MMAE. Enfortumab vedotin is indicated for locally advanced or metastatic urothelial cancer in patients who have received a PD-1/L1 inhibitor and platinum-containing chemotherapy in the neoadjuvant/adjuvant, locally advanced, or metastatic setting.

Class Summary

Nectin-4 is a cell adhesion molecule that is expressed on many solid tumors.

What is bladder cancer?What are the treatment options for non-muscle-invasive bladder cancer?What are some signs and symptoms of bladder cancer?Which urine studies are used to diagnose bladder cancer?What is the role of urinary cytology in the diagnosis of bladder cancer?What is the role of cystoscopy in the diagnosis of bladder cancer?What is the role of upper urinary tract imaging in the diagnosis of bladder cancer?What is the diagnostic strategy for patients with negative cystoscopy for bladder cancer?Which blood tests are performed in the diagnosis of bladder cancer?What are the treatment options for muscle-invasive bladder cancer?What are chemotherapeutic regimens used to treat metastatic bladder cancer?What is squamous cell carcinoma (SCC) bladder cancer?What is bladder cancer?What is transitional (urothelial) cell carcinoma (TCC) bladder cancer?What are the less common types of bladder cancer?What are the phenotypes of urothelial carcinoma bladder cancer?What is the clinical course of bladder cancer?How does the presentation of bladder cancer vary?What are the treatment options for carcinoma in situ (CIS) bladder cancer?How common in bladder cancer recurrence?Where can more information on bladder cancer be found?What anatomy of the bladder is helpful in understanding bladder cancer?How does the anatomy of the bladder differ in males and females?What role does the vascular system play in the anatomy of the bladder?What role does the lymphatic system play in the anatomy of the bladder?Where in the bladder does cancer originate?What is the description of bladder cancer?What terms should be used for bladder cancer?How does the WHO classify bladder cancers?What is the pathophysiology of transitional cell carcinoma (TCC) bladder cancer?What is the most common molecular biologic pathway for transitional cell carcinomas (TCCs) in the bladder?How are urothelial tumors associated with transitional cell carcinomas (TCCs) in the bladder?What is pathophysiology of squamous cell carcinoma (SCC) of the urinary bladder?What is the pathophysiology of rare forms of bladder cancer?What pathways are responsible for the development of noninvasive and invasive bladder tumors?What is the role of genetic mutation in the pathogenesis of bladder cancer?What is the role of TP53 gene alterations in the pathogenesis of bladder cancers?Which genetic alterations are common in bladder cancers?What are mechanisms that can lead to tumor progression in bladder cancer?Which factors promote tumor growth in bladder cancer?What is the role of environmental factors in the etiology of bladder cancer?Which factors increase the risk of developing bladder cancer?Which occupational exposures increase the risk of bladder cancers?How is geography a risk factor for bladder cancers?What is the role of arsenic exposure in the development of bladder cancer?What are the medical risk factors for bladder cancer?What role does genetics play in the development of bladder cancer?What is the role of schistosomiasis in the etiology of bladder cancer?How does schistosomiasis infection cause bladder cancer?What are the risk factors for squamous cell carcinoma (SCC) bladder cancer?What is the incidence of bladder cancers in the US?How does the prevalence of bladder cancer vary by sex in the US?How does the incidence of bladder cancer vary by race in the US?What are the demographic predilections of small cell carcinoma of the bladder?What is the global prevalence of bladder cancer?What are the 5-year survival rates by bladder cancer stage?What are the morbidities of untreated bladder cancer?What is the recurrence rate for superficial transitional cell carcinoma (TCC) bladder cancer?What are the most significant prognostic factors for bladder cancer?What is the prognosis of metastatic urothelial cancer?What are the risks of progression of bladder cancer by tumor grade?What is the prognosis in carcinoma in situ (CIS) bladder cancer?What is the prognosis of squamous cell carcinoma (SCC) bladder cancer?What is the prognosis in small cell carcinoma bladder cancer?What is the recurrence rate for bladder cancers?What are the risk factors for recurrence and progression of bladder cancer?What are the signs and symptoms of bladder cancer?What physical findings are characteristic of bladder cancer?What is the significance of spontaneous resolution of gross or microscopic hematuria during the evaluation of bladder cancer?In what ways do the symptoms of bladder cancer and urinary tract infection (UTI) overlap?How is carcinoma in situ (CIS) bladder cancer diagnosed?Which tests are used to diagnose bladder cancer?What are the main differential diagnoses for small cell carcinoma bladder cancer?What are the differential diagnoses for Bladder Cancer?What is the role of urine studies in the evaluation of bladder cancer?What is the role of urinary cytology in the diagnosis of bladder cancer?Which cytology findings are diagnostic of bladder cancer?What is the role of cystoscopy in the diagnosis of bladder cancer?What is the role of imaging studies in the diagnosis of bladder cancer?What is the role of molecular and genetic markers in the diagnosis of bladder cancer?What is the role of blood tests in the diagnosis of bladder cancer?What is the role of CBC count in the evaluation of bladder cancer?What is the role of a chemistry panel in the diagnosis and management of bladder cancer?What is the role of renal function testing in the management of bladder cancer?What is the significance of hematuria in detecting patients with bladder cancer?What is the AUA recommended definition of microscopic hematuria?What is the role of voided urine cytology in the diagnosis of bladder cancer?What is the sensitivity and specificity of cytology in detecting bladder cancer?Which factors affect the sensitivity of cytology in detecting bladder cancer?What is the role of urine cytology in the diagnosis of carcinoma in situ (CIS) bladder cancer?What effect does instrumentation have on urine cytology findings in the evaluation of bladder cancer?How can the accuracy of urine cytology for detecting bladder cancer be improved?What role does noninvasive urine markers play in the diagnosis of bladder cancer?What are the guidelines for use of urinary tumor markers in the diagnosis of bladder cancer?What is cystoscopy and how is it used for detecting bladder cancer?What are the recommended actions following a negative cystoscopy in the evaluation of bladder cancer?What is the global prevalence of each types of bladder cancer?Which histologic finding is characteristic of carcinoma in situ (CIS) bladder cancer?How is cellular atypia distinguished from carcinoma in situ (CIS) bladder cancer?What is the role of CT scanning in the diagnosis of bladder cancer?What is intravenous pyelography (IVP) and how is it used in detecting bladder cancer?What is the role of ultrasonography in the diagnosis of bladder cancer?How is bladder cancer staged?What is the tumor, node, and metastases (TNM) staging system for bladder cancer?What is the prevalence of each stage of bladder cancer at diagnosis?What is the likelihood of bladder cancer metastasis?What is the significance of stage and grade in bladder cancer?How is carcinoma in situ (CIS) bladder cancer defined and graded?How is carcinoma in situ (CIS) bladder cancer designated in?What are the treatment options for non-muscle-invasive and muscle-invasive bladder cancer?What are the principal treatments for muscle-invasive bladder cancer?What is the role of radical cystectomy in the treatment of bladder cancer?What is the role of trimodality therapy for bladder cancer?What is the efficacy of trimodality therapy for bladder cancer?How might biomarkers affect treatment selection for bladder cancer in the future?What are treatment options for low-grade, low-stage bladder cancer?When are chemotherapies and immunotherapies indicated in the treatment of bladder cancer?What is the role of transurethral resection of bladder tumor (TURBT) in the treatment of bladder cancer?What are the EAU treatment guidelines for bladder cancer?What is the role of radiation therapy in the treatment of bladder cancer?What is the criterion standard for the treatment of patients with stage T2-T4 bladder cancer?When is lymph node dissection indicated in the treatment of bladder cancer?What are the treatment options for small cell carcinoma bladder cancer?What is the treatment for adenocarcinoma bladder cancer?What is the treatment for squamous cell carcinoma (SCC) bladder cancer?Which lifestyle changes decrease the risk of bladder cancer recurrence and progression?What is the treatment for high-risk Ta, T1, and carcinoma in situ (CIS) bladder cancer?How is intravesical instillation of bacillus Calmette-Guérin (BCG) vaccine or chemotherapy administered in the treatment of bladder cancer?What is the role of interferon alpha or gamma in the treatment of bladder cancer?What is the role of bacillus Calmette-Guérin (BCG) plus interferon in the treatment of bladder cancer?What is the role of fluorescence in situ hybridization (FISH) assays in the management of bladder cancer?What is the role of valrubicin (Valstar) in the treatment of bladder cancer?What is the role of docetaxel in the treatment of bladder cancer?What are the guidelines for surgical treatment of bladder cancer?What are the EAU recommendations for surgical treatment of bladder cancer?What is the role of radical cystectomy in the treatment of bladder cancer?What is the role of cystectomy in the treatment of non-muscle-invasive bladder cancer?What is the criterion standard for the treatment of stage T2-T4 bladder cancer?What is the role of radical cystoprostatectomy in the treatment of bladder cancer?What are the benefits of robot-assisted radical cystectomy in the treatment of bladder cancer?What is the impact on survival of a longer interval between diagnosis and radical cystectomy for bladder cancer?When is bilateral pelvic lymphadenectomy (PLND) indicated in the treatment of bladder cancer?What are the benefits of extended bilateral pelvic lymphadenectomy (PLND) in the treatment of bladder cancer?What are the contraindications for continent urinary diversion following cystectomy for bladder cancer?What is the most common incontinent urinary diversion procedure in the treatment of bladder cancer?How is an ileal conduit created following cystectomy for bladder cancer?What is the most common continent urinary diversion procedure in the treatment of bladder cancer?How is an orthotopic neobladder created following cystectomy for bladder cancer?How do continent urinary reservoirs for the treatment of bladder cancer vary?What is the role of cisplatin-based neoadjuvant chemotherapy in the treatment of bladder cancer?What is the evidence of efficacy for neoadjuvant chemotherapy in the treatment of bladder cancer?When is adjuvant chemotherapy indicated for treatment of bladder cancer?What impact does combination therapy have on the mortality risk of bladder cancer?What is the role of epithelial growth factor receptors (EGFRs) in the treatment of bladder cancer?What are the benefits to adjuvant chemotherapy in the treatment of bladder cancer?Which combination regimens are used in the treatment of metastatic bladder cancer?What is the role of programmed cell death inhibitors in the treatment of metastatic bladder cancer?What is the efficacy of second line therapy for metastatic bladder cancer?What is the role of atezolizumab (Tecentriq) in the treatment of metastatic bladder cancer?What is the role of nivolumab in the treatment of metastatic bladder cancer?What is the role of durvalumab (Imfinzi) in the treatment of metastatic bladder cancer?What is the role of avelumab (Bavencio) in the treatment of metastatic bladder cancer?What is the role of pembrolizumab (Keytruda) in the treatment of metastatic bladder cancer?What is the role of erdafitinib in the treatment of metastatic bladder cancer?What is the role of enfortumab vedotin in the treatment of metastatic breast cancer?What is the role of external beam radiation therapy (EBRT) in the treatment of bladder cancer?What is the efficacy of combination EBRT, chemotherapy and surgery in the treatment of bladder cancer?What is recurrence rate of bladder cancer following radiation therapy?What are the most common complications to radical cystectomy for bladder cancer?What is the overall complication rate of radical cystectomy for bladder cancer?What are complications in men of radical cystectomy for bladder cancer?What are complications of urinary diversion following cystectomy for bladder cancer?What are complications of orthotopic neobladder following cystectomy for bladder cancer?Which patients require follow-up following treatment for non-muscle invasive bladder cancer?What are the EAU guidelines for follow-up of patients with bladder cancer?What are the NCCN recommendations for follow-up after radical cystectomy for bladder cancer?Which organizations have issued guidelines for the diagnosis and management of bladder cancer?What is the evidence basis for the AUA-SUO bladder cancer guidelines?What is the evidence basis for the EAU bladder cancer guidelines?What is the evidence basis for the ESMO bladder cancer guidelines?What is the evidence basis for NCCN bladder cancer guidelines?What is the USPSTF recommendation for bladder cancer screening?What is the most established risk factor for bladder cancer?How much does quitting smoking reduce the risk for bladder cancer?What is the second most important risk factor for bladder cancer?Which tests are performed in the diagnosis of bladder cancer?What are the guidelines recommendations for use of urinary cytology in the diagnosis of bladder cancer?What are the AUA-SUO guidelines for use of cystoscopy in the diagnosis of bladder cancer?What are the EAU guidelines for use of cystoscopy in the diagnosis of bladder cancer?What are the NCCN guidelines for use of cystoscopy in the diagnosis of bladder cancer?What is the role of upper tract imaging in the diagnosis of bladder cancer?What are the NCCN diagnostic guidelines for noninvasive bladder cancer?What is the NCCN recommended preferred approach for imaging of the upper tract in the diagnosis of bladder cancer?What are the NCCN diagnostic guidelines for muscle-invasive bladder cancer?What are EAU guidelines for the staging of muscle-invasive bladder cancer?According to guidelines, what is the basis for a definitive diagnosis of bladder cancer?How is bladder cancer staged?What are characteristics of the different grades and stages of bladder cancer?What are the levels of risk stratification used by EAU for non-muscle-invasive bladder tumors?What are the NCCN recommendations for the treatment of low-grade Ta bladder tumors?What are the NCCN and EAU recommendations for the treatment of high-grade Ta bladder tumors?What are the NCCN recommendations for the treatment of T1 (low- and high-grade) bladder tumors?What are the EAU recommendations for the adjuvant treatment of Ta and T1 bladder tumors?What are the AUA/SUO recommendations for repeat TURBT in the treatment of bladder cancer?What are the recommendations AUA/SUO recommendations for intravesical therapy for bladder cancer?What are the NCCN recommendations for treatment of CIS bladder cancer?What are the EAU recommendations for cystoscopic follow-up of Ta, T1 and CIS bladder cancer after TURBT?What are the NCCN guidelines for cystoscopy and urinary cytology monitoring following treatment of bladder cancer?What are the AUA/SUO guidelines for monitoring following treatment of bladder cancer?What are the NCCN recommendations for the treatment of muscle-invasive bladder disease?What are the NCCN recommended first-line chemotherapy regimens advanced or metastatic bladder disease?What are the NCCN second-line therapy recommendations for the treatment of bladder cancer?What does the EAU recommend as the curative treatment of choice for muscle-invasive bladder cancer?What are the EAU recommendations for the treatment of muscle-invasive bladder cancer?What are EAU recommendations for first-line chemotherapy for advanced or metastatic bladder cancer?What are EAU recommended second-line treatments for advanced or metastatic bladder cancer?What are the ESMO guidelines for the treatment of metastatic bladder cancer?What are the NCCN recommendations for treatment of locally advanced/metastatic bladder cancer in platinum-ineligible with no PD-L1 expression patients??What are the NCCN and EAU recommendations for the diagnosis of upper tract urothelial carcinoma (UTUC) bladder cancer?What are the NCCN recommendations for the treatment of UTUC bladder cancer?What are the EAU recommendations for management of UTUC bladder cancer?What are the EAU recommendations for radical nephroureterectomy (RNU) in the treatment of UTUC bladder cancer?What are the indications for conservative management of UTUC bladder cancer?What are EAU recommendations for conservative treatment for UTUC bladder cancer?What are the EAU recommendations for surveillance after treatment for UTUC bladder cancer?Which mediations are the standard of treatment for metastatic bladder cancer?What is the benefit of new combination regimens for the treatment of bladder cancer?Which medications in the drug class PD-1/PD-L1 Inhibitors are used in the treatment of Bladder Cancer?Which medications in the drug class Antineoplastics, Antimicrotubular are used in the treatment of Bladder Cancer?Which medications in the drug class Antineoplastics, Alkylating are used in the treatment of Bladder Cancer?Which medications in the drug class Antineoplastics, Anthracycline are used in the treatment of Bladder Cancer?Which medications in the drug class Antineoplastics, Vinca Alkaloid are used in the treatment of Bladder Cancer?Which medications in the drug class Antineoplastics, Antimetabolite are used in the treatment of Bladder Cancer?Which medications in the drug class FGFR Inhibitors are used in the treatment of Bladder Cancer?Which medications in the drug class Anti-Nectin-4 Monoclonal Antibodies are used in the treatment of Bladder Cancer?

Author

Gary David Steinberg, MD, FACS, Professor and Director, Goldstein Bladder Cancer Program, NYU Langone Health

Disclosure: Received consulting fee from Abbott Molecular for consulting; Received consulting fee from Endo Pharmaceuticals for consulting; Received consulting fee from Bioniche for consulting; Received consulting fee from Tengion for consulting; Received consulting fee from Archimedes for review panel membership; Received consulting fee from PhotoCure for review panel membership; Received consulting fee from Taris Biomedical for review panel membership; Received none from Cold Genesys for other; Received h for: Photocure; Taris Biomedical; Heat Biologics: Cold Genesys; Merck; Roche/Genentech; Karl Storz; Mdx Health, Telesta.

Coauthor(s)

Bagi RP Jana, MD, Professor of Medicine (Genitourinary Oncology), Division of Hematology and Oncology, University of Texas Medical Branch at Galveston

Disclosure: Nothing to disclose.

Kush Sachdeva, MD, Southern Oncology and Hematology Associates, Inspira Health Network

Disclosure: Nothing to disclose.

Chief Editor

Bradley Fields Schwartz, DO, FACS, Professor of Urology, Director, Center for Laparoscopy and Endourology, Department of Surgery, Southern Illinois University School of Medicine

Disclosure: Serve(d) as a director, officer, partner, employee, advisor, consultant or trustee for: AUA Journal of Urology<br/>Serve(d) as a speaker or a member of a speakers bureau for: Cook Medical; Olympus, .

Acknowledgements

Sujeet S Acharya, MD Resident Physician, Department of Surgery, Section of Urology, University of Chicago Division of the Biological Sciences, The Pritzker School of Medicine

Disclosure: Nothing to disclose.

Brendan Curti, MD Director, Genitourinary Oncology Research, Robert W Franz Cancer Research Center, Earle A Chiles Research Institute, Providence Cancer Center

Brendan Curti, MD is a member of the following medical societies: American College of Physicians, American Society of Clinical Oncology, Oregon Medical Association, and Society for Biological Therapy

Disclosure: Nothing to disclose.

Edward M Gong, MD Fellow, Department of Surgery, Division of Urology, Children's Hospital Boston

Disclosure: Nothing to disclose.

Mark H Katz, MD Fellow in Urologic Oncology and Minimally Invasive Surgery, University of Chicago Medical Center

Mark H Katz, MD is a member of the following medical societies: Alpha Omega Alpha, American Urological Association, Endourological Society, and Society of Urologic Oncology

Disclosure: Nothing to disclose.

Hyung L Kim, MD Associate Professor, Cedars-Sinai Medical Center

Disclosure: Nothing to disclose.

Francisco Talavera, PharmD, PhD Adjunct Assistant Professor, University of Nebraska Medical Center College of Pharmacy; Editor-in-Chief, Medscape Drug Reference

Disclosure: Medscape Salary Employment

Dan Theodorescu, MD, PhD Paul A Bunn Professor of Cancer Research, Professor of Surgery and Pharmacology, Director, University of Colorado Comprehensive Cancer Center

Dan Theodorescu, MD, PhD is a member of the following medical societies: American Cancer Society, American College of Surgeons, American Urological Association, Medical Society of Virginia, Society for Basic Urologic Research, and Society of Urologic Oncology

Disclosure: Key Genomics Ownership interest Co-Founder-50% Stock Ownership; KromaTiD, Inc Stock Options Board membership

References

  1. [Guideline] National Comprehensive Cancer Network. NCCN Clinical Practice Guidelines in Oncology. Bladder Cancer Version 3.2019. NCCN.org. Available at http://www.nccn.org/professionals/physician_gls/pdf/bladder.pdf. April 29, 2019; Accessed: June 22, 2019.
  2. Escudero DO, Shirodkar SP, Lokeshwar VB. Bladder Carcinogenesis and Molecular Pathways. Lokeshwar VB. Bladder Tumors: Molecular Aspects and Clinical Management. New York: Springer Science; 2010. 23-41.
  3. Spruck CH 3rd, Ohneseit PF, Gonzalez-Zulueta M, Esrig D, Miyao N, Tsai YC, et al. Two molecular pathways to transitional cell carcinoma of the bladder. Cancer Res. 1994 Feb 1. 54(3):784-8. [View Abstract]
  4. Trias I, Algaba F, Condom E, Español I, Seguí J, Orsola I, et al. Small cell carcinoma of the urinary bladder. Presentation of 23 cases and review of 134 published cases. Eur Urol. 2001 Jan. 39(1):85-90. [View Abstract]
  5. Bessette PL, Abell MR, Herwig KR. A clinicopathologic study of squamous cell carcinoma of the bladder. J Urol. 1974 Jul. 112(1):66-7. [View Abstract]
  6. Faysal MH. Squamous cell carcinoma of the bladder. J Urol. 1981 Nov. 126(5):598-9. [View Abstract]
  7. Lagwinski N, Thomas A, Stephenson AJ, Campbell S, Hoschar AP, El-Gabry E, et al. Squamous cell carcinoma of the bladder: a clinicopathologic analysis of 45 cases. Am J Surg Pathol. 2007 Dec. 31(12):1777-87. [View Abstract]
  8. El-Sebaie M, Zaghloul MS, Howard G, Mokhtar A. Squamous cell carcinoma of the bilharzial and non-bilharzial urinary bladder: a review of etiological features, natural history, and management. Int J Clin Oncol. 2005 Feb. 10(1):20-5. [View Abstract]
  9. Heyns CF, van der Merwe A. Bladder cancer in Africa. Can J Urol. 2008 Feb. 15(1):3899-908. [View Abstract]
  10. Tomlinson DC, Baldo O, Harnden P, Knowles MA. FGFR3 protein expression and its relationship to mutation status and prognostic variables in bladder cancer. J Pathol. 2007 Sep. 213(1):91-8. [View Abstract]
  11. Eswarakumar VP, Lax I, Schlessinger J. Cellular signaling by fibroblast growth factor receptors. Cytokine Growth Factor Rev. 2005 Apr. 16(2):139-49. [View Abstract]
  12. Fadl-Elmula I. Chromosomal changes in uroepithelial carcinomas. Cell Chromosome. 2005 Aug 7. 4:1. [View Abstract]
  13. Knowles MA. Molecular subtypes of bladder cancer: Jekyll and Hyde or chalk and cheese?. Carcinogenesis. 2006 Mar. 27(3):361-73. [View Abstract]
  14. Salinas-Sánchez AS, Lorenzo-Romero JG, Giménez-Bachs JM, Sánchez-Sánchez F, Donate-Moreno MJ, Rubio-Del-Campo A, et al. Implications of p53 gene mutations on patient survival in transitional cell carcinoma of the bladder: a long-term study. Urol Oncol. 2008 Nov-Dec. 26(6):620-6. [View Abstract]
  15. Miyamoto H, Shuin T, Ikeda I, Hosaka M, Kubota Y. Loss of heterozygosity at the p53, RB, DCC and APC tumor suppressor gene loci in human bladder cancer. J Urol. 1996 Apr. 155(4):1444-7. [View Abstract]
  16. Karam JA, Lotan Y, Karakiewicz PI, Ashfaq R, Sagalowsky AI, Roehrborn CG, et al. Use of combined apoptosis biomarkers for prediction of bladder cancer recurrence and mortality after radical cystectomy. Lancet Oncol. 2007 Feb. 8(2):128-36. [View Abstract]
  17. Campbell SC, Volpert OV, Ivanovich M, Bouck NP. Molecular mediators of angiogenesis in bladder cancer. Cancer Res. 1998 Mar 15. 58(6):1298-304. [View Abstract]
  18. Chan KS, Espinosa I, Chao M, Wong D, Ailles L, Diehn M, et al. Identification, molecular characterization, clinical prognosis, and therapeutic targeting of human bladder tumor-initiating cells. Proc Natl Acad Sci U S A. 2009 Aug 18. 106(33):14016-21. [View Abstract]
  19. Sanchez-Carbayo M, Socci ND, Lozano J, Saint F, Cordon-Cardo C. Defining molecular profiles of poor outcome in patients with invasive bladder cancer using oligonucleotide microarrays. J Clin Oncol. 2006 Feb 10. 24(5):778-89. [View Abstract]
  20. Brennan P, Bogillot O, Cordier S, Greiser E, Schill W, Vineis P, et al. Cigarette smoking and bladder cancer in men: a pooled analysis of 11 case-control studies. Int J Cancer. 2000 Apr 15. 86(2):289-94. [View Abstract]
  21. Fortuny J, Kogevinas M, Chang-Claude J, González CA, Hours M, Jöckel KH, et al. Tobacco, occupation and non-transitional-cell carcinoma of the bladder: an international case-control study. Int J Cancer. 1999 Jan 5. 80(1):44-6. [View Abstract]
  22. Freedman ND, Silverman DT, Hollenbeck AR, Schatzkin A, Abnet CC. Association between smoking and risk of bladder cancer among men and women. JAMA. 2011 Aug 17. 306(7):737-45. [View Abstract]
  23. Cumberbatch MG, Cox A, Teare D, Catto JW. Contemporary Occupational Carcinogen Exposure and Bladder Cancer: A Systematic Review and Meta-analysis. JAMA Oncol. 2015 Dec. 1 (9):1282-90. [View Abstract]
  24. Baris D, Waddell R, Beane Freeman LE, Schwenn M, Colt JS, et al. Elevated Bladder Cancer in Northern New England: The Role of Drinking Water and Arsenic. J Natl Cancer Inst. 2016 Sep. 108 (9):[View Abstract]
  25. Nelson R. Arsenic-Contaminated Well Water Boosts Bladder Cancer Risk. Medscape Medical News. Available at http://www.medscape.com/viewarticle/862914. May 5, 2016; Accessed: May 7, 2016.
  26. Stein JP, Skinner EC, Boyd SD, Skinner DG. Squamous cell carcinoma of the bladder associated with cyclophosphamide therapy for Wegener's granulomatosis: a report of 2 cases. J Urol. 1993 Mar. 149(3):588-9. [View Abstract]
  27. Figueroa JD, Koutros S, Colt JS, Kogevinas M, Garcia-Closas M, et al. Modification of Occupational Exposures on Bladder Cancer Risk by Common Genetic Polymorphisms. J Natl Cancer Inst. 2015 Nov. 107 (11):[View Abstract]
  28. El-Bolkainy MN, Mokhtar NM, Ghoneim MA, Hussein MH. The impact of schistosomiasis on the pathology of bladder carcinoma. Cancer. 1981 Dec 15. 48(12):2643-8. [View Abstract]
  29. Botelho M, Ferreira AC, Oliveira MJ, Domingues A, Machado JC, da Costa JM. Schistosoma haematobium total antigen induces increased proliferation, migration and invasion, and decreases apoptosis of normal epithelial cells. Int J Parasitol. 2009 Aug. 39(10):1083-91. [View Abstract]
  30. Ahmad I, Barnetson RJ, Krishna NS. Keratinizing squamous metaplasia of the bladder: a review. Urol Int. 2008. 81(3):247-51. [View Abstract]
  31. Khan MS, Thornhill JA, Gaffney E, Loftus B, Butler MR. Keratinising squamous metaplasia of the bladder: natural history and rationalization of management based on review of 54 years experience. Eur Urol. 2002 Nov. 42(5):469-74. [View Abstract]
  32. Newman DM, Brown JR, Jay AC, Pontius EE. Squamous cell carcinoma of the bladder. J Urol. 1968 Oct. 100(4):470-3. [View Abstract]
  33. Faysal MH, Freiha FS. Primary neoplasm in vesical diverticula. A report of 12 cases. Br J Urol. 1981 Apr. 53(2):141-3. [View Abstract]
  34. Yurdakul T, Avunduk MC, Piskin MM. Pure squamous cell carcinoma after intravesical BCG treatment. A case report. Urol Int. 2005. 74(3):283-5. [View Abstract]
  35. STUART WT. Carcinoma of the bladder associated with exstrophy. Report of a case and review of the literature. Va Med Mon (1918). 1962 Jan. 89:39-42. [View Abstract]
  36. Ribeiro JC, Silva C, Sousa L, García P, Santos A. [Squamous cell carcinoma in bladder extrophy]. Actas Urol Esp. 2005 Jan. 29(1):110-2. [View Abstract]
  37. Gupta S, Gupta IM. Ectopia vesicae complicated by squamous cell carcinoma. Br J Urol. 1976 Aug. 48(4):244. [View Abstract]
  38. Rieder JM, Parsons JK, Gearhart JP, Schoenberg M. Primary squamous cell carcinoma in unreconstructed exstrophic bladder. Urology. 2006 Jan. 67(1):199. [View Abstract]
  39. Sheldon CA, Clayman RV, Gonzalez R, Williams RD, Fraley EE. Malignant urachal lesions. J Urol. 1984 Jan. 131(1):1-8. [View Abstract]
  40. Lin RY, Rappoport AE, Deppisch LM, Natividad NS, Katz W. Squamous cell carcinoma of the urachus. J Urol. 1977 Dec. 118(6):1066-7. [View Abstract]
  41. SHAW RE. Squamous-cell carcinoma in a cyst of the urachus. Br J Urol. 1958 Mar. 30(1):87-9. [View Abstract]
  42. Chow YC, Lin WC, Tzen CY, Chow YK, Lo KY. Squamous cell carcinoma of the urachus. J Urol. 2000 Mar. 163(3):903-4. [View Abstract]
  43. Fujiyama C, Nakashima N, Tokuda Y, Uozumi J. Squamous cell carcinoma of the urachus. Int J Urol. 2007 Oct. 14(10):966-8. [View Abstract]
  44. Cancer Facts & Figures 2019. American Cancer Society. Available at https://www.cancer.org/content/dam/cancer-org/research/cancer-facts-and-statistics/annual-cancer-facts-and-figures/2019/cancer-facts-and-figures-2019.pdf. Accessed: January 31, 2019.
  45. Cancer Stat Facts: Bladder Cancer. National Cancer Institute. Available at https://seer.cancer.gov/statfacts/html/urinb.html. Accessed: January 31, 2019.
  46. Siegel RL, Miller KD, Jemal A. Cancer statistics, 2019. CA Cancer J Clin. 2019 Jan. 69 (1):7-34. [View Abstract]
  47. Dawson C, Whitfield H. ABC of Urology. Urological malignancy--II: Urothelial tumours. BMJ. 1996 Apr 27. 312(7038):1090-4. [View Abstract]
  48. Abrahams NA, Moran C, Reyes AO, Siefker-Radtke A, Ayala AG. Small cell carcinoma of the bladder: a contemporary clinicopathological study of 51 cases. Histopathology. 2005 Jan. 46(1):57-63. [View Abstract]
  49. Lohrisch C, Murray N, Pickles T, Sullivan L. Small cell carcinoma of the bladder: long term outcome with integrated chemoradiation. Cancer. 1999 Dec 1. 86(11):2346-52. [View Abstract]
  50. Gouda I, Mokhtar N, Bilal D, El-Bolkainy T, El-Bolkainy NM. Bilharziasis and bladder cancer: a time trend analysis of 9843 patients. J Egypt Natl Canc Inst. 2007 Jun. 19(2):158-62. [View Abstract]
  51. Felix AS, Soliman AS, Khaled H, Zaghloul MS, Banerjee M, El-Baradie M, et al. The changing patterns of bladder cancer in Egypt over the past 26 years. Cancer Causes Control. 2008 May. 19(4):421-9. [View Abstract]
  52. Elsobky E, El-Baz M, Gomha M, Abol-Enein H, Shaaban AA. Prognostic value of angiogenesis in schistosoma-associated squamous cell carcinoma of the urinary bladder. Urology. 2002 Jul. 60(1):69-73. [View Abstract]
  53. Griffiths TR, Charlton M, Neal DE, Powell PH. Treatment of carcinoma in situ with intravesical bacillus Calmette-Guerin without maintenance. J Urol. 2002 Jun. 167(6):2408-12. [View Abstract]
  54. Pycha A, Mian C, Posch B, Haitel A, Mokhtar AA, El-Baz M, et al. Numerical chromosomal aberrations in muscle invasive squamous cell and transitional cell cancer of the urinary bladder: an alternative to classic prognostic indicators?. Urology. 1999 May. 53(5):1005-10. [View Abstract]
  55. Shaaban AA, Javadpour N, Tribukait B, Ghoneim MA. Prognostic significance of flow-DNA analysis and cell surface isoantigens in carcinoma of bilharzial bladder. Urology. 1992 Mar. 39(3):207-10. [View Abstract]
  56. Ghoneim MA, Ashamallah AK, Awaad HK, Whitmore WF Jr. Randomized trial of cystectomy with or without preoperative radiotherapy for carcinoma of the bilharzial bladder. J Urol. 1985 Aug. 134(2):266-8. [View Abstract]
  57. Cheng L, Pan CX, Yang XJ, Lopez-Beltran A, MacLennan GT, Lin H, et al. Small cell carcinoma of the urinary bladder: a clinicopathologic analysis of 64 patients. Cancer. 2004 Sep 1. 101(5):957-62. [View Abstract]
  58. Shahab N. Extrapulmonary small cell carcinoma of the bladder. Semin Oncol. 2007 Feb. 34(1):15-21. [View Abstract]
  59. Mackey JR, Au HJ, Hugh J, Venner P. Genitourinary small cell carcinoma: determination of clinical and therapeutic factors associated with survival. J Urol. 1998 May. 159(5):1624-9. [View Abstract]
  60. Choong NW, Quevedo JF, Kaur JS. Small cell carcinoma of the urinary bladder. The Mayo Clinic experience. Cancer. 2005 Mar 15. 103(6):1172-8. [View Abstract]
  61. van Rhijn BW, Burger M, Lotan Y, Solsona E, Stief CG, Sylvester RJ, et al. Recurrence and progression of disease in non-muscle-invasive bladder cancer: from epidemiology to treatment strategy. Eur Urol. 2009 Sep. 56(3):430-42. [View Abstract]
  62. Fernandez-Gomez J, Solsona E, Unda M, Martinez-Piñeiro L, Gonzalez M, Hernandez R, et al. Prognostic factors in patients with non-muscle-invasive bladder cancer treated with bacillus Calmette-Guérin: multivariate analysis of data from four randomized CUETO trials. Eur Urol. 2008 May. 53(5):992-1001. [View Abstract]
  63. Cha EK, Tirsar LA, Schwentner C, Christos PJ, Mian C, Hennenlotter J, et al. Immunocytology is a strong predictor of bladder cancer presence in patients with painless hematuria: a multicentre study. Eur Urol. 2012 Jan. 61(1):185-92. [View Abstract]
  64. Strittmatter F, Buchner A, Karl A, Sommer ML, Straub J, Tilki D, et al. Individual learning curve reduces the clinical value of urinary cytology. Clin Genitourin Cancer. 2011 Sep. 9(1):22-6. [View Abstract]
  65. Lotan Y, Roehrborn CG. Cost-effectiveness of a modified care protocol substituting bladder tumor markers for cystoscopy for the followup of patients with transitional cell carcinoma of the bladder: a decision analytical approach. J Urol. 2002 Jan. 167(1):75-9. [View Abstract]
  66. Apolo AB, Vogelzang NJ, Theodorescu D. New and promising strategies in the management of bladder cancer. Am Soc Clin Oncol Educ Book. 2015. 35:105-12. [View Abstract]
  67. Grossfeld GD, Litwin MS, Wolf JS Jr, Hricak H, Shuler CL, Agerter DC, et al. Evaluation of asymptomatic microscopic hematuria in adults: the American Urological Association best practice policy--part II: patient evaluation, cytology, voided markers, imaging, cystoscopy, nephrology evaluation, and follow-up. Urology. 2001 Apr. 57(4):604-10. [View Abstract]
  68. Murphy WM, Crabtree WN, Jukkola AF, Soloway MS. The diagnostic value of urine versus bladder washing in patients with bladder cancer. J Urol. 1981 Sep. 126(3):320-2. [View Abstract]
  69. Lokeshwar VB, Soloway MS. Current bladder tumor tests: does their projected utility fulfill clinical necessity?. J Urol. 2001 Apr. 165(4):1067-77. [View Abstract]
  70. Grossman HB, Soloway M, Messing E, Katz G, Stein B, Kassabian V, et al. Surveillance for recurrent bladder cancer using a point-of-care proteomic assay. JAMA. 2006 Jan 18. 295(3):299-305. [View Abstract]
  71. Al-Sukhun S, Hussain M. Molecular biology of transitional cell carcinoma. Crit Rev Oncol Hematol. 2003 Aug. 47(2):181-93. [View Abstract]
  72. Halling KC, Kipp BR. Bladder cancer detection using FISH (UroVysion assay). Adv Anat Pathol. 2008 Sep. 15(5):279-86. [View Abstract]
  73. Soloway MS, Briggman V, Carpinito GA, Chodak GW, Church PA, Lamm DL, et al. Use of a new tumor marker, urinary NMP22, in the detection of occult or rapidly recurring transitional cell carcinoma of the urinary tract following surgical treatment. J Urol. 1996 Aug. 156(2 Pt 1):363-7. [View Abstract]
  74. van Rhijn BW, van der Poel HG, van der Kwast TH. Urine markers for bladder cancer surveillance: a systematic review. Eur Urol. 2005 Jun. 47 (6):736-48. [View Abstract]
  75. [Guideline] Babjuk M, Burger M, Compérat EM, et al. EAU Guidelines on Non-Muscle-invasive Bladder Cancer--2019 update. UROWEB.org. Available at http://uroweb.org/guideline/non-muscle-invasive-bladder-cancer/. Accessed: June 22, 2019.
  76. Hall MC, Chang SS, Dalbagni G, Pruthi RS, Seigne JD, Skinner EC, et al. Guideline for the management of nonmuscle invasive bladder cancer (stages Ta, T1, and Tis): 2007 update. J Urol. 2007 Dec. 178(6):2314-30. [View Abstract]
  77. American Joint Committee on Cancer. Urinary bladder. Amin MB, Edge S, Greene F, Byrd DR, Brookland RK, et al, eds. AJCC Cancer Staging Manual. 8th ed. New York: Springer; 2017.
  78. [Guideline] Witjes JA, Bruins M, Cathomas R, et al. EAU Guidelines on Muscle-invasive and metastatic bladder cancer--2019 update. Uroweb.org. Available at https://uroweb.org/guideline/bladder-cancer-muscle-invasive-and-metastatic/. Accessed: June 22, 2019.
  79. Chedgy EC, Black PC. Radical Cystectomy and the Multidisciplinary Management of Muscle-Invasive Bladder Cancer. JAMA Oncol. 2016 May 5. [View Abstract]
  80. Mitin T. Rethinking Radical Cystectomy as the Best Choice for Most Patients With Muscle-Invasive Bladder Cancer. JAMA Oncol. 2016 May 5. [View Abstract]
  81. Teo MT, Dyrskjøt L, Nsengimana J, Buchwald C, Snowden H, Morgan J, et al. Next-generation sequencing identifies germline MRE11A variants as markers of radiotherapy outcomes in muscle-invasive bladder cancer. Ann Oncol. 2014 Apr. 25 (4):877-83. [View Abstract]
  82. Serretta V, Galuffo A, Pavone C, Allegro R, Pavone-MacAluso M. Gemcitabine in intravesical treatment of Ta-T1 transitional cell carcinoma of bladder: Phase I-II study on marker lesions. Urology. 2005 Jan. 65(1):65-9. [View Abstract]
  83. Sylvester RJ, van der Meijden AP, Witjes JA, Kurth K. Bacillus calmette-guerin versus chemotherapy for the intravesical treatment of patients with carcinoma in situ of the bladder: a meta-analysis of the published results of randomized clinical trials. J Urol. 2005 Jul. 174(1):86-91; discussion 91-2. [View Abstract]
  84. Witjes JA, Hendricksen K. Intravesical pharmacotherapy for non-muscle-invasive bladder cancer: a critical analysis of currently available drugs, treatment schedules, and long-term results. Eur Urol. 2008 Jan. 53(1):45-52. [View Abstract]
  85. Zaharoff DA, Hoffman BS, Hooper HB, Benjamin CJ Jr, Khurana KK, Hance KW, et al. Intravesical immunotherapy of superficial bladder cancer with chitosan/interleukin-12. Cancer Res. 2009 Aug 1. 69(15):6192-9. [View Abstract]
  86. Islam MA, Bhuiyan ZH, Shameem IA. Intravesical adjuvant therapy using mitomycin C. Mymensingh Med J. 2006 Jan. 15(1):40-4. [View Abstract]
  87. Herr HW, Dalbagni G, Donat SM. Bacillus Calmette-Guérin without maintenance therapy for high-risk non-muscle-invasive bladder cancer. Eur Urol. 2011 Jul. 60(1):32-6. [View Abstract]
  88. Schmidbauer J, Witjes F, Schmeller N, Donat R, Susani M, Marberger M. Improved detection of urothelial carcinoma in situ with hexaminolevulinate fluorescence cystoscopy. J Urol. 2004 Jan. 171(1):135-8. [View Abstract]
  89. Jichlinski P, Guillou L, Karlsen SJ, Malmström PU, Jocham D, Brennhovd B, et al. Hexyl aminolevulinate fluorescence cystoscopy: new diagnostic tool for photodiagnosis of superficial bladder cancer--a multicenter study. J Urol. 2003 Jul. 170(1):226-9. [View Abstract]
  90. Hungerhuber E, Stepp H, Kriegmair M, Stief C, Hofstetter A, Hartmann A, et al. Seven years' experience with 5-aminolevulinic acid in detection of transitional cell carcinoma of the bladder. Urology. 2007 Feb. 69(2):260-4. [View Abstract]
  91. Kausch I, Sommerauer M, Montorsi F, Stenzl A, Jacqmin D, Jichlinski P, et al. Photodynamic diagnosis in non-muscle-invasive bladder cancer: a systematic review and cumulative analysis of prospective studies. Eur Urol. 2010 Apr. 57(4):595-606. [View Abstract]
  92. Waknine Y. FDA Approves Cysview for Cystoscopic Detection of Papillary Bladder Cancer. Medscape Medical News, June 4, 2010. Available at http://www.medscape.com/viewarticle/722923. Accessed: January 10, 2013.
  93. Booth CM, Tannock IF. Benefits of Adjuvant Chemotherapy for Bladder Cancer. JAMA Oncol. 2015 Sep. 1 (6):727-8. [View Abstract]
  94. Standard or Extended Pelvic Lymphadenectomy in Treating Patients Undergoing Surgery for Invasive Bladder Cancer. ClinicalTrials.gov. ClinicalTrials.gov. Available at http://clinicaltrials.gov/ct2/show/NCT01224665. Accessed: August 17, 2018.
  95. Mukesh M, Cook N, Hollingdale AE, Ainsworth NL, Russell SG. Small cell carcinoma of the urinary bladder: a 15-year retrospective review of treatment and survival in the Anglian Cancer Network. BJU Int. 2009 Mar. 103(6):747-52. [View Abstract]
  96. Ehdaie B, Maschino A, Shariat SF, Rioja J, Hamilton RJ, Lowrance WT, et al. Comparative outcomes of pure squamous cell carcinoma and urothelial carcinoma with squamous differentiation in patients treated with radical cystectomy. J Urol. 2012 Jan. 187(1):74-9. [View Abstract]
  97. Brinkman MT, Karagas MR, Zens MS, Schned A, Reulen RC, Zeegers MP. Minerals and vitamins and the risk of bladder cancer: results from the New Hampshire Study. Cancer Causes Control. 2010 Apr. 21(4):609-19. [View Abstract]
  98. O'Donnell MA, Lilli K, Leopold C. Interim results from a national multicenter phase II trial of combination bacillus Calmette-Guerin plus interferon alfa-2b for superficial bladder cancer. J Urol. 2004 Sep. 172(3):888-93. [View Abstract]
  99. Nepple KG, Lightfoot AJ, Rosevear HM, O'Donnell MA, Lamm DL. Bacillus Calmette-Guérin with or without interferon a-2b and megadose versus recommended daily allowance vitamins during induction and maintenance intravesical treatment of nonmuscle invasive bladder cancer. J Urol. 2010 Nov. 184(5):1915-9. [View Abstract]
  100. Kamat AM, Dickstein RJ, Messetti F, Anderson R, Pretzsch SM, Gonzalez GN, et al. Use of fluorescence in situ hybridization to predict response to bacillus Calmette-Guérin therapy for bladder cancer: results of a prospective trial. J Urol. 2012 Mar. 187(3):862-7. [View Abstract]
  101. Barlow L, McKiernan JM, Benson MC. Long-term survival outcomes with intravesical docetaxel for recurrent nonmuscle invasive bladder cancer after previous bacillus Calmette-Guérin therapy. J Urol. 2013 Mar. 189(3):834-9. [View Abstract]
  102. Schmidbauer J, Witjes F, Schmeller N, Donat R, Susani M, Marberger M. Improved detection of urothelial carcinoma in situ with hexaminolevulinate fluorescence cystoscopy. J Urol. 2004 Jan. 171(1):135-8. [View Abstract]
  103. Jichlinski P, Guillou L, Karlsen SJ, Malmström PU, Jocham D, Brennhovd B, et al. Hexyl aminolevulinate fluorescence cystoscopy: new diagnostic tool for photodiagnosis of superficial bladder cancer--a multicenter study. J Urol. 2003 Jul. 170(1):226-9. [View Abstract]
  104. Hungerhuber E, Stepp H, Kriegmair M, Stief C, Hofstetter A, Hartmann A, et al. Seven years' experience with 5-aminolevulinic acid in detection of transitional cell carcinoma of the bladder. Urology. 2007 Feb. 69(2):260-4. [View Abstract]
  105. Fradet Y, Grossman HB, Gomella L, Lerner S, Cookson M, Albala D, et al. A comparison of hexaminolevulinate fluorescence cystoscopy and white light cystoscopy for the detection of carcinoma in situ in patients with bladder cancer: a phase III, multicenter study. J Urol. 2007 Jul. 178(1):68-73; discussion 73. [View Abstract]
  106. Jocham D, Witjes F, Wagner S, Zeylemaker B, van Moorselaar J, Grimm MO, et al. Improved detection and treatment of bladder cancer using hexaminolevulinate imaging: a prospective, phase III multicenter study. J Urol. 2005 Sep. 174(3):862-6; discussion 866. [View Abstract]
  107. Stenzl A, Burger M, Fradet Y, Mynderse LA, Soloway MS, Witjes JA, et al. Hexaminolevulinate guided fluorescence cystoscopy reduces recurrence in patients with nonmuscle invasive bladder cancer. J Urol. 2010 Nov. 184(5):1907-13. [View Abstract]
  108. Hermann GG, Mogensen K, Carlsson S, Marcussen N, Duun S. Fluorescence-guided transurethral resection of bladder tumours reduces bladder tumour recurrence due to less residual tumour tissue in Ta/T1 patients: a randomized two-centre study. BJU Int. 2011 Oct. 108(8 Pt 2):E297-303. [View Abstract]
  109. Tilki D, Reich O, Svatek RS, Karakiewicz PI, Kassouf W, Novara G, et al. Characteristics and outcomes of patients with clinical carcinoma in situ only treated with radical cystectomy: an international study of 243 patients. J Urol. 2010 May. 183(5):1757-63. [View Abstract]
  110. Davis JW, Castle EP, Pruthi RS, Ornstein DK, Guru KA. Robot-assisted radical cystectomy: an expert panel review of the current status and future direction. Urol Oncol. 2010 Sep-Oct. 28(5):480-6. [View Abstract]
  111. Chang SS, Cookson MS. Radical cystectomy for bladder cancer: the case for early intervention. Urol Clin North Am. 2005 May. 32(2):147-55. [View Abstract]
  112. Sánchez-Ortiz RF, Huang WC, Mick R, Van Arsdalen KN, Wein AJ, Malkowicz SB. An interval longer than 12 weeks between the diagnosis of muscle invasion and cystectomy is associated with worse outcome in bladder carcinoma. J Urol. 2003 Jan. 169(1):110-5; discussion 115. [View Abstract]
  113. Raghavan D, Burgess E, Gaston KE, Haake MR, Riggs SB. Neoadjuvant and adjuvant chemotherapy approaches for invasive bladder cancer. Semin Oncol. 2012 Oct. 39(5):588-97. [View Abstract]
  114. Winquist E, Kirchner TS, Segal R, Chin J, Lukka H. Neoadjuvant chemotherapy for transitional cell carcinoma of the bladder: a systematic review and meta-analysis. J Urol. 2004 Feb. 171(2 Pt 1):561-9. [View Abstract]
  115. Grossman HB, Natale RB, Tangen CM, Speights VO, Vogelzang NJ, Trump DL, et al. Neoadjuvant chemotherapy plus cystectomy compared with cystectomy alone for locally advanced bladder cancer. N Engl J Med. 2003 Aug 28. 349(9):859-66. [View Abstract]
  116. Herr HW, Faulkner JR, Grossman HB, Natale RB, deVere White R, Sarosdy MF, et al. Surgical factors influence bladder cancer outcomes: a cooperative group report. J Clin Oncol. 2004 Jul 15. 22(14):2781-9. [View Abstract]
  117. Griffiths G, Hall R, Sylvester R, Raghavan D, Parmar MK. International phase III trial assessing neoadjuvant cisplatin, methotrexate, and vinblastine chemotherapy for muscle-invasive bladder cancer: long-term results of the BA06 30894 trial. J Clin Oncol. 2011 Jun 1. 29(16):2171-7. [View Abstract]
  118. Mooso BA, Vinall RL, Mudryj M, Yap SA, deVere White RW, Ghosh PM. The role of EGFR family inhibitors in muscle invasive bladder cancer: a review of clinical data and molecular evidence. J Urol. 2015 Jan. 193 (1):19-29. [View Abstract]
  119. Sternberg CN, Skoneczna I, Kerst JM, Albers P, Fossa SD, et al. Immediate versus deferred chemotherapy after radical cystectomy in patients with pT3-pT4 or N+ M0 urothelial carcinoma of the bladder (EORTC 30994): an intergroup, open-label, randomised phase 3 trial. Lancet Oncol. 2015 Jan. 16 (1):76-86. [View Abstract]
  120. Saxman SB, Propert KJ, Einhorn LH, Crawford ED, Tannock I, Raghavan D, et al. Long-term follow-up of a phase III intergroup study of cisplatin alone or in combination with methotrexate, vinblastine, and doxorubicin in patients with metastatic urothelial carcinoma: a cooperative group study. J Clin Oncol. 1997 Jul. 15(7):2564-9. [View Abstract]
  121. Sternberg CN, de Mulder P, Schornagel JH, Theodore C, Fossa SD, van Oosterom AT, et al. Seven year update of an EORTC phase III trial of high-dose intensity M-VAC chemotherapy and G-CSF versus classic M-VAC in advanced urothelial tract tumours. Eur J Cancer. 2006 Jan. 42(1):50-4. [View Abstract]
  122. von der Maase H, Sengelov L, Roberts JT, Ricci S, Dogliotti L, Oliver T, et al. Long-term survival results of a randomized trial comparing gemcitabine plus cisplatin, with methotrexate, vinblastine, doxorubicin, plus cisplatin in patients with bladder cancer. J Clin Oncol. 2005 Jul 20. 23(21):4602-8. [View Abstract]
  123. von der Maase H, Sengelov L, Roberts JT, Ricci S, Dogliotti L, Oliver T, et al. Long-term survival results of a randomized trial comparing gemcitabine plus cisplatin, with methotrexate, vinblastine, doxorubicin, plus cisplatin in patients with bladder cancer. J Clin Oncol. 2005 Jul 20. 23(21):4602-8. [View Abstract]
  124. Iwasaki K, Obara W, Kato Y, Takata R, Tanji S, Fujioka T. Neoadjuvant gemcitabine plus carboplatin for locally advanced bladder cancer. Jpn J Clin Oncol. 2013 Feb. 43(2):193-9. [View Abstract]
  125. Bellmunt J, Théodore C, Demkov T, Komyakov B, Sengelov L, Daugaard G, et al. Phase III trial of vinflunine plus best supportive care compared with best supportive care alone after a platinum-containing regimen in patients with advanced transitional cell carcinoma of the urothelial tract. J Clin Oncol. 2009 Sep 20. 27(27):4454-61. [View Abstract]
  126. Rosenberg JE, Hoffman-Censits J, Powles T, van der Heijden MS, Balar AV, Necchi A, et al. Atezolizumab in patients with locally advanced and metastatic urothelial carcinoma who have progressed following treatment with platinum-based chemotherapy: a single-arm, multicentre, phase 2 trial. Lancet. 2016 Mar 4. [View Abstract]
  127. Balar AV, Galsky MD, Rosenberg JE, Powles T, Petrylak DP, Bellmunt J, et al. Atezolizumab as first-line treatment in cisplatin-ineligible patients with locally advanced and metastatic urothelial carcinoma: a single-arm, multicentre, phase 2 trial. Lancet. 2017 Jan 7. 389 (10064):67-76. [View Abstract]
  128. Powles T, Durán I, van der Heijden MS, et al. Atezolizumab versus chemotherapy in patients with platinum-treated locally advanced or metastatic urothelial carcinoma (IMvigor211): a multicentre, open-label, phase 3 randomised controlled trial. Lancet. 2018 Feb 24. 391 (10122):748-757. [View Abstract]
  129. Sharma P, Retz M, Siefker-Radtke A, Baron A, Necchi A, Bedke J, et al. Nivolumab in metastatic urothelial carcinoma after platinum therapy (CheckMate 275): a multicentre, single-arm, phase 2 trial. Lancet Oncol. 2017 Jan 25. [View Abstract]
  130. Powles T, O'Donnell PH, Massard C, Arkenau HT, Friedlander TW, Hoimes CJ, et al. Efficacy and Safety of Durvalumab in Locally Advanced or Metastatic Urothelial Carcinoma: Updated Results From a Phase 1/2 Open-label Study. JAMA Oncol. 2017 Sep 14. 3 (9):e172411. [View Abstract]
  131. Apolo AB, Infante JR, Balmanoukian A, Patel MR, Wang D, Kelly K, et al. Avelumab, an Anti-Programmed Death-Ligand 1 Antibody, In Patients With Refractory Metastatic Urothelial Carcinoma: Results From a Multicenter, Phase Ib Study. J Clin Oncol. 2017 Apr 4. JCO2016716795. [View Abstract]
  132. Balar AV, Castellano D, O'Donnell PH, Grivas P, Vuky J, Powles T, et al. First-line pembrolizumab in cisplatin-ineligible patients with locally advanced and unresectable or metastatic urothelial cancer (KEYNOTE-052): a multicentre, single-arm, phase 2 study. Lancet Oncol. 2017 Nov. 18 (11):1483-1492. [View Abstract]
  133. Bellmunt J, de Wit R, Vaughn DJ, Fradet Y, Lee JL, Fong L, et al. Pembrolizumab as Second-Line Therapy for Advanced Urothelial Carcinoma. N Engl J Med. 2017 Mar 16. 376 (11):1015-1026. [View Abstract]
  134. Siefker-Radtke AO, Necchi A, Park SH, Garcia-Donas J, Huddart RA, Frederick E, et al. First results from the primary analysis population of the phase 2 study of erdafitinib (ERDA; JNJ-42756493) in patients (pts) with metastatic or unresectable urothelial carcinoma (mUC) and FGFR alterations (FGFRalt) (Abstract 4503). Presented at the American Society of Clinical Oncology 2018 Annual Meeting. 2018 Jun 03. Chicago, IL.
  135. Rosenberg JE, O'Donnell PH, Balar AV, McGregor BA, Heath EI, Yu EY, et al. Pivotal Trial of Enfortumab Vedotin in Urothelial Carcinoma After Platinum and Anti-Programmed Death 1/Programmed Death Ligand 1 Therapy. J Clin Oncol. 2019 Oct 10. 37 (29):2592-2600. [View Abstract]
  136. [Guideline] Chang SS, Boorjian SA, Chou R, Clark PE, Daneshmand S, Konety BR, et al. Diagnosis and Treatment of Non-Muscle Invasive Bladder Cancer: AUA/SUO Guideline. J Urol. 2016 Jun 16. [View Abstract]
  137. [Guideline] Rouprêt M, Babjuk M, Böhle A, et al. European Association of Urology guidelines on upper tract urothelial carcinomas: 2019 update. European Association of Urology. Available at http://uroweb.org/guideline/upper-urinary-tract-urothelial-cell-carcinoma/. Accessed: June 22, 2019.
  138. [Guideline] Bellmunt J, Orsola A, Leow JJ, Wiegel T, De Santis M, Horwich A, et al. Bladder cancer: ESMO Practice Guidelines for diagnosis, treatment and follow-up. Ann Oncol. 2014 Sep. 25 Suppl 3:iii40-8. [View Abstract]
  139. [Guideline] U.S. Preventive Services Task Force. Final Recommendation Statement: Bladder Cancer in Adults: Screening. Available at https://www.uspreventiveservicestaskforce.org/Page/Document/RecommendationStatementFinal/bladder-cancer-in-adults-screening. May 2019; Accessed: June 22, 2019.
  140. Recommendations for Clinical Preventive Services: Bladder Cancer, Adults. Leawood, KS: American Academy of Family Physicians; 2011. American Academy of Family Physicians. Available at http://www.aafp.org/patient-care/clinical-recommendations/all/bladder-cancer.html. Accessed: May 11, 2017.
  141. Bladder Cancer. American Cancer Society. Available at https://www.cancer.org/cancer/bladder-cancer.html. January 30, 2019; Accessed: December 19, 2019.
  142. Flaig TW. NCCN Guidelines Updates: Management of Muscle-Invasive Bladder Cancer. J Natl Compr Canc Netw. 2019 May 1. 17 (5.5):591-593. [View Abstract]
  143. Ploeg M, Kums AC, Aben KK, van Lin EN, Smits G, Vergunst H, et al. Prognostic factors for survival in patients with recurrence of muscle invasive bladder cancer after treatment with curative intent. Clin Genitourin Cancer. 2011 Sep. 9(1):14-21. [View Abstract]
  144. Bruins HM, Djaladat H, Ahmadi H, Sherrod A, Cai J, Miranda G, et al. Incidental cancer of the prostate in patients with bladder urothelial carcinoma: comprehensive analysis of 1476 radical cystoprostatectomy specimens. J Urol. 2013 May 23. [View Abstract]
  145. Zehnder P, Studer UE, Daneshmand S, et al. Outcomes of radical cystectomy with extended lymphadenectomy alone in patients with lymph node-positive bladder cancer who are unfit for or who decline adjuvant chemotherapy. BJU Int. 2014 Apr. 113(4):554-60. [View Abstract]

Cross-section through the bladder, uterus, and vagina with squamous cell carcinoma of the bladder infiltrating through the bladder wall into the vaginal wall.

The classic appearance of carcinoma in situ as a flat, velvety patch. However, using special staining techniques such as 5-aminolevulinic acid, it has been shown that significant areas of carcinoma in situ are easily overlooked by conventional cystoscopy. Courtesy of Abbott and Vysis Inc.

Papillary bladder tumors such as this one are typically of low stage and grade (Ta-G1). Courtesy of Abbott and Vysis Inc.

Sessile lesions as shown usually invade muscle, although occasionally a tumor is detected at the T1-G3 stage prior to muscle invasion. Courtesy of Abbott and Vysis Inc.

Histopathology of bladder shows eggs of Schistosoma haematobium surrounded by intense infiltrates of eosinophils and other inflammatory cells.

(A) When infused into the bladder, the optical imaging agent hexaminolevulinate (Cysview) accumulates preferentially in malignant cells. (B) On blue-light cystoscopy, the collection of hexaminolevulinate within tumors is visible as bright red spots. Courtesy of Gary David Steinberg, MD, FACS.

In an ileal conduit, a small segment of ileum is taken out of continuity with the gastrointestinal tract but is maintained on its mesentery. Ureters are anastomosed to one end of this ileal segment, and the other end is brought out as a stoma to the abdominal wall.

In an Indiana pouch, a urinary reservoir is created from detubularized right colon and an efferent limb of terminal ileum. Terminal ileum is plicated and brought to the abdominal wall. The continence mechanism is the ileocecal valve.

In an orthotopic neobladder, a segment of ileum is used to construct a neobladder, which is connected to the urethra. Orthotopic neobladder most closely restores the natural storage and voiding function of the native bladder.

In an ileal conduit, a small segment of ileum is taken out of continuity with the gastrointestinal tract but is maintained on its mesentery. Ureters are anastomosed to one end of this ileal segment, and the other end is brought out as a stoma to the abdominal wall.

In an Indiana pouch, a urinary reservoir is created from detubularized right colon and an efferent limb of terminal ileum. Terminal ileum is plicated and brought to the abdominal wall. The continence mechanism is the ileocecal valve.

In an orthotopic neobladder, a segment of ileum is used to construct a neobladder, which is connected to the urethra. Orthotopic neobladder most closely restores the natural storage and voiding function of the native bladder.

The classic appearance of carcinoma in situ as a flat, velvety patch. However, using special staining techniques such as 5-aminolevulinic acid, it has been shown that significant areas of carcinoma in situ are easily overlooked by conventional cystoscopy. Courtesy of Abbott and Vysis Inc.

Papillary bladder tumors such as this one are typically of low stage and grade (Ta-G1). Courtesy of Abbott and Vysis Inc.

Sessile lesions as shown usually invade muscle, although occasionally a tumor is detected at the T1-G3 stage prior to muscle invasion. Courtesy of Abbott and Vysis Inc.

Photograph in which fluorescence in situ hybridization centromere staining identifies aneuploidy of chromosome 3. Multiple instances of overexpression of the chromosome (note the multiple red dots, which identify centromeres of this chromosome) prove aneuploidy.

Cross-section through the bladder, uterus, and vagina with squamous cell carcinoma of the bladder infiltrating through the bladder wall into the vaginal wall.

High power, Pap stain showing high grade urothelial carcinoma on a bladder wash cytology.

Intermediate power, H and E stain of urothelial carcinoma in situ. The superficial cells shed into the urine and correlate with those seen in cytologic bladder washing or urine cytology.

High power, H and E stain of high grade urothelial carcinoma. This tumor is now invasive into the muscularis propria (smooth muscle seen in center of image).

Histopathology of bladder shows eggs of Schistosoma haematobium surrounded by intense infiltrates of eosinophils and other inflammatory cells.

(A) When infused into the bladder, the optical imaging agent hexaminolevulinate (Cysview) accumulates preferentially in malignant cells. (B) On blue-light cystoscopy, the collection of hexaminolevulinate within tumors is visible as bright red spots. Courtesy of Gary David Steinberg, MD, FACS.

Cystoscopy Findings Urine Cytology Findings FISH* Findings Action
NegativeNegativeNegative†Routine follow-up
NegativeNegativePositive‡Increased frequency of surveillance, whether FISH findings are false positive or anticipatory positive
NegativePositiveNegative or positiveCancer until proven otherwise
  • Upper tract imaging with contrast
  • Cystoscopy with retrograde pyelography, washings, and/or ureteroscopy
  • Evaluate urethra
  • Increased frequency of surveillance upon negative findings
*FISH - Fluorescent in situ hybridization.



†Negative predictive value 95%.



‡Positive predictive value 30%.



Stage Recurrence, % Progression, %
Ta5511
T16131
CIS4523
G1612-4
G2565-7
G350-7030-40
Early Complications Rate, % Late Complications Rate, %
Ileus10Small-bowel obstruction7.4
Wound infection5.5Ureteroenteric stricture7.0
Sepsis4.9Renal calculi3.9
Pelvic abscess4.7Acute pyelonephritis3.1
Hemorrhage3.4Parastomal hernia2.8
Wound dehiscence3.3Stomal stenosis2.8
Bowel obstruction3.0Incisional hernia2.2
Enterocutaneous fistula2.2Fistula1.3
Rectal injury2.2Rectal complications< 1