Esophageal Cancer

Back

Practice Essentials

Esophageal cancer is a disease in epidemiologic transition. Until the 1970s, the most common type of esophageal cancer in the United States was squamous cell carcinoma, which has smoking and alcohol consumption as risk factors. Since then, there has been a steep increase in the incidence of esophageal adenocarcinoma, for which the most common predisposing factor is gastroesophageal reflux disease (GERD). See the image below.



View Image

Cascade of events that lead from gastroesophageal reflux disease to adenocarcinoma.

Signs and symptoms

Presenting signs and symptoms of esophageal cancer include the following:

Physical findings include the following:

See Presentation for more detail.

Diagnosis

Laboratory studies such as complete blood count (CBC) and comprehensive metabolic panel (CMP) focus principally on patient factors that may affect treatment (eg, nutritional status, renal function).

Imaging studies used for diagnosis and staging include the following:

For staging information, see Esophageal Cancer Staging.

See Workup for more detail.

Management

Treatment of esophageal cancer varies by disease stage, as follows:

Indications for surgical treatment of esophageal cancer include the following:

Contraindications for surgical treatment include the following:

Surgical options include the following:

Palliative care options for patients who are not candidates for surgery are as follows:

See Treatment and Medication for more detail.

For patient education information, see the Heartburn and GERD Center and Esophageal Cancer (Cancer of the Esophagus).

Background

Esophageal cancer is a devastating disease. It is the 6th most common cause of cancer deaths worldwide.[3] Although some patients can be cured, the treatment for esophageal cancer is protracted, diminishes quality of life, and is lethal in a significant number of cases.

The principal histologic types of esophageal cancer are squamous cell carcinoma (SCC) and adenocarcinoma. Both are common in men. SCC is the most common histology in Eastern Europe and Asia, while adenocarcinoma is most common in North America and Western European countries.

Adenocarcinoma is diagnosed predominantly in white men and the incidence has risen more steeply in that population. However, adenocarcinoma is gradually increasing in men of all ethnic backgrounds and also in women.[4]

Squamous cells line the entire esophagus, so SCC can occur in any part of the esophagus, but it often arises in the upper half. Adenocarcinoma typically develops in specialized intestinal metaplasia (Barrett metaplasia) that develops as a result of gastroesophageal reflux disease (GERD); thus, adenocarcinoma typically arises in the lower half of the distal esophagus and often involves the esophagogastric junction.

Treatment history and controversy

Surgery has traditionally been the treatment for esophageal carcinoma. The first successful resection was performed in 1913 by Torek.[5] In the 1930s, Ohsawa in Japan and Marshall in the United States were the first to perform successful single-stage transthoracic esophagectomies with continent reconstruction.[6, 7] Nonoperative therapy is usually reserved for patients who are not candidates for surgery because of clinical conditions or advanced disease.

The ideal treatment for localized esophageal cancer is sometimes debated across practice cultures and subspecialties. Proponents of surgical treatment argue that resection is the only treatment modality to offer curative intent; proponents of the nonsurgical approach claim that esophagectomy has a prohibitive index of mortality and that esophageal cancer is an incurable disease.

Most recently, in a study of 12,298 patients with esophageal cancer, median survival in patients who underwent surgery was 35.6 months, versus 24.8 months in the 708 (6%) patients who were advised to have surgery but declined and chose other treatment modalities or no treatment.[8]

A propensity-matched analysis showed that 525 patients who had preoperative therapy followed by surgery had a median survival of 32.3 months, compared with 21.9 months in patients who refused surgery, In a multivariate analysis, refusal of surgery remained a strong predictor of poorer survival (odds ratio, 1.72; P< 0.001).[8]

 

Anatomy

The esophagus is a muscular tube that extends from the level of the 7th cervical vertebra to the 11th thoracic vertebra. The esophagus can be divided into the following anatomic parts:

The blood supply of the cervical esophagus is derived from the inferior thyroid artery, while the blood supply for the thoracic esophagus comes from the bronchial arteries and the aorta. The abdominal esophagus is supplied by branches of the left gastric artery and inferior phrenic artery.

Venous drainage of the cervical esophagus is through the inferior thyroid vein, while the thoracic esophagus drains via the azygous vein, the hemiazygous vein, and the bronchial veins. The abdominal esophagus drains through the coronary vein.

The esophagus is characterized by a rich network of lymphatic channels in the submucosa that can facilitate the longitudinal spread of neoplastic cells along the esophageal wall. Lymphatic drainage is to the following node basins:

Pathophysiology

Major risk factors for SCC include alcohol consumption and tobacco use. Most studies have shown that alcohol is the primary risk factor but smoking in combination with alcohol consumption can have a synergistic effect. 

Alcohol damages the cellular DNA by decreasing metabolic activity within the cell and therefore inhibits detoxification and promotes oxidation.[9] Alcohol is a solvent, specifically of fat-soluble compounds. Therefore, the carcinogens within tobacco are able to penetrate the esophageal epithelium more easily. 

Some of the carcinogens in tobacco include the following:

Other carcinogens, such as nitrosamines found in certain salted vegetables and preserved fish, have also been implicated in esophageal SCC. The pathogenesis appears to be linked to inflammation of the squamous epithelium that leads to dysplasia and in situ malignant transformation.[10]

Adenocarcinoma of the esophagus most commonly occurs in the distal esophagus and has a distinct relationship to GERD. Untreated GERD can progress to Barrett esophagus (BE), in which the stratified squamous epithelium that normally lines the esophagus is replaced by a columnar epithelium.

The chronic reflux of gastric acid and bile at the gastroesophageal junction and the subsequent damage to the esophagus has been implicated in the pathogenesis of Barrett metaplasia. Diagnosis of Barrett esophagus can be confirmed by biopsies of the columnar mucosa during an upper endoscopy.

Barrett esophagus incidence increases with age. The disorder is uncommon in children. It is more common in men than women and more common in whites than in Asians or African Americans.[9]

The progression of Barrett metaplasia to adenocarcinoma is associated with several changes in gene structure, gene expression, and protein structure.[11, 12, 13] The oncosuppressor gene TP53 and various oncogenes, particularly erb-b2, have been studied as potential markers. Casson and colleagues identified mutations in the TP53 gene in patients with Barrett epithelium associated with adenocarcinoma.[14] In addition, alterations in p16 genes and cell cycle abnormalities or aneuploidy appear to be some of the most important and well-characterized molecular changes.

Obesity is another risk factor for esophageal adenocarcinoma, specifically in individuals with central fat distribution. Hypertrophied adipocytes and inflammatory cells within fat deposits create an environment of low-grade inflammation and promote tumor development through the release of adipokines and cytokines.[15]  Adipocytes in the tumor microenvironment supply energy production and support tumor growth and progression.[16]

 

 

Etiology

The etiology of esophageal carcinoma is thought to be related to exposure of the esophageal mucosa to noxious or toxic stimuli, resulting in a sequence of dysplasia to carcinoma in situ to carcinoma. In Western cultures, retrospective evidence has implicated cigarette smoking and chronic alcohol exposure as the most common etiologic factors for squamous cell carcinoma. High body mass index, GERD, and resultant Barrett esophagus are often the associated factors for esophageal adenocarcinoma.[17]

Certain factors can increase or decrease risk for both cancer types. A study by Steevens et al found that in current smokers, increased consumption of specific groups of vegetables and fruits were inversely associated with risk for esophageal SCC and adenocarcinoma.[18]  Total vegetable consumption nonsignificantly reduced risk for both esophageal cancer types. Consumption of raw vegetables and of citrus fruits was inversely associated with risk for esophageal adenocarcinoma. The risk of both SCC and adenocarcinoma of the esophagus was increased in current smokers.[19]

Risk factors for esophageal squamous cell carcinoma

The risk factors and etiologic associations for SCC of the esophagus include the following:

Smoking and alcohol use

The Netherlands Cohort Study, a prospective study in 120,852 participants, demonstrated the combined effects of smoking and alcohol consumption on risk of SCC of the esophagus.[19] In participants who drank 30 g or more of ethanol daily, the multivariable adjusted incidence rate ratio (RR) for esophageal SCC was 4.61 compared with abstainers. The RR for current smokers who consumed more than 15 g/day of ethanol was 8.05 when compared with nonsmokers who consumed less than 5 g/day of ethanol.

No associations were found between alcohol consumption and esophageal adenocarcinoma.

Dietary associations

Esophageal SCC has a wide variety of dietary associations, among them the following:  

A variety of other factors may promote esophageal SCC. These include the following:

A Chinese study found that risk for esophageal cancer was five times higher in individuals who drank very hot tea and drank more than 15 g of alcohol every day compared with those who drank tea less than once a week and consumed fewer than 15 g of alcohol daily (hazard ratio [HR], 5.00). Risk was doubled in those who drank very hot tea daily and smoked tobacco, compared with nonsmokers who drank tea only occasionally (HR, 2.03). The analysis included 456,155 participants aged 30 to 79 years who did not have a prior history of cancer.[32, 33]

A genome-wide association study by Wu et al identified seven susceptibility loci on chromosomes 5q11, 6p21, 10q23, 12q24, and 21q22, suggesting the involvement of multiple genetic loci and gene-environment interaction in the development of esophageal SCC.[34]  

Bisphosphonate use can result in esophagitis and has been suggested as a risk factor for esophageal carcinoma. However, a large study found no significant difference in the frequency of esophageal or gastric cancers between the bisphosphonate cohort and the control group.[35]

Infections

Human papillomavirus (HPV) infection has been recognized as a contributing factor to esophageal cancer. However, Sitas et al reported limited serologic evidence of an association between esophageal SCC and HPV in a study of more than 4000 subjects. The study could not exclude the possibility that certain HPV types may be involved in a small subset of cancers, although HPV does not appear to be an important risk factor.[36] Helicobacter pylori infection, which can cause stomach cancer, has not been associated with esophageal cancer.

Tylosis

Tylosis is a rare autosomal dominant disease caused by a mutation in TEC (tylosis with esophageal cancer), a tumor suppressor gene located on chromosome 17q25. Tylosis is associated with hyperkeratosis of the palms and soles (see the images below) and a high rate of esophageal SCC (40% to 90% by the age of 70 years).[37] The inherited type of tylosis (Howell-Evans syndrome) has been most strongly linked to esophageal SCC.[38, 39] Surveillance by upper GI endoscopy is recommended for family members with tylosis after 20 years of age.[40]



View Image

Palmoplantar keratoderma (Tylosis) of palms (A) and soles (B). Courtesy of The American Journal of Gastroenterology, Nature Publishing Group.

Risk factors for adenocarcinoma

The principal risk factors and etiologic associations for esophageal adenocarcinoma include the following:

Gastroesophageal reflux disease

GERD is the most common predisposing factor for adenocarcinoma of the esophagus. Adenocarcinoma may represent the last event of a sequence that starts with irritation caused by the reflux of acid and bile and progresses to specialized intestinal (Barrett) metaplasia, low-grade dysplasia, high-grade dysplasia, and finally adenocarcinoma see the image below). Approximately 10%-15% of patients who undergo endoscopy for evaluation of GERD symptoms are found to have Barrett epithelium.



View Image

Cascade of events that lead from gastroesophageal reflux disease to adenocarcinoma.

In 1952, Morson and Belcher published the first description of a patient with adenocarcinoma of the esophagus arising in a columnar epithelium with goblet cells.[41] In 1975, Naef et al emphasized the malignant potential of Barrett esophagus.[42]

The risk of adenocarcinoma in patients with Barrett metaplasia has been estimated to be 30-60 times that of the general population. A nationwide population-based case-control study performed in Sweden found an odds ratio of 7.7 for adenocarcinoma in persons with recurrent symptoms of reflux, as compared with persons without such symptoms, and an odds ratio of 43.5 in patients with long-standing and severe symptoms of reflux.[17]

Although the annual risk of developing esophageal adenocarcinoma in people with GERD has been reported at 0.5%, some studies have found lower risk. Data from the Northern Ireland Barrett Esophagus Register, which is one of the largest population-based registries in the world, found that malignant progression in patients with Barrett esophagus was 0.22% per year. This suggests that current surveillance approaches may not be cost effective.[43]

A study by Hvid-Jensen et al examined a large Danish registry (11,028 patients over a median of 5.2 years) and found the incidence of esophageal adenocarcinoma to be 1.2 cases per 1000 person-years (or 0.12% annual risk). Low-grade dysplasia detected on index endoscopy was associated with an incidence rate of 5.1 cases per 1000 person-years, compared with 1 per 1000 person-years in those without dysplasia.[44]

Additional factors that increase the risk for esophageal adenocarcinoma, particularly in patients with Barrett esophagus, include cigarette smoking[45] and certain polymorphisms of the epidermal growth factor gene that have been associated with higher serum levels of epidermal growth factor.[46] Freedman et al suggested a possible association between cholecystectomy and esophageal adenocarcinoma, likely due to the toxic effect of refluxed duodenal juice containing bile on esophageal mucosa [47]

Obesity and metabolic syndrome

Obesity has been linked to a higher risk for Barrett esophagus and esophageal adenocarcinoma.[48, 49] Individuals in the highest quartile for body mass index (BMI) have a 7.6-fold higher risk of developing esophageal adenocarcinoma compared with those in the lowest quartile. A meta-analysis of case control and cohort studies revealed a relative risk for esophageal adenocarcinoma of 1.71 (95% confidence index [CI] 1.5-1.96) for BMI between 25 and 30 kg/m2, and 2.34 (95% CI 1.95-2.81) for BMI ≥30 kg/m2.[50] Obesity does not appear to increase the risk of esophageal SCC.[51]

Obesity increases the risk of GERD and subsequently of esophageal adenocarcinoma by a "mechanical" process that consists of an amplification of intragastric pressure, disruption of normal esophageal sphincter function, and increased risk of a hiatal hernia.[52] Obesity also has an inflammatory effect mediated by the release of various proinflammatory cytokines, which can lead to metabolic syndrome, a constellation of metabolic disorders that includes obesity, impaired fasting glucose, high blood pressure, and dyslipidemia. Like obesity, metabolic syndrome is also linked with the risk of esophageal adenocarcinoma.[53]

Epidemiology

United States statistics

The American Cancer Society estimates that 17,650 new cases of esophageal cancer (13,750 in men and 3,900 in women) will occur in the United States in 2019, and that 16,080 persons (13, 020 men and 3,060 women) will die of the disease.[54] Esophageal cancer is the seventh most common cause of cancer death in males.[54] The 5-year survival rate from 2007 to 2013 was 18.8%.[55]

The incidence of esophageal carcinoma is approximately 3-6 cases per 100,000 population, although certain endemic areas appear to have higher per-capita rates. The age-adjusted annual incidence is 4.3 cases per 100,000 population.[55]

The epidemiology of esophageal carcinoma has changed markedly over the past several decades in the United States.[56] Until the 1970s, squamous cell carcinoma was the most common type of esophageal cancer (90-95%). It was typically located in the thoracic esophagus and most frequently affected African-American men with a long history of smoking and alcohol consumption.

Subsequently, rates of esophageal adenocarcinoma rose markedly, particularliy in whites. In white men, the incidence rate of esophageal adenocarcinoma exceeded that of squamous cell carcinoma around 1990, while in white women aged 45–59 years, adenocarcinoma overtook squamous cell carcinoma in 2006–2010.[57]

From 1973 to 1996, the incidence of esophageal adenocarcinoma increased by 8.2% annually. From 1996 to 2006, the rate of increase fell to 1.3% annually, principally because of a plateau in the incidence of early-stage disease. Prior to 1996, early-stage cases increased by 10% annually; subsequently, they declined by 1.6% annually.[58] From 2004 to 2014, incidence rates of esophageal adenocarcinoma in the United States fell on average 1.4% each year.[55]

International statistics

Esophageal cancer is the ninth most common cancer and the sixth most common cause of cancer deaths worldwide.[59] It is endemic in many parts of the world, particularly in the third world countries, where it is the fourth most common cause of cancer deaths.[59] Incidence rates are variable worldwide, with the highest rates found in southern and eastern Africa and eastern Asia and the lowest rates in western and central Africa and Central America in both men amd women.[59]

In some regions, such as areas of northern Iran, some areas of southern Russia, and northern China (sometimes called an "esophageal cancer belt"), the incidence of esophageal carcinoma may be as high as 800 cases per 100,000 population. Major risk factors in these areas are not well known but are probably related to the poor nutritional status, including low intake of fruits and vegetables and drinking very hot beverages. Unlike in the United States, squamous cell carcinoma is responsible for 95% of all esophageal cancers worldwide.

Sex- and age-related demographics

Esophageal cancer is more common in men than in women. The male-to-female ratio is 3-4:1.

Esophageal cancer occurs most commonly during the sixth and seventh decades of life. The disease becomes more common with advancing age; it is about 20 times more common in persons older than 65 years than it is in individuals below that age. Median age at diagnosis is 68 years.[55]

Prognosis

Survival in patients with esophageal cancer depends on the stage of the disease. Squamous cell carcinoma and adenocarcinoma, stage-by-stage, appear to have equivalent survival rates.

Lymph node or solid organ metastases are associated with low survival rates. In 2009-2015, the overall 5-year survival rate for esophageal cancer was 19.9%.[55]  Patients without lymph node involvement have a significantly better prognosis and 5-year survival rate than patients with involved lymph nodes. Stage IV lesions with distant metastasis are associated with a 5-year survival rate of around 5%. (See the table below.)

Table 2. Five-year esophageal cancer survival rates  by stage at diagnosis in the US, 2009-2015[55]



View Table

See Table

The 5-year survival rate in 2015 was 21.5% in whites and 13.5% in blacks.[60] A report of 1085 patients who underwent transhiatal esophagectomy for cancer showed that the operation was associated with a 4% operative mortality rate and a 23% 5-year survival rate. A better 5-year survival rate (48%) was identified in a subgroup of patients who had a complete response (ie, disappearance of the tumor) following preoperative radiation and chemotherapy (ie, neoadjuvant therapy).[61]

Transhiatal and transthoracic esophagectomies have equivalent long-term survival rates.[62, 63]

Imaging and prognosis

Suzuki et al found that a higher initial standardized uptake value on positron emission tomography (PET) scanning is associated with poorer overall survival in patients with esophageal or gastroesophageal carcinoma receiving chemoradiation. The authors suggested that PET scanning may become useful for individualizing therapy.[64]

A study by Gillies et al also found that PET–computed tomography (CT) scanning can be used to predict survival; in this study, the presence of fluorodeoxyglucose (FDG)-avid lymph nodes was an independent adverse prognostic factor.[65]

HER-2 and prognosis

A study by Prins et al of human epidermal growth factor 2 (HER-2) protein overexpression and HER-2 gene amplification in esophageal carcinomas found that HER-2 positivity and gene amplification are independently associated with poor survival. In their study, which involved 154 patients with esophageal adenocarcinoma, HER-2 positivity was seen in 12% of these patients and overexpression was seen in 14% of them.[66]

History

Dysphagia, the most common presenting symptom of esophageal cancer, is initially experienced for solids but eventually progresses to include liquids. It usually occurs when esophageal lumen diameter is under 13 mm and indicates locally advanced disease. A complaint of dysphagia in an adult should always prompt an endoscopy to help rule out the presence of esophageal cancer. A barium swallow study is also indicated in these cases.

Other symptoms include the following:

Physical Examination

Physical examination findings in patients with esophageal cancer are typically normal, unless the cancer has metastasized to neck nodes or the liver. Lymphadenopathy in the laterocervical or supraclavicular area or the presence of hepatomegaly often indicates unresectable disease.

Approach Considerations

In 2013, the Society of Thoracic Surgeons released clinical practice guidelines to assist in the diagnosis and treatment of localized esophageal cancer. Their recommendations for diagnosis include the following[67] :

Imaging Studies

Imaging studies used in the diagnosis and staging of esophageal cancer include the following:

Computed tomography

Abdominal and chest computed tomography (CT) scans are useful for helping to exclude the presence of metastases (M staging) to the lungs and liver and may be useful for helping to determine whether adjacent structures have been invaded.[68] (See the image below.)



View Image

Chest CT scan showing invasion of the trachea by esophageal cancer.

Positron emission tomography

PET scanning is also a useful baseline imaging technique and is increasingly becoming standard in the staging of esophageal cancer. It may be particularly useful in detecting occult distant lymph node metastases and bone spread. In addition, the intensity of radiopharmaceutical uptake on PET scans may reflect the biology of the cancer and thus may have prognostic significance.[64]

Endoscopic ultrasound

EUS is the most sensitive test for determining the depth of tumor penetration (T staging) and the presence of enlarged periesophageal lymph nodes (N staging).[69, 70]

Following are the characteristic features of malignant or inflammatory lymph nodes detected on EUS:

The accuracy of diagnosing nodal disease is significantly increased with the combination of above-mentioned features, but also is confirmed with the use of fine needle aspiration (FNA) biopsy for cytology assessment.[71] The combined use of EUS and FNA (EUS-FNA) has a greater accuracy than EUS alone in the evaluation of lymph node metastasis.[72] In a study that compared the role of CT, EUS, and EUS-FNA for preoperative nodal staging in 125 patients with esophageal cancer, EUS-FNA was more sensitive than CT (83% vs. 29%) and more accurate than CT (87% vs. 51%) or EUS (87% vs. 74%) for nodal staging.[73]

Patients with obstructing tumors are at increased risk for perforation during staging EUS. The risk of perforation can be reduced with the use of wire-guided or mini-EUS probes. In certain cases, the malignant stricture is dilated prior to the staging EUS.The review of CT and PET scans prior to EUS is recommended to evaluate the nodal distribution for a possible FNA biopsy.

Bronchoscopy

Bronchoscopy is indicated for cancers of the middle and upper third of the thoracic esophagus (tumor at or above carina) to help exclude invasion of the trachea or bronchi. It should be performed only if the patient shows no evidence of M1 disease. Laparoscopy and thoracoscopy have a greater than 92% accuracy in staging regional nodes.

Barium swallow

Barium swallow is very sensitive for detecting strictures (see the first image below) and intraluminal masses (see the second image below) but does not allow staging and biopsy. It is now rarely used, but it may be helpful for studying the distal anatomy in obstructive tumors that are inaccessible by endoscopy.



View Image

Barium swallow demonstrating stricture due to cancer.



View Image

Barium swallow demonstrating an endoluminal mass in the mid esophagus.

For more information, see the Medscape article Esophageal Carcinoma Imaging.

Staging

Esophageal cancer staging follows the tumor-node-metastasis (TNM) classification of the American Joint Cancer Committee/Union for International Cancer Control/ (AJCC/UICC).[74]

No completely satisfactory method is available to clinically stage esophageal cancer. The difficulty of clinically assessing the disease is reflected by changes over time in the AJCC staging system. The 1983 system was based on the length of the intraluminal esophageal tumor, the presence of esophageal obstruction, and the involvement of palpable lymph nodes. This clinical staging system proved to have limited value.

The 1988 revision defined a clinical and pathologic staging system based entirely on the depth of esophageal wall invasion and the presence or absence of local nodal involvement. Neither of those parameters is assessed easily on a clinical basis. Hofstetter et al therefore proposed incorporating the number of involved lymph nodes with regional and nonregional node location.[75] This modification, which seemed to be simpler and to better predict long-term survival, was adopted into the revised system.

The revised 2010 AJCC staging classification was based on the risk-adjusted random forest analysis of the data generated by the Worldwide Esophageal Cancer Collaboration (WECC) for 4627 patients who were treated with primary esophagectomy without preoperative or postoperative therapy.[76] In the data reported by the WECC, survival decreased with increasing depth of tumor invasion (T), presence of regional lymph node metastases (N), and the presence of distant metastases (M).[77]

The 2017 TNM classification for esophageal cancer is shown below (staging is detailed in Tables 2-4, below. T staging is illustrated in the image below). For more information, see Esophageal Cancer Staging.

TNM staging is as follows:



View Image

Diagram showing T1,T2 and T3 stages of esophageal cancer. Courtesy of Cancer Research UK and Wikimedia Commons.

 

Table 3. Clinical Staging Classification (Squamous Cell Carcinoma)



View Table

See Table

 

Table 4. Clinical Staging Classification (Adenocarcinoma)



View Table

See Table

 

All esophageal tumors, as well as tumors with epicenters within 5 cm of the esophagogastric junction that also extend into the esophagus, are classified and staged according to the AJCC/UICC esophageal scheme. Tumors with an epicenter in the stomach that are more than 5 cm from the esophagogastric junction or those within 5 cm of the esophagogastric junction without extension into the esophagus are staged using the gastric carcinoma scheme.

However, this classification may not work well for patients who have received preoperative therapy. Some other shortcomings associated with the current staging classification are as follows:

Other classifications—such as that of the Japanese Society for Esophageal Diseases, which is widely used in Asia—differ from that of the AJCC/UICC, especially regarding lymph node distribution and nomenclature.[78]

Laboratory Studies

Laboratory studies in patients with esophageal cancer focus principally on patient factors that may affect treatment. These include complete blood count (CBC) and comprehensive metabolic panel (CMP). Nutritional status should be evaluated in patients with dysphagia; liver function studies should be performed in patients who abuse alcohol.

Procedures

Upper GI endoscopy

Upper GI endoscopy allows direct visualization and biopsies of the tumor. (See the image below.)



View Image

Endoscopy demonstrating intraluminal esophageal cancer.

Endoscopy is a very important tool in the diagnosis, staging, and surveillance of patients with esophageal cancer. Most endoscopy procedures are performed under conscious sedation. Patients who are at risk of aspiration during endoscopy may require general anesthesia. 

Diagnostic endoscopies are performed to determine the following:

Esophageal tumor length, as assessed by preoperative endoscopy, has been identified as an independent predictor of long-term survival in patients with adenocarcinoma of the esophagus.[79] The 5-year survival rate was significantly higher for patients with a tumor length of 2 cm or less (78% vs 29% for those with a tumor length of more than 2 cm).

Endoscopic resection

Endoscopic resection (ER) of focal nodules should be performed in the setting of early-stage disease (T1a or T1b) to provide accurate assessment of depth of invasion, degree of differentiation, and the presence of lymphovascular invasion.[80] Thus, ER is an essential procedure for the accurate staging of early-stage cancer especially in patients with small nodular lesions (≤2 cm).[81] ER can become a therapeutic procedure if a small lesion (under 2 cm) is fully removed and histopathology reveals that the lesion is well differentiated, with penetration limited to submucosa, absence of lymphovascular invasion, and clear margins.[82]

Histologic Findings

Histologically, esophageal squamous cell carcinoma is characterized microscopically by keratinocyte-like cells with intercellular bridges or keratinization. Adenocarcinomas that arise from Barrett esophagus mucosa are typically well- or moderately differentiated and have well-formed tubular or papillary structures. In poorly differentiated adenocarcinomas, glandular structures are only sloightly formed; in undifferentiated adenocarcinomas, glandular structures are absent. See the images below.



View Image

Micrograph of squamous cell carcinoma of the esophagus (H&E Stain). Courtesy of Wikimedia Commons.



View Image

Low magnification micrograph of an intramucosal esophageal adenocarcinoma (H&E stain). Endoscopic mucosal resection specimen. Courtesy of Wikimedia Co....

Approach Considerations

Treatment of esophageal cancer varies according to stage—locoregional (stages I-III) versus metastatic cancer (stage IV)—and  histologic subtype—squamous cell carcinoma (SCC) versus adenocarcinoma.

National Comprehensive Cancer Network (NCCN) treatment recommendations for esophageal cancer include the following[83] :

Surgical Indications and Contraindications

Surgery remains the cornerstone of treatment for esophageal cancer. Indications for surgery include the following:

Contraindications to surgery include the following:

In addition, the presence of severe, associated comorbid conditions (eg, cardiovascular disease, respiratory disease) can decrease a patient's chances of surviving an esophageal resection. Consequently, cardiac and respiratory function must be carefully evaluated preoperatively. A forced expiratory volume in 1 second of less than 1.2 L and a left ventricular ejection fraction of less than 0.4 are relative contraindications to the operation.

Esophagectomy

Esophageal resection (esophagectomy) remains a critical component of multimodality therapy for patients with tumors of any stage. Endoscopic mucosal resection is an experimental approach to patients with T1a disease or high-grade dysplasia that is limited to certain centers and performed only under protocol. Esophagectomy is no longer is used for palliation of symptoms because other treatment modalities have become available for relieving dysphagia.

An esophagectomy can be performed by using an abdominal and a cervical incision with blunt mediastinal dissection through the esophageal hiatus (ie, transhiatal esophagectomy [THE]) or by using an abdominal and a right thoracic incision (ie, transthoracic esophagectomy [TTE]).

THE offers the advantage of avoiding a chest incision, which can cause prolonged discomfort and can further aggravate the condition of patients with compromised respiratory function. After removal of the esophagus, continuity of the gastrointestinal tract is usually reestablished using the stomach.

Some authors have questioned the validity of THE as a cancer operation because part of the operation is not performed under direct vision and fewer lymph nodes are removed than with TTE. However, many retrospective studies and 2 prospective ones have shown no difference in survival between the operations, suggesting that the factor influencing survival is not the type of operation but, rather, the stage of the cancer at the time the operation is performed.[62, 63, 86, 87, 88, 89]

Morbidity and mortality

Complications from esophagectomy occur in approximately 40% of patients. The morbidity associated with the surgery consists mostly of respiratory, cardiac, and septic complications, including the following:

Anastomotic leaks and stricture may require dilatation (20%). Leaks may be treated with endoscopic placement of self-expanding, removable plastic stents.[90]

Leak rates vary depending on whether the anastomosis is in the chest (3%-12%) or the neck (10%-25%).[91] The choice of location for the anastomosis is based mostly on the location of the tumor and the surgeon’s assessment of the risks and benefits of a thoracic anastomosis. Such anastomoses have a lower leak rate, but an intrathoracic leak following esophagectomy can lead to sepsis and death.

A retrospective review of 1223 esophagectomies for cancer found that surgical management of intrathoracic leaks did not increase the patient mortality rate or effect long-term survival.[91]

As with other complex operations (eg, cardiac operations, resection of the pancreas or liver), the lowest mortality rate with esophagectomy is achieved when the procedure is performed in high-volume centers by high-volume surgeons. In California from 1990-1994, for instance, 5 high-volume centers had a mortality rate of 5% or less for esophageal resection for cancer, while the state’s average mortality rate for this surgery was approximately 18%.[92]

The better results in high-volume centers are due to a team approach. In these facilities, expert surgeons work with intensivists, cardiologists, pulmonologists, radiologists, and nurses who have experience and expertise.

Transthoracic esophagectomy

There are two types of TTE, as follows:

For TTE, the patient is placed supine on the operating room table. An arterial line, a central venous catheter, a Foley catheter, and a dual-lumen endotracheal tube are placed. Preoperative antibiotics are administered. An upper midline incision is made.

After exploring the peritoneal cavity for metastatic disease (if metastases are found, the operation is not continued), the stomach is mobilized. The right gastric and the right gastroepiploic arteries are preserved, while the short gastric vessels and the left gastric artery are divided.

Next, the gastroesophageal junction is mobilized, and the esophageal hiatus is enlarged. A pyloromyotomy is performed, and a feeding jejunostomy is placed for postoperative nutritional support.

After closure of the abdominal incision, the patient is repositioned in the left lateral decubitus position and a right posterolateral thoracotomy is performed in the fifth intercostal space.

The azygos vein is divided to allow full mobilization of the esophagus. The stomach is delivered into the chest through the hiatus and is then divided approximately 5 cm below the gastroesophageal junction.

An anastomosis (hand-sewn or stapled) is performed between the esophagus and the stomach at the apex of the right chest cavity. Then, the chest incision is closed.

McKeown esophagectomy, with an anastomosis in the cervical region, is similar in conduct, but with the advantage of being applicable for tumors in the upper, middle, and lower thoracic esophagus.

Transhiatal esophagectomy

For THE, the preoperative details are similar to those of TTE, except that a single-lumen, rather than a double-lumen, endotracheal tube is used. The neck is prepared in the operative field.

The abdominal part of the operation is identical to the TTE; however, dissection of the esophagus is performed through the enlarged esophageal hiatus without opening the right chest. The esophagus is mobilized in this fashion all the way to the thoracic inlet.

Then, a 6-cm incision is made in the left side of the neck. The internal jugular vein and carotid artery are retracted laterally, and the esophagus is identified and isolated posterior to the airway. To prevent injury to the left recurrent laryngeal nerve, no mechanical retractors are used to retract the trachea.

Next, after resection of the proximal stomach and thoracic esophagus, the remaining stomach is pulled up through the posterior mediastinum until it reaches the remaining esophagus at the cervical level. Then, a hand-sewn anastomosis is performed, and a small drain is placed in the neck alongside the anastomosis. The abdominal and neck incisions are closed. (See the image below.)



View Image

Transhiatal esophagectomy in which (a) is the abdominal incision, (b) is the cervical incision, and (c) is the stomach stretching from abdomen to the ....

Minimally invasive esophagectomy

The use of laparoscopic and thoracoscopic techniques has revolutionized the treatment of benign esophageal disorders such as achalasia and gastroesophageal reflux disease (GERD). Advantages of minimally invasive surgery include a shorter hospital stay, less postoperative discomfort, and much faster recovery time than with open surgery. Minimally invasive esophagectomy (MIE) is finding a place in the treatment of esophageal cancer.[93]

In a study of MIE (mainly using thoracoscopic mobilization) in 222 patients, the mortality rate was only 1.4% and hospital stay was only 7 days, which is less than with most open procedures; only 16 patients (7.2%) required conversion to an open procedure.[94] A report by Luketich et al involving 56 patients also showed that MIE was comparable to open esophagectomy but the use of neoadjuvant treatment slightly increased the surgical mortality from 1.5% to 1.8%.[94]

In a randomized French study that compared transthoracic open esophagectomy (n=104) with MIE (hybrid procedure; n=103), .Mariette et al reported that the incidence of intraoperative and postoperative major complications, specifically pulmonary complications, was 69% lower with the hybrid procedure, while 3- and 5-year overall survival and disease-free survival were noninferior. For the hybrid procedure, the abdominal portion of the operation was performed through five small incisions, rather than the long abdominal incision used in TTE.[95, 96]

Video-assisted thoracoscopy (VATS) is being used in many centers for the thoracic mobilization of the esophagus, reducing the size of the chest incision. In addition, laparoscopy can be used to mobilize the gastric conduit in the abdomen, reducing abdominal incision size as well.

A study by Uenosono et al found that sentinel node mapping can be applied to patients with clinical T1 and N0 esophageal cancer. Use of this technique may facilitate less invasive surgery, with reduction of lymphadenectomy.[97]

Endoscopic mucosal resection (EMR) is a modern, attractive option for the treatment of superficial esophageal cancers. High-grade dysplasia and mucosa-limited neoplasms are candidates for EMR, because of the low risk of node metastasis in these cases. A population-based study of 1618 patients with grade Tis, T1a, or T1b esophageal cancer found that overall survival times and esophageal-cancer-specific survival times with endoscopic therapy were similar to those with surgery, after adjustment for patient and tumor factors.[98, 99]

Salvage endoscopic resection

In patients with local failure after definitive chemoradiotherapy (CRT) for esophageal squamous cell carcinoma (SCC), salvage endoscopic treatment (SET) may be a viable option, according to a study reported at the 2014 Gastrointestinal Cancers Symposium.[100] The study included 716 esophageal SCC patients treated with CRT, 417 of whom experienced local failure (incomplete response or local recurrence); of these 417 patients, 164 underwent SET (either photodynamic therapy or endoscopic resection).

Curative resection was achieved in 88% of the patients who underwent endoscopic resection, and a complete response was achieved in 57.5% of those who underwent photodynamic therapy.[100] Overall survival and relapse-free survival rates at 5 years were 38.6% and 28%, respectively. The factors most strongly predictive of improved survival were (1) an absence of lymph node metastasis before CRT and (2) an elapsed time of 6 months or longer between the initiation of CRT and the performance of SET.[100]

Chemotherapy and Radiotherapy

Chemotherapy and radiotherapy for esophageal cancer are delivered preoperatively. No survival benefit is obtained when radiation and chemotherapy are administered postoperatively; however, postoperative continuance of chemotherapy started preoperatively may be beneficial.[101] The aims of preoperative (neoadjuvant) chemotherapy and radiotherapy are to reduce the bulk of the primary tumor before surgery to facilitate higher curative resection rates and to eliminate or delay the appearance of distant metastases.

Most chemotherapy that is currently used for the treatment of esophageal cancer, including alkylating, antimetabolite, anthracycline, and antimicrotubular agents, are not approved for this indication by the US Food and Drug Administration (FDA). Chemotherapy for squamous cell esophageal carcinoma, as with squamous cell carcinomas in general, is based on cisplatin, while chemotherapy for esophageal adenocarcinoma has been extrapolated from experience in patients with adenocarcinoma of the stomach.

Neoadjuvant chemotherapy alone appears to offer a limited benefit at best. A North American randomized trial found that preoperative chemotherapy with a combination of cisplatin and fluorouracil did not improve overall survival among patients with squamous cell cancer or adenocarcinoma of the esophagus. In a larger trial, British investigators found that preoperative chemotherapy with those 2 agents resulted in a 5-year survival rate of 23.0%, compared with 17.1% for surgery alone.[102]

In contrast, the Chemoradiotherapy for Oesophageal Cancer Followed by Surgery Study (CROSS) demonstrated considerable benefit from preoperative chemoradiation over surgery alone in selected patients with esophageal or esophagogastric junction cancer (tumor stage T1N1 or T2–T3 with any N).[103] Median overall survival with chemoradiation therapy followed by surgery was 49.4 months, compared with 24.0 months with surgery alone.

An analysis of data from CROSS I and II showed that preoperative chemoradiotherapy (CRT) plus surgery was superior to surgery alone in preventing local, regional, and distant recurrence, particularly hematogenous metastasis and peritoneal carcinomatosis.[104, 105, 106] Overall recurrence rates were 35% for CRT plus surgery and 58% for surgery alone. The rates of locoregional recurrence, peritoneal carcinomatosis, and hematogenous dissemination were all lower for the former as well (14% vs 34%, 4% vs 14%, and 29% vs 35%, respectively).

Neoadjuvant therapy consists of a combination of radiotherapy (approximately 45 Gy) and chemotherapy with cisplatin and 5-fluorouracil. While the radiotherapy acts locally at the tumor site, the chemotherapy acts on tumor cells that have already spread. This combination therapy is usually administered over a 45-day period; esophageal resection is performed after an interval of approximately 4 weeks.

In a study of patients with distal esophageal carcinoma, Franko and colleagues reported that a longer interval between neoadjuvant chemoradiation and esophagectomy was significantly associated with worse outcomes. The median interval between radiation therapy and esophagectomy was 7.1 weeks; when the delay was 9 weeks or more, perioperative mortality was increased and overall survival decreased.[107]

Rohatgi et al reported that the response to preoperative chemoradiotherapy correlated strongly with overall survival and disease-free survival in patients with esophageal cancer. In their review of 235 cases, survival decreased progressively between patients who achieved a pathologic complete response, those who had a partial response (1-50% residual carcinoma in the resected specimen), and those with no response (greater than 50% residual carcinoma).[108]

A study by Rao et al identified groups of esophageal tumors with distinct gene expression profiles, which in future may allow for tailored treatment protocols.[109] Similarly, Alexander et al identified DNA-repair biomarkers that predict response to neoadjuvant chemotherapy.[110]

A trial involving 111 patients undergoing chemoradiotherapy for head-and-neck or esophageal cancer indicated that enteral nutrition enriched with n-3 fatty acids helps to preserve body mass and improve nutritional and functional status parameters during chemoradiotherapy.[111, 112]

Pembrolizumab

In July 2019, the FDA approved pembrolizumab (Keytruda) for patients with recurrent, locally advanced or metastatic, squamous cell carcinoma of the esophagus (ESCC) whose tumors express PD-L1 (Combined Positive Score [CPS] ≥10), as determined by an FDA-approved test, with disease progression after ≥1 prior lines of systemic therapy.[113]

Efficacy was observed in two clinical trials, KEYNOTE‑181 and KEYNOTE‑180. KEYNOTE-181 (n=628) was a phase 3 randomized, open-label, active-controlled trial that enrolled patients with recurrent locally advanced or metastatic esophageal cancer who progressed on or after one prior line of systemic treatment for advanced or metastatic disease. Patients were randomized to receive either pembrolizumab or the investigator’s choice of a paclitaxel, docetaxel, or irinotecan regimen as a control arm. The median overall survival was 9.3 months and 6.7 months in the pembrolizumab and control arms, respectively.[114]

KEYNOTE‑180 was a single arm, open-label trial that enrolled patients with locally advanced or metastatic esophageal cancer who progressed on or after at least 2 prior systemic treatments for advanced disease. In the patients with ESCC expressing PD-L1 CPS ≥10, ORR was 20% and response durations ranged from 4.2 to 25.1+ months, with 71% having responses of 6 months or longer and 57% having responses of 12 months or longer.[115]

Tipiracil/trifluridine

The FDA approved tipiracil/trifluridine in February 2019 for metastatic gastric or gastroesophageal junction (GEJ) adenocarcinoma previously treated with at least 2 prior lines of chemotherapy that included a fluoropyrimidine, a platinum, either a taxane or irinotecan, and if appropriate, HER2/neu-targeted therapy.

Approval of tipiracil/trifluridine was based on the phase 3 TAGS (TAS-102 Gastric Study) clinical trial (n=507). Patients in the tipiracil/trifluridine group had a median overall survival 5.7 months compared with 3.6 months in the placebo group (one-sided P=0·00029, two-sided P=0·00058).[116]

Palliative Care

In patients who are not candidates for surgery, because of their clinical condition or advanced disease, treatment focuses on control of dysphagia. The goal of palliative care is to prevent and relieve suffering and improve quality of life for patients and their caregivers regardless of the disease stage. In patients with unresectable or locally advanced cancer, palliative interventions provide symptomatic relief and may result in significant prolongation of life, improvement in nutritional status, the sensation of well-being, and overall quality of life.

Although various treatment options are available for the management of dysphagia, optimal treatment is still debated. A multimodality interdisciplinary approach is strongly encouraged.

Dysphagia is the most common symptom in patients with esophageal cancer. Assessing the severity of the condition and swallowing impairment is essential to initiate appropriate interventions for long-term palliation of dysphagia in patients with esophageal cancer.

Available palliative methods for the management of dysphagia include the following:

National Comprehensive Cancer Network guidelines for best supportive/palliative care

For patients with complete esophageal obstruction, the NCCN guidelines[83] recommend the following:

Surgical or radiologic placement of jejunostomy or gastronomy tubes may be necessary to provide adequate hydration and nutrition, if endoscopic lumen restoration is not undertaken or is unsuccessful. Brachytherapy may be considered instead of EBRT if lumen can be restored using appropriate applicators during the delivery of brachytherapy to avoid an excessive dose on mucosal surfaces. In a multicenter randomized trial, single-dose brachytherapy was associated with fewer complications and better long-term relief of dysphagia compared with metal stents.[117]  

For patients with severe esophageal obstruction (those able to swallow liquids only), some additional options include endoscopic lumen enhancement (wire-guided or balloon dilation) and endoscopy or fluoroscopy-guided placement of covered expandable metal stents. While some data suggest a lower migration and re-obstruction rate with the larger-diameter covered expandable metal stents, there may be a higher risk of stent-related complications[118]  

Stents

Long-term palliation of dysphagia can be achieved with endoscopic, radiographic-assisted insertion of expandable metal or plastic stents.[119, 120] Temporary placement of self-expanding metal stents (SEMS) with concurrent EBRT was found to increase survival rates compared with permanent stent placement.[121] SEMS is the preferred treatment for patients with tracheoesophageal fistula and those who are not candidates for chemoradiation or who failed to achieve adequate palliation with such therapy.[122] Membrane-covered stents have significantly better palliation than conventional bare metal stents because of a lower rate of tumor ingrowth.[120]

Radiotherapy

Radiation therapy is successful in relieving dysphagia in approximately 50% of patients. In patients with advanced esophageal cancer, the preoperative combination of chemotherapy and radiotherapy has shown good results.

In a large, multicenter study, Herskovic and colleagues reported a 2-year survival rate of 38%, with a median survival period of 12.5 months, for patients treated with radiotherapy in combination with chemotherapy (fluorouracil and cisplatin), compared with a 2-year-survival rate of 10% and a median survival period of 8.5 months in patients treated with radiotherapy alone.[123, 124]

Folkert et al found that high-dose-rate (HDR) endoluminal brachytherapy was well tolerated in medically inoperable patients with superficial primary or recurrent esophageal cancer. Over the course of 3 years, 14 patients were treated with HDR intraluminal brachytherapy; 10 had recurrent esophageal cancer and 4 had previously unirradiated lesions. The overall freedom from failure (OFFF) and the overall survival (OS) rate at 18 months were 30.8% and 72.7%, respectively. Patients with recurrent disease had an 11.1% OFFF and a 55.6% OS rate at 18 months. For patients with previously unirradiated disease, the OFFF was 75% and the OS rate was 100%.[125]

Chemotherapy

Chemotherapy as a single modality has limited use. Only a few patients achieve a modest and short-lived response.

A phase 3 study from the United Kingdom suggests that docetaxel may be useful as a second-line treatment for patients with esophageal cancer who have progressed after first-line chemotherapy.[126, 127] Median overall survival was significantly better in patients treated with docetaxel than in those managed with active symptom control (5.2 versus 3.6 months, respectively).

Because survival in these patients is measured in months, quality of life is an important consideration. In the UK study, quality-of-life questionnaires demonstrated no differences between the 2 groups on global and functioning scores but did indicate an improvement in symptom scores, with the docetaxel group reporting less pain.[126, 127]

In 2006, a Cochrane review tried to assess the effectiveness of chemotherapy versus best supportive care, as well as that of different chemotherapy regimens against each other, in metastatic esophageal carcinoma. The authors found that no consistent benefit with any specific chemotherapy regimen.[128] Cisplatin, 5-fluorouracil (5-FU), paclitaxel, and anthracyclines had promising response rates and tolerable toxicity.

Laser therapy

Laser therapy (Nd:YAG laser) can help to achieve temporary relief of dysphagia in as many as 70% of patients. Multiple sessions are usually required to keep the esophageal lumen patent.

The photosensitizer porfimer (Photofrin) is FDA approved for palliation of patients with completely obstructing esophageal cancer or partially obstructing cancer that cannot be satisfactorily treated with Nd:YAG laser therapy. Intravenous injection of porfimer is followed 40–50 hours later with delivery of 630 nm wavelength laser light; a second laser light treatment may be given 96–120 hours after the injection. Photodynamic therapy (PDT) with porfimer sodium has overall equal efficacy to Nd:YAG laser thermal ablation for palliation of dysphagia in esophageal cancer, and equal or better objective tumor response rate. Temporary photosensitivity is a limitation, but PDT is carried out with greater ease and is associated with fewer acute perforations than Nd:YAG laser therapy[119] .

Postoperative Care and Follow-up

The average length of postoperative hospital stay for patients with esophageal cancer is 9-14 days. Patients usually spend the first postoperative night in the intensive care unit (ICU).

Patients can be extubated immediately after the operation, but mechanical ventilation should be continued if any concerns about the respiratory status are present. Respiratory complications (eg, atelectasis, pleural effusion, pneumonia) and cardiac complications (eg, cardiac arrhythmias) usually occur in the first postoperative days. Patients leave the ICU and are transferred to the surgical ward only when their respiratory status and cardiac status are satisfactory.

Feeding through the feeding jejunostomy begins on postoperative day 1. On postoperative day 6, a swallow study is performed to check for anastomotic leakage. If no leak is present, patients start oral feedings. If a leak is present, the drainage tubes are left in place and nutrition is provided entirely through the feeding jejunostomy until the leak closes spontaneously.

Approximately 85%-90% of patients go home after discharge. The remaining patients may need additional time in a skilled nursing facility if they live alone and if they cannot take care of themselves.

Patients are seen by the responsible surgeon at 2 weeks and 4 weeks after discharge from the hospital and subsequently every 6 months by an oncologist. Most patients return to their regular level of activities within 2 months.

Prevention of Esophageal Cancer

For squamous cell carcinoma, prevention consists of smoking cessation, efforts to reduce alcohol abuse, and consumption of a diet containing an adequate amount of fruits, vegetables, and vitamins. Both the National Comprehensive Cancer Network (NCCN) guidelines and the National Cancer Institute (NCI) cancer information summary for esophageal cancer prevention conclude that smoking cessation decreases the risk of squamous cell carcinoma.[83, 129]

For esophageal adenocarcinoma, prevention involves stopping the sequence of events leading from gastroesophageal reflux disease (GERD) to Barrett esophagus to adenocarcinoma. Better control of gastroesophageal reflux can prevent the development of Barrett metaplasia in patients with GERD and can discourage the development of high-grade dysplasia in patients with metaplasia. Endoscopic follow-up evaluations should be performed at 1- to 2-year interval to detect the presence of dysplasia, allowing intervention before cancer develops

Hereditary Cancer Predisposition Syndromes Associated with an Increased Risk for Esophageal and EGJ Cancers

Surveillance upper endoscopy with biopsy (see the table below) should be considered for patients who have one of the following hereditary cancer predisposition syndromes:

Table 5. Hereditary Cancer Predisposition Syndromes Associated with an Increased Risk for Esophageal Cancers.



View Table

See Table

 

Screening and Surveillance

Recommendations for screening and surveillance of patients with GERD and/or Barrett esophagus have been issued by the following organizations:

None of the organizations recommend endoscopic screening of the general population with GERD. There is general agreement among the guidelines that patients with chronic GERD and multiple other risk factors associated with esophageal adenocarcinoma should undergo upper gastrointestinal endoscopy to screen for Barrett esophagus or esophageal adenocarcinoma.[137, 138, 139, 140, 141] Those additional risk factors include the following:

The 2015 American Society for Gatrointestinal Endoscopy (ASGE) guidelines for the use of endoscopy in the management of GERD recommends endoscopic screening in select patients with multiple risk factors for Barrett esophagus be considered, but also advises that patients be informed that there is insufficient evidence that this practice prevents cancer or prolongs survival.[141]

The American College of Physicians' (ACP) best practice advice, issued in 2012, offers the following recommendations for upper endoscopy for GERD[140] :

The American College of Gastroenterology (ACG) guidelines for the surveillance and therapy of Barrett esophagus[139] are listed in the table below.

 

Table 6. ACG Recommendations on surveillance of Barrett esophagus



View Table

See Table

The ACG recommends that endoscopic ablative therapies should not be routinely used in patients with nondysplastic Barrett esophagus, because of their low risk of progression to esophageal adenocarcinoma (strong recommendation, very low level of evidence). However, endoscopic eradication therapy is the procedure of choice for patients with confirmed dysplasia, whether low or high grade.

The ACG advises that antireflux surgery should not be used as an antineoplastic measure.[139] Endoscopic resection (ER) and mucosal radiofrequency ablation (RFA) has become the preferred treatment for most patients with Barrett esophagus and HGD. Alternative strategies include cryoablation or photodynamic therapy (PDT)[142, 143] . Surgical resection is reserved for patients with HGD and characteristics that are unfavorable for non-surgical therapy, such as nodularity or long-segment involvement. For patients with metaplasia or LGD, gastroesophageal reflux is controlled with histamine receptor antagonists or proton pump inhibitors.

Radiofrequency ablation

RFA has been gaining popularity as a treatment for Barrett esophagus with dysplasia. Shaheen et al reported complete eradication of all dysplasia 2 years after RFA in 101 of 106 patients (95%). After 3 years, dysplasia remained eradicated in more than 85% of patients, without maintenance RFA. The rate of serious adverse events was 3.4%, and the rate of esophageal stricture was 7.6%.[144]

Photodynamic therapy

PDT involves the administration of photosensitizing chromophores, which are selectively retained by dysplastic malignant tissue. Light is then delivered in the area. The photosensitizer absorbs photons, becomes photoexcited, and transfers its energy to a chemical substrate that causes biologic damage to the abnormal tissue. A drawback of PDT is the formation of esophageal strictures in 34% of patients. The photosensitizer porfimer is FDA approved for ablation of high-grade dysplasia in Barrett esophagus.

Surveillance of Esophageal Cancer

Surveillance strategies after successful local therapy of esophageal cancers remain controversial since very limited prospective data is available on effective surveillance strategies.

In general, for asymptomatic patients, follow-up should include a complete history and physical examination every 3 to 6 months for 1 to 2 years, then every 6 to 12 months for 3 to 5 years, and annually thereafter. CBC, comprehensive serum chemistry evaluation, upper GI endoscopy with biopsy, and imaging studies should be obtained as clinically indicated. In addition, some patients may require dilatation of an anastomotic or a chemoradiation-induced stricture. Nutritional assessment and counseling may be extremely valuable. 

The stage-specific recommendations for surveillance included in the NCCN Guidelines are based on the available evidence from retrospective studies[145, 146]  and the expertise of the panel members.

Stage (0-I / Tis, T1a, T1b): Evidence-based guidelines have not been established for all stages of completely treated early stage esophageal cancer. The recommendations outlined in the guidelines are based on available evidence from clinical trials and current practice. Endoscopic surveillance with upper GI endoscopy (EGD) is recommended for patients with Tis, T1a and T1b tumors, after completion of endoscopic therapy. It is recommended to perform EGD  every 3 months for the first year, every 4-6 months for the second year, and annually for 3 more years. In patients with T1b tumors treated with esophagectomy, endoscopic surveillance with EGD should be done as clinically indicated based on the symptoms and radiographic findings. Routine imaging studies are not recommended for patients with Tis and T1a tumors.

Stage (II-III / T2-T4, N0-N+, T4b): Locoregional recurrences are common after bimodality therapy. Therefore, EGD is a valuable surveillance tool following bimodality therapy. In patients treated with bimodality therapy, the majority of recurrences (95%) occur within 24 months. Thus, surveillance for at least 24 months is recommended following bimodality therapy[147]

Locoregional relpases are uncommon following trimodality therapy.[148] Therfore, EGD for surveillance is not recommended after trimodality therapy. Other imaging modalities (e.g., PET/CT or CT chest/abdomen with contrast) are used for most luminal recurrences. In patients treated with trimodality therapy, the majority of recurrences (90%) occur within 36 months of surgery. Therefore, surveillance for at least 36 months is recommended following trimodality therapy.

 

Evaluation

In 2013, the Society of Thoracic Surgeons released clinical practice guidelines to assist in the diagnosis and treatment of localized esophageal cancer. Their recommendations include the following[67] :

The 2013 American Society for Gastrointestinal Endoscopy (ASGE) guidelines for use of endoscopy in the assessment and treatment of esophageal cancer make the following recommendations for accurate staging[149] :

Guidelines from the College of American Pathologists, American Society for Clinical Pathology, and American Society of Clinical Oncology, issued in 2016, include the following recommendations for clinicians[150] :

National Comprehensive Cancer Network (NCCN) guidelines for workup include the following[83] :

Management

American College of Gastroenterology (ACG) recommendations regarding endoscopic treatment of Barrett esophagus include the following[139] :

National Comprehensive Cancer Network (NCCN) treatment recommendations for esophageal cancer include the following[83] :

NCCN guidelines for surgical therapy of esophageal cancer are as follows[83] :

NCCN recommendations for first-line systemic therapy of advanced or metastatic disease are as follows[83] :

NCCN recommendations for second-line and subsequent systemic therapy of advanced or metastatic disease are as follows[83] :

NCCN recommendations for targeted therapy are as follows[83] :

In 2014, the Society of Thoracic Surgeons (STS) released clinical practice guidelines for multimodal treatment of esophageal cancer with the following recommendations[167] :

The 2013 American Society for Gastrointestinal Endoscopy (ASGE) guidelines note that argon plasma coagulation (APC), heater probe, cryotherapy, or radiofrequency ablation should not be used as monotherapy with curative intent for T1a tumors. However, those techniques may have a role in ablation of remaining high-risk tissue following resection[149] .

Palliative Care

NCCN guidelines recommend the following for palliative/best supportive care[83] ::

The 2013 American Society for Gastrointestinal Endoscopy (ASGE) guidelines have following recommendations for palliative care[149] :

Medication Summary

Most of the chemotherapy regimens currently used for the treatment of esophageal cancer, including alkylating, antimetabolite, anthracycline, and antimicrotubular agents, have not been approved by the US Food and Drug Administration (FDA). Following are the recommended regimens for the treatment of esophageal cancer. These are based on extrapolations from the published literature and clinical practice. 

For preoperative chemoradiation, preferred regimens (category 1) are as follows:

Other preoperative chemoradiation regimens (category 2B) are as follows:

For definitive chemoradiation, preferred regimens are as follows:

Other definitive chemoradiation regimens are as follows:

Perioperative chemotherapy is used only for adenocarcinoma of the thoracic esophagus. Three preoperative cycles and posteroperative  cycles of the following may be given:

Preoperative chemotherapy is used only for adenocarcinoma of the thoracic esophagus. A total of two cycles of fluorouracil and cisplatin (category 2B) is given.

With postoperative chemoradiation, fluoropyrimidine (infusional fluorouracil or capecitabine) is given before and after fluoropyrimidine-based chemoradiation.

For postoperative chemotherapy, capecitabine plus oxaliplatin is used.

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.

Capecitabine (Xeloda)

Clinical Context:  Capecitabine is a pyrimidine antimetabolite and a prodrug of fluorouracil. It forms the active moiety, fluorouracil, by undergoing hydrolysis in the liver and tissues. Capecitabine is used for the treatment of esophageal cancer, which is an off-label indication.

Class Summary

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

Cisplatin

Clinical Context:  Intrastrand cross-linking of DNA and inhibition of DNA precursors are among the proposed mechanisms of action for cisplatin. This agent is used in combination with radiation therapy. Cisplatin has black box warnings for adverse reactions, 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 esophageal cancer, which is an off-label indication. Carboplatin has black box warnings including bone marrow suppression, anaphylactic reactions, and vomiting.

Oxaliplatin (Eloxatin)

Clinical Context:  Oxaliplatin is a platinum alkylating agent that inhibits DNA replication and transcription, resulting in cell death. It can be used in combination chemotherapy for the treatment of esophageal cancer, which is an off-label indication. It has a black box warning for anaphylactic reactions, which can be managed with epinephrine, corticosteroids, and antihistamines.

Class Summary

These agents inhibit cell growth and proliferation, interfering with DNA synthesis by the formation of DNA cross-links. Alkylating agents can have serious adverse effects such as 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 in combination with cisplatin and fluorouracil for the treatment of esophageal cancer, which is an off-label indication. It has several black box warnings such as bone marrow suppression, fluid retention, and hypersensitivity reactions.

Use of docetaxel is not recommended in certain patients with hepatic impairment. Patients receiving treatment with docetaxel should be premedicated with corticosteroids the day before administration to help reduce fluid retention and hypersensitivity reactions.

Paclitaxel

Clinical Context:  Paclitaxel promotes microtubule assembly, interferes with the G2 mitotic phase, and inhibits cell replication. Although not FDA approved, it has been used in combination chemotherapy for the treatment of esophageal cancer. Paclitaxel has an off-label indication for the treatment of adenocarcinoma. Black box warnings for this drug include bone marrow suppression and hypersensitivity reactions.

Class Summary

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

Epirubicin (Ellence)

Clinical Context:  Epirubicin inhibits DNA and RNA synthesis. It can be used off label as part of a combination chemotherapy regimen for the treatment of esophageal cancer. It has several black box warnings, including bone marrow suppression, extravasation, myocardial toxicity, and secondary malignancy. Dosage reduction is recommended in patients with mild to moderate hepatic impairment.

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.

Irinotecan (Camptosar)

Clinical Context:  Irinotecan binds reversibly to the topoisomerase I–DNA complex and prevents the ligation of the cleaved DNA strand. It can be used as part of combination chemotherapy for the treatment of esophageal cancer, which is an off-label indication. Black box warnings for irinotecan include bone marrow suppression and diarrhea.

Class Summary

These agents prevent cell growth and proliferation. They work by binding to topoisomerase and causing single-strand DNA breaks.

Tipiracil/trifluridine (Lonsurf)

Clinical Context:  Tipiracil is a thymidine phosphorylase inhibitor that increases trifluridine exposure by inhibiting its metabolism. Trifluridine is a thymidine-based nucleoside analog that incorporates into DNA, interferes with DNA synthesis, and inhibits cell proliferation. It is indicated for metastatic gastric or gastroesophageal junction (GEJ) adenocarcinoma previously treated with at least 2 prior lines of chemotherapy that included a fluoropyrimidine, a platinum, either a taxane or irinotecan, and if appropriate, HER2/neu-targeted therapy.

Class Summary

Following uptake into cancer cells, trifluridine is incorporated into DNA, interferes with DNA synthesis and inhibits cell proliferation.

Porfimer (Photofrin)

Clinical Context:  Porfimer is a photodynamic therapy that causes cytotoxic activity by producing oxygen free-radicals in the presence of laser light. It also can release thromboxane A2, leading to necrosis and vascular occlusion. It is indicated for palliation in patients with partially or completely obstructing esophageal cancer.

Class Summary

This category includes miscellaneous antineoplastic agents that cause cytotoxic activity by various mechanisms of action.

What is esophageal cancer?What are the signs and symptoms of esophageal cancer?Which physical findings suggest esophageal cancer?What is the role of lab tests in the evaluation of esophageal cancer?What is the role of imaging studies in the evaluation of esophageal cancer?What are the treatment options for esophageal cancer?When is surgery indicated for esophageal cancer?What are the contraindications for surgery for esophageal cancer?What are surgical options for the treatment of esophageal cancer?What are palliative care options for esophageal cancer?What are patient education resources for esophageal cancer?How common is esophageal cancer?What are the principal histologic types of esophageal cancer?Who is at highest risk for adenocarcinoma esophageal cancer?Where in the esophagogastric junction do squamous cell carcinoma (SCC) and adenocarcinoma typically occur?What is preferred treatment for esophageal cancer?What is the anatomy of the esophagus?From which artery is the blood supply for the esophagus derived?How does venous drainage of the esophagus occur?What are the lymphatic channels in the esophagus that facilitate esophageal cancer?What are the major risk factors for SCC esophageal cancer?What is the role of alcohol in the pathophysiology of esophageal cancer?Which tobacco carcinogens have a role in the pathogenesis of esophageal cancer?What is the role of carcinogens in the pathogenesis of SCC esophageal cancer?What is the pathophysiology of adenocarcinoma esophageal cancer?What is the role of Barrett metaplasia in esophageal adenocarcinoma?What is the role of obesity in the pathogenesis of esophageal adenocarcinoma?What is the role of tylosis in the pathogenesis of SCC esophageal cancer?What is the etiology of esophageal cancer?What are the risk factors for esophageal cancer?What are the risk factors for SCC esophageal cancer?How much does smoking and alcohol consumption increase the risk of SCC esophageal cancer?Which dietary factors increase or decrease the risk of SCC esophageal cancer?Which genetic factors may increase the risk of SCC esophageal cancer?Which infections may be etiological factors in esophageal cancer?What are the principal etiologic factors for esophageal adenocarcinoma?What is the most common predisposing factor for esophageal adenocarcinoma?What is the risk of esophageal adenocarcinoma in patients with Barrett metaplasia?What is the risk of developing esophageal adenocarcinoma in patients with gastroesophageal reflux disease (GERD)?What are risk factors for esophageal adenocarcinoma?How does obesity affect the risk for esophageal adenocarcinoma?What is the incidence of esophageal cancer in the US?What is the global incidence of esophageal cancer?How does the incidence of esophageal cancer vary by sex?How does the incidence of esophageal cancer vary by age?What is the survival rate for esophageal cancer?Which PET scan finding suggests poorer overall survival in patients with esophageal cancer?What can be used to predict survival in esophageal cancer?Which genetic mutation is a negative prognostic indicator in esophageal cancer?What is the most common presenting symptom of esophageal cancer?What are the signs and symptoms of esophageal cancer?Which physical findings suggest patients with esophageal cancer?Which esophageal lesions should be included in the differential diagnosis of esophageal cancer?How is achalasia differentiated from esophageal cancer?How is esophageal stricture differentiated from esophageal cancer?What are the differential diagnoses for Esophageal Cancer?What are the Society of Thoracic Surgeons recommendations for the diagnosis of esophageal cancer?Which imaging studies are performed in the diagnosis and staging of esophageal cancer?What is the role of a barium swallow in the diagnosis and staging of esophageal cancer?What is the role of CT scanning in the diagnosis of esophageal cancer?What is the role of PET scanning in the evaluation of esophageal cancer?What is the role of EUS in the evaluation of esophageal cancer?Which EUS findings are characteristic of esophageal cancer?What is the accuracy of EUS for the diagnosis of esophageal cancer?What are the risks of EUS for staging esophageal cancer?What is the role of bronchoscopy in the evaluation of esophageal cancer?What classification systems are used outside the US for staging esophageal cancer?Which staging system is used for esophageal cancer?How is esophageal cancer staging evolved?What is the 2017 AJCC/UICC TNM classification for esophageal cancer?What are the limitations of the AJCC/UICC TNM staging system for esophageal cancer?What is the role of lab testing in the evaluation of esophageal cancer?What is the role of upper GI endoscopy in the evaluation and management of esophageal cancer?What do diagnostic endoscopies for esophageal cancer determine?How does tumor length predict long-term survival in esophageal adenocarcinoma?What is the role of endoscopic resection (ER) of focal nodules is the staging of esophageal cancer?Which histologic findings are characteristic of esophageal cancer?Which factors impact the choice of treatment for esophageal cancer?What are the NCCN treatment recommendations for esophageal cancer?What are the indications for surgery for esophageal cancer?What are the contraindications for surgery of esophageal cancer?Which comorbidities decrease survival rates from esophageal resection of esophageal cancer?When is esophageal resection (esophagectomy) indicated in the treatment of esophageal cancer?How is esophageal resection (esophagectomy) performed in the treatment of esophageal cancer?What are complications of esophageal resection (esophagectomy) in esophageal cancer?What is the treatment for anastomotic leaks from esophageal resection (esophagectomy)?How do leak rates vary by location in an esophageal resection (esophagectomy) for esophageal cancer?What has been shown to decrease the mortality rate of esophageal resection (esophagectomy) for esophageal cancer?What are the benefits of a team approach to surgery of esophageal cancer?What are the types of transthoracic esophagectomy (TTE) used for the treatment of esophageal cancer?How is a transthoracic esophagectomy (TTE) performed in esophageal cancer?How is a transhiatal esophagectomy (THE) performed in esophageal cancer?What are the advantages of minimally invasive surgery for the treatment of esophageal cancer?What is the role of video-assisted thoracoscopy (VATS) in the treatment of esophageal cancer?What is the role of sentinel node mapping for the treatment of esophageal cancer?What is the role of endoscopic mucosal resection (EMR) in the treatment of esophageal cancer?What is the indication for a salvage endoscopic therapy (SET) in esophageal cancer and what is its efficacy?What is the role of CRT in the treatment of esophageal cancer?What is the role of neoadjuvant chemotherapy for esophageal cancer?What is the benefit of preoperative CRT for the treatment of esophageal cancer?What is neoadjuvant therapy for esophageal cancer and what is its efficacy?What is the response rate to preoperative CRT for esophageal cancer and how affect survival rates?What is the role of biomarkers in the treatment of esophageal cancer?What is the role of enteral nutrition in the treatment of esophageal cancer?What is the efficacy of laser therapy for palliative treatment of esophageal cancer?What is the goal of palliative care for esophageal cancer?How does dysphagia affect palliative care selection for esophageal cancer?What palliative care options for the management of dysphagia in esophageal cancer?What are the NCCN recommendations for palliative care of complete esophageal obstruction in esophageal cancer?What is the role of stents in the palliation of dysphagia in esophageal cancer?What is the efficacy of radiotherapy for palliative treatment of esophageal cancer?What is the efficacy of chemotherapy for palliative treatment of esophageal cancer?What is the typical length of hospitalization for esophageal cancer treatment?When are patients extubated after surgery to treat esophageal cancer?When does postoperative feeding through the jejunostomy begin in esophageal cancer?How frequently do patients return home after surgery for esophageal cancer?What postoperative follow-up for esophageal cancer is required?How can esophageal cancer be prevented?What is the indication for surveillance upper endoscopy with biopsy for esophageal cancer?What are the ACP recommendations for upper endoscopy for gastroesophageal reflux disease (GERD)?Which organizations have released screening guidelines for GERD) and/or Barrett esophagus?What are the recommendation for screening for Barrett esophagus or esophageal adenocarcinoma?What are the ASGE guidelines for the use of endoscopy in the management of gastroesophageal reflux disease (GERD)?What are the ACG guidelines for the surveillance and therapy of Barrett esophagus?What is the indication for radiofrequency ablation (RFA) for the treatment of Barrett esophagus with dysplasia?What is PDT for the treatment of Barrett esophagus?How do surveillance strategies for esophageal cancer vary by disease stage?What are the STS recommendations for diagnosis and treatment of localized esophageal cancer?What are the ASGE recommendations for the use of endoscopy in the assessment and treatment of esophageal cancer?What are the CAP-ASCP-ASCO recommendations for HER2 testing in esophageal cancer?What are the NCCN guidelines for HER-2 testing in the evaluation for esophageal cancer?What are the ACG recommendations for endoscopic treatment of Barrett esophagus in esophageal cancer?What are the NCCN treatment recommendations for esophageal cancer?What are the NCCN guidelines for surgical therapy of esophageal cancer?What are the NCCN treatment guidelines for metastatic or recurrent esophageal cancer?What are the STS recommendations for multimodal treatment of esophageal cancer?What are the NCCN guidelines for palliative care for esophageal cancer?What are the ASGE recommendations for palliative care for esophageal cancer?Which chemotherapy regimens have been approved by the FDA for the treatment of esophageal cancer?Which preoperative chemoradiation regimens are used in the treatment of esophageal cancer?What are the definitive chemoradiation regimens used in esophageal cancer?When is perioperative chemotherapy used for the treatment of esophageal cancer?When is preoperative chemotherapy used for the treatment of esophageal cancer?Which regimen is used for preoperative and postoperative chemotherapy for esophageal cancer?Which medications in the drug class Thymidine Analog are used in the treatment of Esophageal Cancer?Which medications in the drug class Antineoplastics, Topoisomerase Inhibitors are used in the treatment of Esophageal Cancer?Which medications in the drug class Antineoplastics, Anthracycline are used in the treatment of Esophageal Cancer?Which medications in the drug class Antineoplastics, Antimicrotubular are used in the treatment of Esophageal Cancer?Which medications in the drug class Antineoplastics, Alkylating are used in the treatment of Esophageal Cancer?Which medications in the drug class Antineoplastics, Antimetabolite are used in the treatment of Esophageal Cancer?Which medications in the drug class Antineoplastics, Other are used in the treatment of Esophageal Cancer?

Author

Muhammad Masab, MD, Resident Physician, Department of Internal Medicine, Einstein Medical Center

Disclosure: Nothing to disclose.

Coauthor(s)

Claudia Mattos da Costa Dourado, MD, Clinical Assistant Professor of Medicine, Sidney Kimmel Medical College of Thomas Jefferson University; Associate Program Director, Hematology and Medical Oncology Fellowship Program, Attending Physician, Division of Hematology and Medical Oncology, Department of Medicine, Einstein Medical Center

Disclosure: Nothing to disclose.

Chief Editor

N Joseph Espat, MD, MS, FACS, Harold J Wanebo Professor of Surgery, Assistant Dean of Clinical Affairs, Boston University School of Medicine; Chairman, Department of Surgery, Director, Adele R Decof Cancer Center, Roger Williams Medical Center

Disclosure: Nothing to disclose.

Additional Contributors

Fernando AM Herbella, MD, PhD, TCBC, Affiliate Professor, Attending Surgeon in Gastrointestinal Surgery, Esophagus and Stomach Division, Department of Surgery, Federal University of Sao Paulo Medical School; Private Practice; Medical Examiner, Sao Paulo's Medical Examiner's Office Headquarters, Brazil

Disclosure: Nothing to disclose.

Keith M Baldwin, DO, IMPH, Assistant Professor of Surgery, Boston University School of Medicine; Endocrine and Surgical Oncologist, Department of General Surgery, Roger Williams Cancer Center

Disclosure: Nothing to disclose.

Marco G Patti, MD, Surgeon, UNC Hospitals Multispecialty Surgery Clinic

Disclosure: Nothing to disclose.

Acknowledgements

Philip Schulman, MD Chief, Medical Oncology, Department of Medicine, Memorial Sloan-Kettering Cancer Center

Philip Schulman, MD, is a member of the following medical societies: American Association for Cancer Research, American College of Physicians, American Society of Hematology, and Medical Society of the State of New York

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

References

  1. Edwards MJ, Gable DR, Lentsch AB, et al. The rationale for esophagectomy as the optimal therapy for Barrett's esophagus with high-grade dysplasia. Ann Surg. 1996 May. 223(5):585-9; discussion 589-91. [View Abstract]
  2. Ferguson MK, Durkin A. Long-term survival after esophagectomy for Barrett's adenocarcinoma in endoscopically surveyed and nonsurveyed patients. J Gastrointest Surg. 2002 Jan-Feb. 6(1):29-35; discussion 36. [View Abstract]
  3. Torre LA, Bray F, Siegel RL, Ferlay J, Lortet-Tieulent J, Jemal A. Global cancer statistics, 2012. CA Cancer J Clin. 2015 Mar. 65 (2):87-108. [View Abstract]
  4. Brown LM, Devesa SS, Chow WH. Incidence of adenocarcinoma of the esophagus among white Americans by sex, stage, and age. J Natl Cancer Inst. 2008 Aug 20. 100 (16):1184-7. [View Abstract]
  5. Torek F. The first successful resection of the thoracic portion of the esophagus for carcinoma. Surg Gynecol Obstet. 1913. 16:614-17.
  6. Ohsawa T. Surgery of the esophagus. Arch Jpn Chir. 1933. 10:605-8.
  7. Marshall SF. Carcinoma of the esophagus: successful resection of lower end of esophagus with reestablishment of esophageal gastric continuity. Surg Clin North Amer. 1938. 18:643.
  8. Brooks M. 'Striking' Survival Benefit With Surgery for Esophageal Cancer. Available at https://www.medscape.com/viewarticle/892133. February 2, 2018; Accessed: February 7, 2018.
  9. Napier KJ, Scheerer M, Misra S. Esophageal cancer: A Review of epidemiology, pathogenesis, staging workup and treatment modalities. World J Gastrointest Oncol. 2014 May 15. 6 (5):112-20. [View Abstract]
  10. Mao WM, Zheng WH, Ling ZQ. Epidemiologic risk factors for esophageal cancer development. Asian Pac J Cancer Prev. 2011. 12 (10):2461-6. [View Abstract]
  11. Tilanus HW. Changing patterns in the treatment of carcinoma of the esophagus. Scand J Gastroenterol Suppl. 1995. 212:38-42. [View Abstract]
  12. Jankowski JA, Wright NA, Meltzer SJ, et al. Molecular evolution of the metaplasia-dysplasia-adenocarcinoma sequence in the esophagus. Am J Pathol. 1999 Apr. 154(4):965-73. [View Abstract]
  13. Koppert LB, Wijnhoven BP, van Dekken H, et al. The molecular biology of esophageal adenocarcinoma. J Surg Oncol. 2005 Dec 1. 92(3):169-90. [View Abstract]
  14. Casson AG, Manolopoulos B, Troster M, et al. Clinical implications of p53 gene mutation in the progression of Barrett's epithelium to invasive esophageal cancer. Am J Surg. 1994 Jan. 167(1):52-7. [View Abstract]
  15. Nieman KM, Romero IL, Van Houten B, Lengyel E. Adipose tissue and adipocytes support tumorigenesis and metastasis. Biochim Biophys Acta. 2013 Oct. 1831 (10):1533-41. [View Abstract]
  16. Wang KK, Sampliner RE, Practice Parameters Committee of the American College of Gastroenterology. Updated guidelines 2008 for the diagnosis, surveillance and therapy of Barrett's esophagus. Am J Gastroenterol. 2008 Mar. 103 (3):788-97. [View Abstract]
  17. Lagergren J, Bergstrom R, Lindgren A, et al. Symptomatic gastroesophageal reflux as a risk factor for esophageal adenocarcinoma. N Engl J Med. 1999 Mar 18. 340(11):825-31. [View Abstract]
  18. Steevens J, Schouten LJ, Goldbohm RA, van den Brandt PA. Vegetables and fruits consumption and risk of esophageal and gastric cancer subtypes in the Netherlands Cohort Study. Int J Cancer. 2011 Dec 1. 129(11):2681-93. [View Abstract]
  19. Steevens J, Schouten LJ, Goldbohm RA, van den Brandt PA. Alcohol consumption, cigarette smoking and risk of subtypes of oesophageal and gastric cancer: a prospective cohort study. Gut. 2010 Jan. 59(1):39-48. [View Abstract]
  20. Lu SH, Montesano R, Zhang MS, et al. Relevance of N-nitrosamines to esophageal cancer in China. J Cell Physiol Suppl 1986; 4:51.
  21. Chu FS, Li GY. Simultaneous occurrence of fumonisin B1 and other mycotoxins in moldy corn collected from the People's Republic of China in regions with high incidences of esophageal cancer. Appl Environ Microbiol 1994; 60:847.
  22. Trivedy C, Baldwin D, Warnakulasuriya S, et al. Copper content in Areca catechu (betel nut) products and oral submucous fibrosis. Lancet 1997; 349:1447.
  23. Cross AJ, Freedman ND, Ren J, Ward MH, Hollenbeck AR, Schatzkin A, et al. Meat consumption and risk of esophageal and gastric cancer in a large prospective study. Am J Gastroenterol. 2011 Mar. 106 (3):432-42. [View Abstract]
  24. Mark SD, Qiao YL, Dawsey SM, et al. Prospective study of serum selenium levels and incident esophageal and gastric cancers. J Natl Cancer Inst 2000; 92:1753.
  25. Li B, Taylor PR, Li JY, et al. Linxian nutrition intervention trials. Design, methods, participant characteristics, and compliance. Ann Epidemiol 1993; 3:577.
  26. Abnet CC, Lai B, Qiao YL, et al. Zinc concentration in esophageal biopsy specimens measured by x-ray fluorescence and esophageal cancer risk. J Natl Cancer Inst 2005; 97:301.
  27. Fong LY, Sivak A, Newberne PM. Zinc deficiency and methylbenzylnitrosamine-induced esophageal cancer in rats. J Natl Cancer Inst 1978; 61:145.
  28. Fong LY, Zhang L, Jiang Y, Farber JL. Dietary zinc modulation of COX-2 expression and lingual and esophageal carcinogenesis in rats. J Natl Cancer Inst 2005; 97:40.
  29. Larsson SC, Giovannucci E, Wolk A. Folate intake, MTHFR polymorphisms, and risk of esophageal, gastric, and pancreatic cancer: a meta-analysis. Gastroenterology 2006; 131:1271.
  30. Liu J, Wang J, Leng Y, Lv C. Intake of fruit and vegetables and risk of esophageal squamous cell carcinoma: a meta-analysis of observational studies. Int J Cancer. 2013 Jul 15. 133 (2):473-85. [View Abstract]
  31. Wysowski DK. Reports of esophageal cancer with oral bisphosphonate use. N Engl J Med 2009; 360:89.
  32. Yu C, Tang H, Guo Y, Bian Z, Yang L, Chen Y, et al. Effect of Hot Tea Consumption and Its Interactions With Alcohol and Tobacco Use on the Risk for Esophageal Cancer: A Population-Based Cohort Study. Ann Intern Med. 2018 Feb 6. [View Abstract]
  33. Jenkins K. Hot Tea and Increased Risk for Esophageal Cancer. Medscape Medical News. Available at https://www.medscape.com/viewarticle/892285. February 5, 2018; Accessed: February 7, 2018.
  34. Wu C, Hu Z, He Z, et al. Genome-wide association study identifies three new susceptibility loci for esophageal squamous-cell carcinoma in Chinese populations. Nat Genet. 2011 Jun 5. 43(7):679-84. [View Abstract]
  35. Cardwell CR, Abnet CC, Cantwell MM, Murray LJ. Exposure to oral bisphosphonates and risk of esophageal cancer. JAMA. 2010 Aug 11. 304(6):657-63. [View Abstract]
  36. Sitas F, Egger S, Urban MI, et al. InterSCOPE study: associations between esophageal squamous cell carcinoma and human papillomavirus serological markers. J Natl Cancer Inst. 2012 Jan 18. 104(2):147-58. [View Abstract]
  37. Ellis A, Field JK, Field EA, Friedmann PS, Fryer A, Howard P, Leigh IM, Risk J, Shaw JM, Whittaker J. Tylosis associated with carcinoma of the oesophagus and oral leukoplakia in a large Liverpool family—a review of six generations. European Journal of Cancer Part B: Oral Oncology. 1994 Dec 31;30(2):102-12.
  38. Stevens HP, Kelsell DP, Bryant SP, et al. Linkage of an American pedigree with palmoplantar keratoderma and malignancy (palmoplantar ectodermal dysplasia type III) to 17q24. Literature survey and proposed updated classification of the keratodermas. Arch Dermatol 1996; 132:640.
  39. Iwaya T, Maesawa C, Ogasawara S, Tamura G. Tylosis esophageal cancer locus on chromosome 17q25.1 is commonly deleted in sporadic human esophageal cancer. Gastroenterology 1998; 114:1206.
  40. Lindor NM, Greene MH. The concise handbook of family cancer syndromes. Journal of the National Cancer Institute. 1998 Jul 15;90(14):1039-71.
  41. Morson BC, Belcher JR. Adenocarcinoma of the oesophagus and ectopic gastric mucosa. Br J Cancer. 1952 Jun. 6(2):127-30. [View Abstract]
  42. Naef AP, Savary M, Ozzello L. Columnar-lined lower esophagus: an acquired lesion with malignant predisposition. Report on 140 cases of Barrett's esophagus with 12 adenocarcinomas. J Thorac Cardiovasc Surg. 1975 Nov. 70(5):826-35. [View Abstract]
  43. Bhat S, Coleman HG, Yousef F, et al. Risk of malignant progression in Barrett's esophagus patients: results from a large population-based study. J Natl Cancer Inst. 2011 Jul 6. 103(13):1049-57. [View Abstract]
  44. Hvid-Jensen F, Pedersen L, Drewes AM, et al. Incidence of adenocarcinoma among patients with Barrett's esophagus. N Engl J Med. 2011 Oct 13. 365(15):1375-83. [View Abstract]
  45. Cook MB, Kamangar F, Whiteman DC, et al. Cigarette smoking and adenocarcinomas of the esophagus and esophagogastric junction: a pooled analysis from the international BEACON consortium. J Natl Cancer Inst 2010; 102:1344.
  46. Lanuti M, Liu G, Goodwin JM, Zhai R, Fuchs BC, Asomaning K, et al. A functional epidermal growth factor (EGF) polymorphism, EGF serum levels, and esophageal adenocarcinoma risk and outcome. Clin Cancer Res. 2008 May 15. 14 (10):3216-22. [View Abstract]
  47. Freedman J, Ye W, Näslund E, Lagergren J. Association between cholecystectomy and adenocarcinoma of the esophagus. Gastroenterology 2001; 121:548.
  48. Lauby-Secretan B, Scoccianti C, Loomis D, Grosse Y, Bianchini F, Straif K, et al. Body Fatness and Cancer--Viewpoint of the IARC Working Group. N Engl J Med. 2016 Aug 25. 375 (8):794-8. [View Abstract]
  49. Thrift AP, Shaheen NJ, Gammon MD, Bernstein L, Reid BJ, Onstad L, et al. Obesity and risk of esophageal adenocarcinoma and Barrett's esophagus: a Mendelian randomization study. J Natl Cancer Inst. 2014 Nov. 106 (11):[View Abstract]
  50. Turati F, Tramacere I, La Vecchia C, Negri E. A meta-analysis of body mass index and esophageal and gastric cardia adenocarcinoma. Ann Oncol 2013; 24:609.
  51. Lagergren J, Bergström R, Nyrén O. Association between body mass and adenocarcinoma of the esophagus and gastric cardia. Ann Intern Med 1999; 130:883.
  52. Cook MB, Greenwood DC, Hardie LJ, et al. A systematic review and meta-analysis of the risk of increasing adiposity on Barrett's esophagus. Am J Gastroenterol 2008; 103:292.
  53. Drahos J, Ricker W, Pfeiffer RM, Cook MB. Metabolic syndrome and risk of esophageal adenocarcinoma in elderly patients in the United States: An analysis of SEER-Medicare data. Cancer 2017; 123:657.
  54. 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 14, 2019.
  55. Cancer Stat Facts: Esophageal Cancer. National Cancer Institute. Available at https://seer.cancer.gov/statfacts/html/esoph.html. Accessed: August 5, 2019.
  56. Kirby TJ, Rice TW. The epidemiology of esophageal carcinoma. The changing face of a disease. Chest Surg Clin N Am. 1994 May. 4(2):217-25. [View Abstract]
  57. Raman R, Deorah S, McDowell BD, Abu Hejleh T, Lynch CF, Gupta A. Changing incidence of esophageal cancer among white women: analysis of SEER data (1992-2010). Contemp Oncol (Pozn). 2015. 19 (4):338-40. [View Abstract]
  58. Pohl H, Sirovich B, Welch HG. Esophageal adenocarcinoma incidence: are we reaching the peak?. Cancer Epidemiol Biomarkers Prev. 2010 Jun. 19 (6):1468-70. [View Abstract]
  59. Bray F, Ferlay J, Soerjomataram I, Siegel RL, Torre LA, Jemal A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2018 Nov. 68 (6):394-424. [View Abstract]
  60. Howlader N, Noone AM, Krapcho M, Miller D, Brest A, Yu M, et al. SEER Cancer Statistics Review (CSR) 1975-2016. National Cancer Institute: Surveillance, Epidemiology, and End Results Program. Available at https://seer.cancer.gov/csr/1975_2016/. April 15, 2019; Accessed: August 5, 2019.
  61. Orringer MB, Marshall B, Iannettoni MD. Transhiatal esophagectomy: clinical experience and refinements. Ann Surg. 1999 Sep. 230(3):392-400; discussion 400-3. [View Abstract]
  62. Chang AC, Ji H, Birkmeyer NJ, et al. Outcomes after transhiatal and transthoracic esophagectomy for cancer. Ann Thorac Surg. 2008 Feb. 85(2):424-9. [View Abstract]
  63. Chu KM, Law SY, Fok M, et al. A prospective randomized comparison of transhiatal and transthoracic resection for lower-third esophageal carcinoma. Am J Surg. 1997 Sep. 174(3):320-4. [View Abstract]
  64. Suzuki A, Xiao L, Hayashi Y, et al. Prognostic significance of baseline positron emission tomography and importance of clinical complete response in patients with esophageal or gastroesophageal junction cancer treated with definitive chemoradiotherapy. Cancer. 2011 Nov 1. 117(21):4823-33. [View Abstract]
  65. Gillies RS, Middleton MR, Han C, et al. Role of positron emission tomography–computed tomography in predicting survival after neoadjuvant chemotherapy and surgery for oesophageal adenocarcinoma. Brit J Surg. Feb 2012. 99(2):239-45.
  66. Prins MJ, Ruurda JP, van Diest PJ, van Hillegersberg R, Ten Kate FJ. The significance of the HER-2 status in esophageal adenocarcinoma for survival: an immunohistochemical and an in situ hybridization study. Ann Oncol. 2013 Jan 18. [View Abstract]
  67. [Guideline] Varghese TK Jr, Hofstetter WL, Rizk NP, Low DE, Darling GE, Watson TJ, et al. The society of thoracic surgeons guidelines on the diagnosis and staging of patients with esophageal cancer. Ann Thorac Surg. 2013 Jul. 96(1):346-56. [View Abstract]
  68. O'Donovan PB. The radiographic evaluation of the patient with esophageal carcinoma. Chest Surg Clin N Am. 1994 May. 4(2):241-56. [View Abstract]
  69. Dittler HJ, Siewert JR. Role of endoscopic ultrasonography in esophageal carcinoma. Endoscopy. 1993 Feb. 25(2):156-61. [View Abstract]
  70. Barbour AP, Rizk NP, Gerdes H, Bains MS, Rusch VW, Brennan MF, Coit DG. Endoscopic ultrasound predicts outcomes for patients with adenocarcinoma of the gastroesophageal junction. Journal of the American College of Surgeons. 2007 Oct 31;205(4):593-601.
  71. Keswani RN, Early DS, Edmundowicz SA, Meyers BF, Sharma A, Govindan R, Chen J, Kohlmeier C, Azar RR. Routine positron emission tomography does not alter nodal staging in patients undergoing EUS-guided FNA for esophageal cancer. Gastrointestinal endoscopy. 2009 Jun 30;69(7):1210-7.
  72. Vazquez-Sequeiros E, Norton ID, Clain JE, Wang KK, Affi A, Allen M, Deschamps C, Miller D, Salomao D, Wiersema MJ. Impact of EUS-guided fine-needle aspiration on lymph node staging in patients with esophageal carcinoma. Gastrointestinal endoscopy. 2001 Jun 30;53(7):751-7.
  73. Vazquez-Sequeiros E, Wiersema MJ, Clain JE, Norton ID, Levy MJ, Romero Y, Salomao D, Dierkhising R, Zinsmeister AR. Impact of lymph node staging on therapy of esophageal carcinoma. Gastroenterology. 2003 Dec 31;125(6):1626-35.
  74. Rice TW, Rusch VW, Ishwaran H, Blackstone EH. Cancer of the esophagus and esophagogastric junction: data-driven staging for the seventh edition of the American Joint Committee on Cancer/International Union Against Cancer Cancer Staging Manuals. Cancer. 2010 Aug 15. 116(16):3763-73. [View Abstract]
  75. Hofstetter W, Correa AM, Bekele N, et al. Proposed modification of nodal status in AJCC esophageal cancer staging system. Ann Thorac Surg. 2007 Aug. 84(2):365-73; discussion 374-5. [View Abstract]
  76. American Joint Committee on Cancer. Esophagus and Esophagogastric Junction. Amin MB, Edge S, Greene F, Byrd DR, Brookland RK, et al, eds. AJCC Staging Manual. 8th edition. New York, NY: Springer; 2017. 185-202.
  77. Rice TW, Rusch VW, Apperson-Hansen C, Allen MS, Chen LQ, Hunter JG, et al. Worldwide esophageal cancer collaboration. Dis Esophagus. 2009. 22 (1):1-8. [View Abstract]
  78. Herbella FA, Del Grande JC, Colleoni R. Japanese Society for Disease of the Esophagus. Anatomical analysis of the mediastinal lymph nodes of normal Brazilian subjects according to the classification of the Japanese Society for Diseases of the Esophagus. Surg Today. 2003. 33(4):249-53. [View Abstract]
  79. Gaur P, Sepesi B, Hofstetter WL, Correa AM, Bhutani MS, Watson TJ, et al. Endoscopic esophageal tumor length: a prognostic factor for patients with esophageal cancer. Cancer. 2011 Jan 1. 117 (1):63-9. [View Abstract]
  80. Maish MS, DeMeester SR. Endoscopic mucosal resection as a staging technique to determine the depth of invasion of esophageal adenocarcinoma. Ann Thorac Surg. 2004 Nov. 78 (5):1777-82. [View Abstract]
  81. Thosani N, Singh H, Kapadia A, Ochi N, Lee JH, Ajani J, et al. Diagnostic accuracy of EUS in differentiating mucosal versus submucosal invasion of superficial esophageal cancers: a systematic review and meta-analysis. Gastrointest Endosc. 2012 Feb. 75 (2):242-53. [View Abstract]
  82. Pennathur A, Farkas A, Krasinskas AM, Ferson PF, Gooding WE, Gibson MK, et al. Esophagectomy for T1 esophageal cancer: outcomes in 100 patients and implications for endoscopic therapy. Ann Thorac Surg. 2009 Apr. 87 (4):1048-54; discussion 1054-5. [View Abstract]
  83. [Guideline] NCCN Clinical Practice Guidelines in Oncology: Esophageal and Esophagogastric Junction Cancers. National Comprehensive Cancer Network. Available at http://www.nccn.org/professionals/physician_gls/pdf/esophageal.pdf. Version 2.2019 — May 29, 2019; Accessed: August 5, 2019.
  84. Bang YJ, Van Cutsem E, Feyereislova A, Chung HC, Shen L, Sawaki A, et al. Trastuzumab in combination with chemotherapy versus chemotherapy alone for treatment of HER2-positive advanced gastric or gastro-oesophageal junction cancer (ToGA): a phase 3, open-label, randomised controlled trial. Lancet. 2010 Aug 28. 376 (9742):687-97. [View Abstract]
  85. Schomas DA, Quevedo JF, Donahue JM, Nichols FC 3rd, Romero Y, Miller RC. The prognostic importance of pathologically involved celiac node metastases in node-positive patients with carcinoma of the distal esophagus or gastroesophageal junction: a surgical series from the Mayo Clinic. Dis Esophagus. 2010 Apr. 23(3):232-9. [View Abstract]
  86. Gluch L, Smith RC, Bambach CP, et al. Comparison of outcomes following transhiatal or Ivor Lewis esophagectomy for esophageal carcinoma. World J Surg. 1999 Mar. 23(3):271-5; discussion 275-6. [View Abstract]
  87. Goldminc M, Maddern G, Le Prise E, et al. Oesophagectomy by a transhiatal approach or thoracotomy: a prospective randomized trial. Br J Surg. 1993 Mar. 80(3):367-70. [View Abstract]
  88. Hankins JR, Attar S, Coughlin TR Jr, et al. Carcinoma of the esophagus: a comparison of the results of transhiatal versus transthoracic resection. Ann Thorac Surg. 1989 May. 47(5):700-5. [View Abstract]
  89. Stiles BM, Altorki NK. Traditional techniques of esophagectomy. Surg Clin North Am. 2012 Oct. 92 (5):1249-63. [View Abstract]
  90. Dai Y, Chopra SS, Kneif S, Hunerbein M. Management of esophageal anastomotic leaks, perforations, and fistulae with self-expanding plastic stents. J Thorac Cardiovasc Surg. 2011 May. 141(5):1213-7. [View Abstract]
  91. Martin LW, Swisher SG, Hofstetter W, et al. Intrathoracic leaks following esophagectomy are no longer associated with increased mortality. Ann Surg. 2005 Sep. 242(3):392-9; discussion 399-402. [View Abstract]
  92. Patti MG, Corvera CU, Glasgow RE, et al. A hospital's annual rate of esophagectomy influences the operative mortality rate. J Gastrointest Surg. 1998 Mar-Apr. 2(2):186-92. [View Abstract]
  93. Nguyen NT, Roberts P, Follette DM, et al. Thoracoscopic and laparoscopic esophagectomy for benign and malignant disease: lessons learned from 46 consecutive procedures. J Am Coll Surg. 2003 Dec. 197(6):902-13. [View Abstract]
  94. Luketich JD, Alvelo-Rivera M, Buenaventura PO, Christie NA, McCaughan JS, Litle VR, et al. Minimally invasive esophagectomy: outcomes in 222 patients. Ann Surg. 2003 Oct. 238 (4):486-94; discussion 494-5. [View Abstract]
  95. Mariette C, et al, for the Fédération de Recherche en Chirurgie (FRENCH) and French Eso-Gastric Tumors (FREGAT) Working Group. Hybrid Minimally Invasive Esophagectomy for Esophageal Cancer. N Engl J Med. 2019 Jan 10. 380 (2):152-162. [View Abstract]
  96. Castellino AM. Should Minimally Invasive Esophagectomy Now Be Standard?. Medscape Medical News. Available at https://www.medscape.com/viewarticle/907435. January 09, 2019; Accessed: January 14, 2019.
  97. Uenosono Y, Arigami T, Yanagita S, et al. Sentinel node navigation surgery is acceptable for clinical T1 and N0 esophageal cancer. Ann Surg Oncol. 2011 Jul. 18(7):2003-9. [View Abstract]
  98. Ngamruengphong S, Wolfsen HC, Wallace MB. Survival of Patients With Superficial Esophageal Adenocarcinoma After Endoscopic Treatment vs Surgery. Clin Gastroenterol Hepatol. 2013 Nov. 11(11):1424-1429.e2. [View Abstract]
  99. Brooks M. Endoscopic Therapy a Good Option in Early Esophageal Cancer. Medscape [serial online]. Available at http://www.medscape.com/viewarticle/814817. Accessed: November 25, 2013.
  100. Novak B. Salvage endoscopy viable for some esophageal cancers. Medscape Medical News. January 29, 2014.
  101. Cunningham D, Allum WH, Stenning SP, et al. Perioperative chemotherapy versus surgery alone for resectable gastroesophageal cancer. N Engl J Med. 2006 Jul 6. 355(1):11-20. [View Abstract]
  102. Allum WH, Stenning SP, Bancewicz J, Clark PI, Langley RE. Long-term results of a randomized trial of surgery with or without preoperative chemotherapy in esophageal cancer. J Clin Oncol. 2009 Oct 20. 27(30):5062-7. [View Abstract]
  103. van Hagen P, Hulshof MC, van Lanschot JJ, Steyerberg EW, van Berge Henegouwen MI, Wijnhoven BP, et al. Preoperative chemoradiotherapy for esophageal or junctional cancer. N Engl J Med. 2012 May 31. 366(22):2074-84. [View Abstract]
  104. Brooks M. Preop chemoradiation benefits confirmed in esophageal cancer. Medscape Medical News. January 16, 2014.
  105. Oppedijk V, van der Gaast A, van Lanschot JJ, van Hagen P, van Os R, van Rij CM, et al. Patterns of Recurrence After Surgery Alone Versus Preoperative Chemoradiotherapy and Surgery in the CROSS Trials. J Clin Oncol. 2014 Jan 13. [View Abstract]
  106. Mamon HJ, Tepper JE. Combination Chemoradiation Therapy: The Whole Is More Than the Sum of the Parts. J Clin Oncol. 2014 Jan 13. [View Abstract]
  107. Franko J, Voynov G, Goldman CD. Esophagectomy Timing After Neoadjuvant Therapy for Distal Esophageal Adenocarcinoma. Ann Thorac Surg. 2016 Mar. 101 (3):1123-30. [View Abstract]
  108. Rohatgi PR, Swisher SG, Correa AM, Wu TT, Liao Z, Komaki R, et al. Failure patterns correlate with the proportion of residual carcinoma after preoperative chemoradiotherapy for carcinoma of the esophagus. Cancer. 2005 Oct 1. 104(7):1349-55. [View Abstract]
  109. Rao S, Welsh L, Cunningham D, et al. Correlation of overall survival with gene expression profiles in a prospective study of resectable esophageal cancer. Clin Colorectal Cancer. 2011 Mar 1. 10(1):48-56. [View Abstract]
  110. Alexander BM, Wang XZ, Niemierko A, Weaver DT, Mak RH, Roof KS, et al. DNA repair biomarkers predict response to neoadjuvant chemoradiotherapy in esophageal cancer. Int J Radiat Oncol Biol Phys. 2012 May 1. 83(1):164-71. [View Abstract]
  111. Fietkau R, Lewitzki V, Kuhnt T, et al. A disease-specific enteral nutrition formula improves nutritional status and functional performance in patients with head and neck and esophageal cancer undergoing chemoradiotherapy: results of a randomized, controlled, multicenter trial. Cancer. 2013 Jun 13. [View Abstract]
  112. Tailored Enteral Formula During Chemoradiotherapy Curbs Body Mass Loss. Medscape. Jun 21 2013.
  113. Zosia Chustecka. First Approval for Immunotherapy for Esophageal Cancer. Medscape Medical News. Available at https://www.medscape.com/viewarticle/916257. Jul 31, 2019; Accessed: August 1, 2019.
  114. Kojima T, et al. Pembrolizumab versus chemotherapy as second-line therapy for advanced esophageal cancer: Phase III KEYNOTE-181 study. Presented at the Gastrointestinal Cancers Symposium 2019. Journal of Clinical Oncology. February 1, 2019. 37 (4):2-2.
  115. Shah MA, Kojima T, Hochhauser D, Enzinger P, Raimbourg J, Hollebecque A, et al. Efficacy and Safety of Pembrolizumab for Heavily Pretreated Patients With Advanced, Metastatic Adenocarcinoma or Squamous Cell Carcinoma of the Esophagus: The Phase 2 KEYNOTE-180 Study. JAMA Oncol. 2018 Dec 20. [View Abstract]
  116. Shitara K, Doi T, Dvorkin M, Mansoor W, Arkenau HT, Prokharau A, et al. Trifluridine/tipiracil versus placebo in patients with heavily pretreated metastatic gastric cancer (TAGS): a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol. 2018 Nov. 19 (11):1437-1448. [View Abstract]
  117. Homs MY, Steyerberg EW, Eijkenboom WM, Tilanus HW, Stalpers LJ, Bartelsman JF, van Lanschot JJ, Wijrdeman HK, Mulder CJ, Reinders JG, Boot H. Single-dose brachytherapy versus metal stent placement for the palliation of dysphagia from oesophageal cancer: multicentre randomised trial. The Lancet. 2004 Oct 29;364(9444):1497-504.
  118. Verschuur EM, Steyerberg EW, Kuipers EJ, Siersema PD. Effect of stent size on complications and recurrent dysphagia in patients with esophageal or gastric cardia cancer. Gastrointestinal endoscopy. 2007 Apr 30;65(4):592-601.
  119. Lightdale CJ, Heier SK, Marcon NE, McCaughan JS, Gerdes H, Overholt BF, Sivak MV, Stiegmann GV, Nava HR. Photodynamic therapy with porfimer sodium versus thermal ablation therapy with Nd: YAG laser for palliation of esophageal cancer: a multicenter randomized trial. Gastrointestinal endoscopy. 1995 Dec 31;42(6):507-12.
  120. Vakil N, Morris AI, Marcon N, Segalin A, Peracchia A, Bethge N, Zuccaro G, Bosco JJ, Jones WF. A prospective, randomized, controlled trial of covered expandable metal stents in the palliation of malignant esophageal obstruction at the gastroesophageal junction. The American journal of gastroenterology. 2001 Jun 1;96(6):1791-6.
  121. Shin JH, Song HY, Kim JH, Kim SB, Lee GH, Park SI, Han YM, Kang W. Comparison of temporary and permanent stent placement with concurrent radiation therapy in patients with esophageal carcinoma. Journal of vascular and interventional radiology. 2005 Jan 31;16(1):67-74.
  122. Ross WA, Alkassab F, Lynch PM, Ayers GD, Ajani J, Lee JH, Bismar M. Evolving role of self-expanding metal stents in the treatment of malignant dysphagia and fistulas. Gastrointestinal endoscopy. 2007 Jan 31;65(1):70-6.
  123. [Guideline] Wong RK, Malthaner RA, Zuraw L, Rumble RB,. Combined modality radiotherapy and chemotherapy in nonsurgical management of localized carcinoma of the esophagus: a practice guideline. Int J Radiat Oncol Biol Phys. 2003 Mar 15. 55(4):930-42. [View Abstract]
  124. Herskovic A, Martz K, al-Sarraf M, et al. Combined chemotherapy and radiotherapy compared with radiotherapy alone in patients with cancer of the esophagus. N Engl J Med. 1992 Jun 11. 326(24):1593-8. [View Abstract]
  125. Folkert MR, Cohen GN, Wu AJ, et al. Endoluminal high-dose-rate brachytherapy for early stage and recurrent esophageal cancer in medically inoperable patients. Brachytherapy. 2013 Feb 20. [View Abstract]
  126. Chustecka Z. Second-line docetaxel affirmed for esophagogastric cancer. Medscape Medical News. Jan 23, 2013. Available at http://www.medscape.com/viewarticle/778031. Accessed: January 28, 2013.
  127. Ford H, Marshall A, Wadsley J, et al. Cougar-02: a randomized phase III study of docetaxel versus active symptom control in advanced esophagogastric adenocarcinoma. J Clin Oncol. 2012. 30:(suppl 34 abstr LBA4).
  128. Homs MY, v d Gaast A, Siersema PD, et al. Chemotherapy for metastatic carcinoma of the esophagus and gastro-esophageal junction. Cochrane Database Syst Rev. 2006 Oct 18. CD004063. [View Abstract]
  129. Bethesda, MD: National Cancer Institute. National Cancer Institute: PDQ® Esophageal Cancer Prevention. Available at http://www.cancer.gov/cancertopics/pdq/prevention/esophageal/HealthProfessional. February 11, 2016; Accessed: August 7, 2016.
  130. Romero YV, Cameron AJ, Locke G3, Schaid DJ, Slezak JM, Branch CD, Melton LJ. Familial aggregation of gastroesophageal reflux in patients with Barrett's esophagus and esophageal adenocarcinoma. Gastroenterology. 1997 Nov 1;113(5):1449-56.
  131. Verbeek RE, Spittuler LF, Peute A, van Oijen MG, Fiebo J, Vermeijden JR, Oberndorff A, van Baal JW, Siersema PD. Familial clustering of Barrett's esophagus and esophageal adenocarcinoma in a European cohort. Clinical Gastroenterology and Hepatology. 2014 Oct 31;12(10):1656-63.
  132. Ellis NA, German J. Molecular genetics of Bloom's syndrome. Human molecular genetics. 1996 Sep 1;5(Supplement 1):1457-63.
  133. de Winter JP, Joenje H. The genetic and molecular basis of Fanconi anemia. Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis. 2009 Jul 31;668(1):11-9.
  134. Rosenberg PS, Alter BP, Ebell W. Cancer risks in Fanconi anemia: findings from the German Fanconi Anemia Registry. Haematologica. 2008 Apr 1;93(4):511-7.
  135. van Zeeburg HJ, Snijders PJ, Wu T, Gluckman E, Soulier J, Surralles J, Castella M, van der Wal JE, Wennerberg J, Califano J, Velleuer E. Clinical and molecular characteristics of squamous cell carcinomas from Fanconi anemia patients. Journal of the National Cancer Institute. 2008 Nov 19;100(22):1649-53.
  136. Oostra AB, Nieuwint AW, Joenje H, De Winter JP. Diagnosis of fanconi anemia: chromosomal breakage analysis. Anemia. 2012 May 24;2012.
  137. [Guideline] ASGE Standards of Practice Committee, Evans JA, Early DS, Fukami N, et al, Standards of Practice Committee of the American Society for Gastrointestinal Endoscopy. The role of endoscopy in Barrett's esophagus and other premalignant conditions of the esophagus. Gastrointest Endosc. 2012 Dec. 76 (6):1087-94. [View Abstract]
  138. [Guideline] American Gastroenterological Association, Spechler SJ, Sharma P, Souza RF, Inadomi JM, Shaheen NJ. American Gastroenterological Association medical position statement on the management of Barrett's esophagus. Gastroenterology. 2011 Mar. 140 (3):1084-91. [View Abstract]
  139. [Guideline] Shaheen NJ, Falk GW, Iyer PG, Gerson LB, American College of Gastroenterology. ACG Clinical Guideline: Diagnosis and Management of Barrett's Esophagus. Am J Gastroenterol. 2016 Jan. 111 (1):30-50; quiz 51. [View Abstract]
  140. [Guideline] Shaheen NJ, Weinberg DS, Denberg TD, Chou R, Qaseem A, Shekelle P, et al. Upper endoscopy for gastroesophageal reflux disease: best practice advice from the clinical guidelines committee of the American College of Physicians. Ann Intern Med. 2012 Dec 4. 157 (11):808-16. [View Abstract]
  141. [Guideline] ASGE Standards of Practice Committee, Muthusamy VR, Lightdale JR, Acosta RD, Chandrasekhara V, Chathadi KV, et al. The role of endoscopy in the management of GERD. Gastrointest Endosc. 2015. 81 (6):1305-10. [View Abstract]
  142. Chennat J, Waxman I. Endoscopic treatment of Barrett’s esophagus: From metaplasia to intramucosal carcinoma. World J Gastroenterol. 2010 Aug 14;16(30):3780-5.
  143. Nealis TB, Washington K, Keswani RN. Endoscopic therapy of esophageal premalignancy and early malignancy. Journal of the National Comprehensive Cancer Network. 2011 Aug 1;9(8):890-9.
  144. Shaheen NJ, Overholt BF, Sampliner RE, Wolfsen HC, Wang KK, Fleischer DE, et al. Durability of radiofrequency ablation in Barrett's esophagus with dysplasia. Gastroenterology. 2011 Aug. 141(2):460-8. [View Abstract]
  145. Oppedijk V, van der Gaast A, van Lanschot JJ, van Hagen P, van Os R, van Rij CM, van der Sangen MJ, Beukema JC, Rütten H, Spruit PH, Reinders JG. Patterns of recurrence after surgery alone versus preoperative chemoradiotherapy and surgery in the CROSS trials. Journal of Clinical Oncology. 2014 Jan 13;32(5):385-91.
  146. Lou F, Sima CS, Adusumilli PS, Bains MS, Sarkaria IS, Rusch VW, Rizk NP. Esophageal cancer recurrence patterns and implications for surveillance. Journal of Thoracic Oncology. 2013 Dec 31;8(12):1558-62.
  147. Sudo K, Xiao L, Wadhwa R, Shiozaki H, Elimova E, Taketa T, Blum MA, Lee JH, Bhutani MS, Weston B, Ross WA. Importance of surveillance and success of salvage strategies after definitive chemoradiation in patients with esophageal cancer. Journal of Clinical Oncology. 2014 Sep 15;32(30):3400-5.
  148. Dorth JA, Pura JA, Palta M, Willett CG, Uronis HE, D'amico TA, Czito BG. Patterns of recurrence after trimodality therapy for esophageal cancer. Cancer. 2014 Jul 15;120(14):2099-105.
  149. [Guideline] ASGE Standards of Practice Committee, Evans JA, Early DS, Chandraskhara V, Chathadi KV, Fanelli RD, et al. The role of endoscopy in the assessment and treatment of esophageal cancer. Gastrointest Endosc. 2013 Mar. 77 (3):328-34. [View Abstract]
  150. [Guideline] Bartley AN, Washington MK, Ismaila N, Ajani JA. HER2 Testing and Clinical Decision Making in Gastroesophageal Adenocarcinoma: Guideline Summary From the College of American Pathologists, American Society for Clinical Pathology, and American Society of Clinical Oncology. J Oncol Pract. 2016 Nov 29. JOP2016018929. [View Abstract]
  151. Stahl M, Walz MK, Stuschke M, Lehmann N, Meyer HJ, Riera-Knorrenschild J, Langer P, Engenhart-Cabillic R, Bitzer M, Königsrainer A, Budach W. Phase III comparison of preoperative chemotherapy compared with chemoradiotherapy in patients with locally advanced adenocarcinoma of the esophagogastric junction. Journal of Clinical Oncology. 2009 Jan 12;27(6):851-6.
  152. Minsky BD, Pajak TF, Ginsberg RJ, Pisansky TM, Martenson J, Komaki R, Okawara G, Rosenthal SA, Kelsen DP. INT 0123 (Radiation Therapy Oncology Group 94-05) phase III trial of combined-modality therapy for esophageal cancer: high-dose versus standard-dose radiation therapy. Journal of clinical oncology. 2002 Mar 1;20(5):1167-74.
  153. Cooper JS, Guo MD, Herskovic A, Macdonald JS, Martenson Jr JA, Al-Sarraf M, Byhardt R, Russell AH, Beitler JJ, Spencer S, Asbell SO. Chemoradiotherapy of locally advanced esophageal cancer: long-term follow-up of a prospective randomized trial (RTOG 85-01). Jama. 1999 May 5;281(17):1623-7.
  154. Meerten EV, van Rij C, Tesselaar ME, Neelis K, Richel D, Hulshof M, Gaast AV. Definitive concurrent chemoradiation (CRT) with weekly paclitaxel and carboplatin for patients (pts) with irresectable esophageal cancer: A phase II study. Journal of Clinical Oncology. 2010 May 20;28(15_suppl):e14508-.
  155. van Hagen P, Hulshof MC, van Lanschot JJ, et al. Preoperative chemoradiotherapy for esophageal or junctional cancer. N Engl J Med. 2012 May 31. 366 (22):2074-84. [View Abstract]
  156. Cunningham D, Allum WH, Stenning SP, Thompson JN, Van de Velde CJ, Nicolson M, et al. Perioperative chemotherapy versus surgery alone for resectable gastroesophageal cancer. N Engl J Med. 2006 Jul 6. 355 (1):11-20. [View Abstract]
  157. Cooper JS, Guo MD, Herskovic A, et al. Chemoradiotherapy of locally advanced esophageal cancer: long-term follow-up of a prospective randomized trial (RTOG 85-01). Radiation Therapy Oncology Group. JAMA. 1999 May 5. 281(17):1623-7. [View Abstract]
  158. Forastiere AA, Orringer MB, Perez-Tamayo C, et al. Preoperative chemoradiation followed by transhiatal esophagectomy for carcinoma of the esophagus: final report. J Clin Oncol. 1993 Jun. 11(6):1118-23. [View Abstract]
  159. Sjoquist KM, Burmeister BH, Smithers BM, Zalcberg JR, Simes RJ, Barbour A, et al. Survival after neoadjuvant chemotherapy or chemoradiotherapy for resectable oesophageal carcinoma: an updated meta-analysis. Lancet Oncol. 2011 Jul. 12(7):681-92. [View Abstract]
  160. Le Prise E, Etienne PL, Meunier B, et al. A randomized study of chemotherapy, radiation therapy, and surgery versus surgery for localized squamous cell carcinoma of the esophagus. Cancer. 1994 Apr 1. 73(7):1779-84. [View Abstract]
  161. Leichman L, Steiger Z, Seydel HG, et al. Preoperative chemotherapy and radiation therapy for patients with cancer of the esophagus: a potentially curative approach. J Clin Oncol. 1984 Feb. 2(2):75-9. [View Abstract]
  162. Nygaard K, Hagen S, Hansen HS, et al. Pre-operative radiotherapy prolongs survival in operable esophageal carcinoma: a randomized, multicenter study of pre-operative radiotherapy and chemotherapy. The second Scandinavian trial in esophageal cancer. World J Surg. 1992 Nov-Dec. 16(6):1104-9; discussion 1110. [View Abstract]
  163. Reynolds JV, Muldoon C, Hollywood D, Ravi N, Rowley S, O'Byrne K, et al. Long-term outcomes following neoadjuvant chemoradiotherapy for esophageal cancer. Ann Surg. 2007 May. 245(5):707-16. [View Abstract]
  164. Macdonald JS, Smalley SR, Benedetti J, Hundahl SA, Estes NC, Stemmermann GN, et al. Chemoradiotherapy after surgery compared with surgery alone for adenocarcinoma of the stomach or gastroesophageal junction. N Engl J Med. 2001 Sep 6. 345 (10):725-30. [View Abstract]
  165. Fuchs CS, Tomasek J, Yong CJ, Dumitru F, Passalacqua R, et al. Ramucirumab monotherapy for previously treated advanced gastric or gastro-oesophageal junction adenocarcinoma (REGARD): an international, randomised, multicentre, placebo-controlled, phase 3 trial. Lancet. 2014 Jan 4. 383 (9911):31-9. [View Abstract]
  166. Wilke H, Muro K, Van Cutsem E, Oh SC, Bodoky G, Shimada Y, et al. Ramucirumab plus paclitaxel versus placebo plus paclitaxel in patients with previously treated advanced gastric or gastro-oesophageal junction adenocarcinoma (RAINBOW): a double-blind, randomised phase 3 trial. Lancet Oncol. 2014 Oct. 15 (11):1224-35. [View Abstract]
  167. [Guideline] Little AG, Lerut AE, Harpole DH, Hofstetter WL, Mitchell JD, Altorki NK, et al. The Society of Thoracic Surgeons practice guidelines on the role of multimodality treatment for cancer of the esophagus and gastroesophageal junction. Ann Thorac Surg. 2014 Nov. 98 (5):1880-5. [View Abstract]
  168. Bartley AN, Washington MK, Colasacco C, Ventura CB, Ismaila N, Benson III AB, Carrato A, Gulley ML, Jain D, Kakar S, Mackay HJ. HER2 Testing and Clinical Decision Making in Gastroesophageal Adenocarcinoma: Guideline From the College of American Pathologists, American Society for Clinical Pathology, and the American Society of Clinical Oncology. Journal of Clinical Oncology. 2016 Nov 14:JCO-2016.
  169. Siegel RL, Miller KD, Jemal A. Cancer Statistics, 2017. CA Cancer J Clin. 2017 Jan. 67 (1):7-30. [View Abstract]
  170. Du C, Zhou Y, Huang K, Zhao G, Fu H, Shi Y. Defining a high-risk subgroup of pathological T2N0 gastric cancer by prognostic risk stratification for adjuvant therapy. J Gastrointest Surg. 2011 Dec. 15 (12):2153-8. [View Abstract]
  171. Kubo A, Corley DA. Body mass index and adenocarcinomas of the esophagus or gastric cardia: a systematic review and meta-analysis. Cancer Epidemiol Biomarkers Prev. 2006 May. 15 (5):872-8. [View Abstract]
  172. PDQ® Adult Treatment Editorial Board. PDQ Esophageal Cancer Treatment. National Cancer Institute. Available at http://www.cancer.gov/types/esophageal/hp/esophageal-treatment-pdq. February 4, 2016; Accessed: August 8, 2016.
  173. Knyrim K, Wagner HJ, Bethge N, Keymling M, Vakil N. A controlled trial of an expansile metal stent for palliation of esophageal obstruction due to inoperable cancer. N Engl J Med. 1993 Oct 28. 329(18):1302-7. [View Abstract]
  174. Conroy T, Galais MP, Raoul JL, et al. Definitive chemoradiotherapy with FOLFOX versus fluorouracil and cisplatin in patients with oesophageal cancer (PRODIGE5/ACCORD17): final results of a randomised, phase 2/3 trial. Lancet Oncol. 2014 Mar. 15(3):305-14. [View Abstract]
  175. Gebski V, Burmeister B, Smithers BM, et al. Survival benefits from neoadjuvant chemoradiotherapy or chemotherapy in oesophageal carcinoma: a meta-analysis. Lancet Oncol. 2007 Mar. 8(3):226-34. [View Abstract]
  176. Kelsen DP, Ginsberg R, Pajak TF, Sheahan DG, Gunderson L, Mortimer J, et al. Chemotherapy followed by surgery compared with surgery alone for localized esophageal cancer. N Engl J Med. 1998 Dec 31. 339(27):1979-84. [View Abstract]
  177. Scheer RV, Fakiris AJ, Johnstone PA. Quantifying the benefit of a pathologic complete response after neoadjuvant chemoradiotherapy in the treatment of esophageal cancer. Int J Radiat Oncol Biol Phys. 2011 Jul 15. 80(4):996-1001. [View Abstract]

Cascade of events that lead from gastroesophageal reflux disease to adenocarcinoma.

Palmoplantar keratoderma (Tylosis) of palms (A) and soles (B). Courtesy of The American Journal of Gastroenterology, Nature Publishing Group.

Cascade of events that lead from gastroesophageal reflux disease to adenocarcinoma.

Chest CT scan showing invasion of the trachea by esophageal cancer.

Barium swallow demonstrating stricture due to cancer.

Barium swallow demonstrating an endoluminal mass in the mid esophagus.

Diagram showing T1,T2 and T3 stages of esophageal cancer. Courtesy of Cancer Research UK and Wikimedia Commons.

Endoscopy demonstrating intraluminal esophageal cancer.

Micrograph of squamous cell carcinoma of the esophagus (H&E Stain). Courtesy of Wikimedia Commons.

Low magnification micrograph of an intramucosal esophageal adenocarcinoma (H&E stain). Endoscopic mucosal resection specimen. Courtesy of Wikimedia Commons.

Transhiatal esophagectomy in which (a) is the abdominal incision, (b) is the cervical incision, and (c) is the stomach stretching from abdomen to the neck.

Endoscopy demonstrating intraluminal esophageal cancer.

Cascade of events that lead from gastroesophageal reflux disease to adenocarcinoma.

Barium swallow demonstrating stricture due to cancer.

Barium swallow demonstrating an endoluminal mass in the mid esophagus.

Chest CT scan showing invasion of the trachea by esophageal cancer.

Transhiatal esophagectomy in which (a) is the abdominal incision, (b) is the cervical incision, and (c) is the stomach stretching from abdomen to the neck.

Five-year survival for esophageal cancer based on TNM stage.

H and E, high power, showing junction of benign glands in the lower right, Barrett's columnar cell metaplasia with a large goblet cell containing blue mucin in the lower center and adenocarcinoma on the left.

Macroscopic image of a resection of the gastroesophageal junction. On the right is non-neoplastic esophagus, consisting of tan, smooth mucosa. On the left is the non-neoplastic rugal folds of the stomach. In the center of the picture is an ulcer with a yellow-green fibrinous exudate surrounded by irregular, heaped-up margins with almost a cobblestone appearance. The latter represents mucosal adenocarcinoma with probably some Barrett's metaplasia in the background.

H and E, high power, demonstrating invasive esophageal squamous cell carcinoma. This carcinoma does not form glands and instead shows features of squamous differentiation, including keratinization and intercellular bridges.

Macroscopic image of an esophageal resection. A polypoid squamous cell carcinoma is visible protruding from the esophageal mucosal surface (left center of specimen).

On this positron emission computed tomography (PET) scan, esophageal cancer is evident as a golden lesion in the chest.

Palmoplantar keratoderma (Tylosis) of palms (A) and soles (B). Courtesy of The American Journal of Gastroenterology, Nature Publishing Group.

Diagram showing T1,T2 and T3 stages of esophageal cancer. Courtesy of Cancer Research UK and Wikimedia Commons.

Micrograph of squamous cell carcinoma of the esophagus (H&E Stain). Courtesy of Wikimedia Commons.

Low magnification micrograph of an intramucosal esophageal adenocarcinoma (H&E stain). Endoscopic mucosal resection specimen. Courtesy of Wikimedia Commons.

Stage

Survival Rate (%)
Localized46.7
Regional25.1
Distant4.8
All Stages19.9
Stage 0TisN0M0
Stage IT1N0-1M0
Stage IIT2N0M0
 T3N0M0
Stage IIIT3N1M0
 T1-3N2M0
Stage IVAT4N0-2M0
 Any TN3M0
Stage IVBAny TAny NM1
Stage 0TisN0M0
Stage IT1N0M0
Stage IIAT1N1M0
Stage IIBT2N0M0
Stage IIIT2N1M0
 T3N0-1M0
 T4aN0-1M0
Stage IVAT1-4aN2M0
 T4bN0-2M0
 Any TN3M0
Stage IVBAny TAny NM1
SyndromeGene(s)Inheritance PatternClinical ManifestationSurveillance Recommendations
Tylosis (non-Epidermolytic palmoplantar keratosis (PPK) or Howel Evans syndrome)RHBDF2Autosomal dominantSkin thickening of palms and soles. Non-epidermolytic PPK is associated with high risk of developing SCC of esophagus (40-90% by age of 70 years)[40, 37] Surveillance by upper GI endoscopy is recommended for family members with tylosis after 20 years of age[40]
Familial Barrett esophagus (FBE)Genes not identified Autosomal dominantFamilial aggregation of Barrett esophagus, adenocarcinoma of the esophagus and EGJ[130, 131] Potential family history of Barrett esophagus, and adenocarcinoma of the esophagus,should be determined for patients presenting with GERD, especially white men older than 40 years of age.
Bloom syndromeBLM/RECQL3Autosomal recessive Increased predisposition to a wide variety of malignancies[132] . Acute myeloid leukemia, acute lymphoblastic leukemia, lymphoid neoplasms, and Wilms tumor are the predominant cancers diagnosed before 25 years of age, whereas carcinomas of different anatomic sites including SCC of the esophagus are diagnosed after 20 years of age. Chromosomal quadraradials with breakage may be used for the diagnosis of BS [40] Screening for GERD with or without endoscopy after 20 years of age may be considered to detect cancer early. 
Fanconi anemia (FA)FANC[133] Autosomal recessive Characterized by congenital malformations, progressive pancytopenia, easy bruising, chromosomal breakage and an increased predisposition to the development of hematologic malignancies as well as solid tumors[40] . AML is the most common cancer type in patients with FA. However, patients with FA are also at an increased risk of developing SCC of head, neck and esophagus, cervical cancer, and brain tumors.[40, 134, 135] Enhanced mitomycin C–induced chromosomal breakage analysis can help in diagnosis.[136] Endoscopy of the esophagus may be considered as a surveillance strategy in individuals identified with FA.
Biopsy Finding Recommended Surveillance/Intervention
Barrett Esophagus (BE) without dysplasiaEndoscopic surveillance every 3 years for patients with BE without dysplasia on 2 consecutive endoscopies with biopsies within a year. 
BE with low-grade dysplasia (LGD)For patients with confirmed low-grade dysplasia and without life-limiting comorbidity, endoscopic therapy is considered as the preferred treatment modality, although endoscopic surveillance is an acceptable alternative (Endoscopy within 6 months is warranted to ensure that no HGD is present in the esophagus. Follow-up endoscopy is recommended annually until no dysplasia is detected on 2 consecutive endoscopies with biopsies)
BE with high- grade dysplasia (HGD)Patients with confirmed high-grade dysplasia should be managed with endoscopic therapy. In case of life-limiting comorbidity endoscopy within 3 months is recommended to rule out adenocarcinoma of the esophagus. Follow-up endoscopy every 3 months is recommended thereafter.