Stomach


Anatomy

Gross Anatomy

Divisions

The stomach is derived from the tubular embryonic foregut and begins as a dilation during the fifth week of gestation in the caudal portion. The embryonic stomach is invested by two mesenteries: dorsal (which becomes the gastrosplenic, gastrocolic, and gastrophrenic ligaments) and ventral (which becomes the hepatoduodenal and gastrohepatic ligaments of the lesser omentum and the falciform ligament). By the seventh week of gestation, the stomach descends, rotates, and further dilates with a disproportionate elongation of the greater curvature into its normal anatomic shape and position.

The most proximal region of the stomach is called the cardia and attaches to the esophagus. Immediately proximal to the cardia is the physiologically competent lower esophageal sphincter. The stomach is fixed at the gastroesophageal (GE) junction and pylorus, but its large midportion is mobile. The fundus represents the superior-most part of the stomach and is floppy and distensible. The angle of His is an important anatomic angle formed by the fundus with the left margin of the esophagus. The body of the stomach represents the largest portion and is also referred to as the corpus. The body is bounded on the right by the relatively straight lesser curvature and on the left by the longer greater curvature. At the angularis incisura, the lesser curvature abruptly angles to the right. The body of the stomach ends here and the antrum begins. Distally, the pylorus connects the distal stomach (antrum) to the proximal duodenum ( Fig. 49.1 ).

Fig. 49.1, Divisions of the stomach.

The left lateral segment of the liver covers a large portion of the stomach anteriorly. Posteriorly, the stomach is bounded by the diaphragm, left kidney, pancreas, aorta, and celiac trunk. Inferiorly, the stomach is attached to the transverse colon via the gastrocolic ligament. Superiorly, the GE junction is found approximately 2 to 3 cm below the diaphragmatic esophageal hiatus in the horizontal plane of the seventh chondrosternal articulation, a plane only slightly cephalad to the plane containing the pylorus. The gastrosplenic ligament attaches the proximal greater curvature to the spleen.

Blood Supply

The celiac artery provides the majority of the blood supply to the stomach ( Fig. 49.2 ). There are four main arteries—the left and right gastric arteries along the lesser curvature and the left and right gastroepiploic arteries along the greater curvature, with the left gastric artery being the largest. In addition, a substantial quantity of blood may be supplied to the proximal stomach by the inferior phrenic arteries and by the short gastric arteries from the spleen. Approximately 15% to 20% of patients have an aberrant left hepatic artery originating from the left gastric artery. Consequently, proximal ligation of the left gastric artery can result in acute left-sided hepatic ischemia. The right gastric artery arises from the hepatic artery (or sometimes the gastroduodenal artery). The left gastroepiploic artery originates from the splenic artery, and the right gastroepiploic artery originates from the gastroduodenal artery. The extensive anastomotic connections between these major vessels ensures that, in most cases, the stomach will survive if three out of four arteries are ligated, provided that the arcades along the greater and lesser curvatures are not disturbed. In general, the veins of the stomach parallel the arteries. The left gastric (coronary) and right gastric veins usually drain into the portal vein. The right gastroepiploic vein drains into the superior mesenteric vein, and the left gastroepiploic vein drains into the splenic vein.

Fig. 49.2, Blood supply to the stomach and duodenum showing anatomic relationships to the spleen and pancreas. The stomach is reflected cephalad.

Lymphatic Drainage

The lymphatic drainage of the stomach parallels the vasculature and drains into four zones of lymph nodes ( Fig. 49.3 ). The superior gastric group drains lymph from the upper lesser curvature into the left gastric and paracardial nodes. The suprapyloric group of nodes drains the antral segment on the lesser curvature of the stomach into the right suprapancreatic nodes. The pancreaticolienal group of nodes drains lymph high on the greater curvature into the left gastroepiploic and splenic nodes. The inferior gastric and subpyloric group of nodes drains lymph along the right gastroepiploic vascular pedicle. All four zones of lymph nodes drain into the celiac nodes and eventually into the thoracic duct. Although these lymph nodes drain different areas of the stomach, gastric cancers may metastasize to any of the four nodal groups, regardless of the cancer location. In addition, the extensive submucosal plexus of lymphatics accounts for the fact that there is frequently microscopic evidence of malignant cells several centimeters from gross disease.

Fig. 49.3, Lymphatic drainage of the stomach.

Innervation

As shown in Fig. 49.4 , the extrinsic innervation of the stomach is both parasympathetic (via the vagus nerve) and sympathetic (via the celiac plexus). The vagus nerve originates in the vagal nucleus in the floor of the fourth ventricle and traverses the neck in the carotid sheath to enter the mediastinum, where it divides into several branches around the esophagus. These branches coalesce above the esophageal hiatus to form the left and right vagus nerves. It is not uncommon to find more than two vagal trunks at the distal esophagus. At the GE junction, the left vagus is anterior, and the right vagus is posterior.

Fig. 49.4, Vagal innervation of the stomach. The line of division for truncal vagotomy is shown; it is above the hepatic and celiac branches of the left and right vagus nerves, respectively. The line of division for selective vagotomy is shown; this is below the hepatic and celiac branches.

The left vagus gives off the hepatic branch to the liver and continues along the lesser curvature as the nerve of Latarjet. The so-called criminal nerve of Grassi is the first branch of the right posterior vagus nerve; it is recognized as a potential cause of recurrent ulcers when left undivided. The right vagus nerve gives a branch off to the celiac plexus and continues posteriorly along the lesser curvature. A truncal vagotomy is performed above the celiac and hepatic branches of the vagi, whereas a selective vagotomy is performed below. A highly selective vagotomy is performed by dividing the crow’s feet to the proximal stomach while preserving the innervation of the antral and pyloric parts of the stomach. Most (90%) of the vagal fibers are afferent, carrying stimuli to the brain. Efferent vagal fibers originate in the dorsal nucleus of the medulla and synapse with neurons in the myenteric and submucosal plexuses. These neurons influence gastric motor function and gastric secretion. In contrast, the sympathetic nerve supply comes from T5 to T10, traveling in the splanchnic nerve to the celiac ganglion. Postganglionic fibers travel with the arterial system to innervate the stomach.

The intrinsic or enteric nervous system of the stomach consists of neurons in Auerbach and Meissner autonomic plexuses. In these locations, cholinergic, serotoninergic, and peptidergic neurons are present. The exact function of these neurons is poorly understood. Nevertheless, numerous neuropeptides have been localized to these neurons, including acetylcholine, serotonin, substance P, calcitonin gene–related peptide, bombesin, cholecystokinin (CCK), and somatostatin.

Gastric Morphology

The stomach is covered by peritoneum, which forms the outer serosa of the stomach. Below this is the thicker muscularis propria, or muscularis externa, which is composed of three layers of smooth muscles. The middle layer of smooth muscle is circular and is the only complete muscle layer of the stomach wall. At the pylorus, this middle circular muscle layer becomes progressively thicker and functions as a true anatomic sphincter. The outer muscle layer is longitudinal and predominates in the distal two thirds of the stomach. Within the layers of the muscularis externa is a rich plexus of autonomic nerves and ganglia, called Auerbach myenteric plexus. The submucosa lies between the muscularis externa and the mucosa and is a collagen-rich layer of connective tissue that forms the strongest layer of the gastric wall. In addition, it contains the rich anastomotic network of blood vessels and lymphatics as well as the Meissner plexus of autonomic nerves. The mucosa consists of surface epithelium, lamina propria, and muscularis mucosae. The muscularis mucosae is on the luminal side of the submucosa and is probably responsible for the rugae that greatly increases the stomach’s epithelial surface area. It also marks the microscopic boundary for invasive and noninvasive gastric carcinoma. The lamina propria represents a small connective tissue layer and contains capillaries, vessels, lymphatics, and nerves necessary to support the surface epithelium.

Gastric Microscopic Anatomy

Gastric mucosa consists of simple columnar epithelia interrupted by gastric pits containing one or more gastric glands. The cellular populations (and functions) of the cells forming this glandular epithelium vary based on their location in the stomach ( Table 49.1 ). The glandular epithelium is divided into cells that secrete products into the gastric lumen for digestion (parietal cells, chief cells, mucus-secreting cells) and cells that control function (gastrin-secreting G cells, somatostatin-secreting D cells). In the cardia, the mucosa is arranged in branched glands, and the pits are short. In the fundus and body, the glands are more tubular, and the pits are longer. In the antrum, the glands are more branched. The luminal ends of the gastric glands and pits are lined with mucus-secreting surface epithelial cells, which extend down into the necks of the glands for variable distances. In the cardia, the glands are predominantly mucus-secreting. In the body, the glands are mostly lined from the neck to the base with parietal and chief cells ( Fig. 49.5 ). There are a few parietal cells in the fundus and proximal antrum, but none in the cardia or prepyloric antrum. The endocrine G cells are present in greatest quantity in the antral glands.

Table 49.1
Gastric cell types, location, and function.
Cell Type Location Function
Parietal Body Secretion of acid and intrinsic factor
Mucus Body, antrum Mucus
Chief Body Pepsin
Surface epithelial Diffuse Mucus, bicarbonate, prostaglandins
Enterochromaffin-like Body Histamine
G Antrum Gastrin
D Body, antrum Somatostatin
Gastric mucosal interneurons Body, antrum Gastrin-releasing peptide
Enteric neurons Diffuse Calcitonin gene–related peptide, others
Endocrine Body Ghrelin

Fig. 49.5, Cells residing within a gastric gland.

Physiology

The principal function of the stomach is to prepare ingested food for digestion and absorption as it is propulsed into the small intestine. Receptive relaxation of the proximal stomach with ingestion of food enables the stomach to function as a storage organ. This relaxation enables liquids to pass easily from the stomach along the lesser curvature, whereas the solid food settles along the greater curvature of the fundus. In contrast to liquids, emptying of solid food is facilitated by the antrum, which propels solid food components into and through the pylorus. The antrum and pylorus function in a coordinated fashion, returning material to the proximal stomach until the size is suitable for delivery into the duodenum.

In addition to storing food, the stomach begins digestion of a meal. Starches undergo enzymatic breakdown through the activity of salivary amylase. Pepsin initiates protein digestion, although this hydrolysis is not completed in the stomach. The small intestine is primarily responsible for digestion of a meal and nutrient absorption.

Regulation of Gastric Function

Gastric function is under neural (sympathetic and parasympathetic) and hormonal control (peptides or amines that interact with target cells in the stomach). An understanding of the roles of endocrine and neural regulation of digestion is critical to understanding gastric physiology and the resultant physiologic effects of gastric surgical procedures on digestion. We initially focus here on peptide regulation of gastric function and then describe the interactions of these peptides with neural inputs in regard to acid secretion and gastric function.

Gastric Peptides

Gastrin

Gastrin is produced by G cells located in the gastric antrum and is the primary endocrine regulator of the secretory phase of a protein meal (see Table 49.1 ). It is synthesized as a prepropeptide and undergoes posttranslational processing by enzymatic cleavage in the rough endoplasmic reticulum and secretory vesicles to produce biologically reactive gastrin peptides. Several molecular forms of gastrin exist. The two major forms are G34 (big gastrin) and G17 (little gastrin). Ninety percent of antral gastrin is released as the 17–amino acid peptide, although G34 predominates in the circulation because its metabolic half-life is longer. The pentapeptide sequence contained at the carboxy terminus of gastrin is the biologically active component and is identical to that found on another gut peptide, CCK. CCK and gastrin differ by their tyrosine sulfation sites. Gastrin initiates its biologic actions by activation of surface membrane receptors. These receptors are members of the classic G protein–coupled receptor family and are classified as type A or B CCK receptors. The gastrin or CCK-B receptor has high affinity for gastrin and CCK, whereas type A CCK receptors have an affinity for sulfated CCK analogues and a low affinity for gastrin.

The release of gastrin is stimulated mostly by gastric distension, gastrin-releasing peptide (bombesin), and protein digestion products. Luminal acid inhibits the release of gastrin, specifically when the intragastric pH is below 3.0, via somatostatin release. In the antral location, somatostatin and gastrin release are functionally linked, and an inverse reciprocal relationship exists between these two peptides.

Gastrin is the major hormonal regulator of the gastric phase of acid secretion. Gastrin primarily does this by stimulating enterochromaffin-like (ECL) cells to synthesize and release histamine. However, gastrin also exerts direct actions on the parietal cell to stimulate acid release. Gastrin also has considerable trophic effects on both parietal cells and ECL cells. Prolonged hypergastrinemia from any cause leads to mucosal hyperplasia and an increase in the number of ECL cells and, under some circumstances, is associated with the development of gastric carcinoid tumors.

The detection of hypergastrinemia may suggest a pathologic state of acid hypersecretion but most commonly is the result of treatment with agents to reduce acid secretion, such as proton pump inhibitors (PPIs). Table 49.2 lists common causes of chronic hypergastrinemia. Hypergastrinemia that results from the administration of acid-reducing drugs is an appropriate response caused by loss of feedback inhibition of gastrin release by luminal acid. Lack of acid causes a reduction in somatostatin release, which causes increased release of gastrin from antral G cells. Hypergastrinemia can also occur in the setting of pernicious anemia, uremia, or after surgical procedures such as vagotomy. In contrast, gastrin levels increase inappropriately in patients with gastrinoma (Zollinger-Ellison syndrome [ZES]). These gastrin-secreting tumors are not located in the antrum and secrete gastrin autonomously.

Table 49.2
Causes of hypergastrinemia.
Ulcerogenic Causes Nonulcerogenic Causes
Antral G cell hyperplasia or hyperfunction Antisecretory agents (PPIs)
Retained gastric antrum Atrophic gastritis
Zollinger-Ellison syndrome Pernicious anemia
Gastric outlet obstruction Acid-reducing procedure (vagotomy)
Short-gut syndrome Gastric cancer
Chronic renal failure
PPIs , Proton pump inhibitors.

Somatostatin

Somatostatin is produced by delta cells and exists endogenously as either a 14–amino acid peptide or 28–amino acid peptide. The predominant molecular form in the stomach is somatostatin-14. It is produced by diffuse neuroendocrine cells located in the fundus and antrum. The principal stimulus for somatostatin release is antral acidification as well as gastrin-releasing peptide, whereas acetylcholine from vagal fibers inhibits its release. Somatostatin inhibits parietal cell acid secretion directly but also indirectly decreases acid secretion through inhibition of gastrin release from G cells and downregulation of histamine release from ECL cells.

Somatostatin receptors are also G protein–coupled receptors. Binding of somatostatin with its receptors is coupled to one or more inhibitory guanine nucleotide–binding proteins. Parietal cell somatostatin receptors appear to be a single subunit of glycoproteins with a molecular weight of 99 kDa, with equal affinity for somatostatin-14 and somatostatin-28. Somatostatin can inhibit parietal cell secretion through G protein–dependent and G protein–independent mechanisms. However, the ability of somatostatin to exert its inhibitory actions on cellular function is primarily thought to be mediated through the inhibition of adenylate cyclase, with a resultant reduction in cyclic adenosine monophosphate (cAMP) levels.

Histamine

Histamine (H 2 ) plays a prominent role in parietal cell stimulation. Administration of H 2 -receptor antagonists almost completely abolishes gastric acid secretion in response to gastrin and acetylcholine, suggesting that histamine may be a necessary intermediary of these pathways. Histamine is stored in the acidic granules of ECL cells and in resident mast cells. ECL cells are located in the oxyntic mucosa in direct proximity to the parietal cell. Histamine release is stimulated by gastrin, vasoactive intestinal peptide, ghrelin, acetylcholine, and epinephrine through receptor-ligand interactions on ECL cells. In contrast, somatostatin inhibits gastrin-stimulated histamine release through interactions with somatostatin receptors located on the ECL cell, with other inhibitors including peptide YY and prostaglandins. The ECL cell plays an essential role in parietal cell activation, possessing stimulatory and inhibitory feedback pathways that modulate the release of histamine.

Ghrelin

Ghrelin is a 28–amino acid peptide predominantly produced by endocrine cells of the oxyntic glands in the stomach. Ghrelin appears to be under both endocrine and metabolic control, has a diurnal rhythm, and likely plays a major role in the neuroendocrine and metabolic responses to changes in nutritional status. Ghrelin has been shown to stimulate growth hormone release as well. Ghrelin levels are high during fasting and increases shortly before meals, with decreased levels after meals. Decreased ghrelin levels have been associated with gastritis. Within the stomach, ghrelin increases gastric emptying and motility and increases gastric acid secretion.

In human volunteers, ghrelin administration enhances appetite and increases food intake. In patients who have undergone a gastric bypass or sleeve gastrectomy, ghrelin levels are lower. Although the mechanism responsible for suppression of ghrelin levels after bariatric surgery is unknown, it is suggested that ghrelin may be responsive to the normal flow of nutrients across the stomach. Other studies have suggested that ghrelin leads to a switch toward glycolysis and away from fatty acid oxidation, which would favor fat deposition. Ghrelin appears to be upregulated in times of negative energy balance and downregulated in times of positive energy balance. Ghrelin may come to have a role in the treatment and prevention of obesity.

Gastric Acid Secretion

The hydrogen-potassium-adenosine triphosphatase (ATPase) acid-secreting pump is located in the parietal cell. Gastric acid secretion by the parietal cell is regulated mainly by three stimuli—acetylcholine, gastrin, and histamine. Acetylcholine is the principal neurotransmitter modulating acid secretion and is released from the vagus and parasympathetic ganglion cells; it mainly exerts effects on M3 receptors. Vagal fibers not only have a direct effect on parietal cells, but also modulate peptide release from G cells and ECL cells as well as inhibit somatostatin secretion. Gastrin has direct hormonal effects on the parietal cell and also stimulates histamine release. Histamine has paracrine-like effects on the parietal cell and, as shown in Fig. 49.6 , plays a central role in the regulation of acid secretion by the parietal cell after its release from ECL cells. As depicted, somatostatin exerts inhibitory actions on gastric acid secretion. Release of somatostatin from antral D cells is stimulated in the presence of low intraluminal pH as well as vasoactive intestinal peptide and gastrin. After its release, somatostatin inhibits gastrin release through paracrine effects and modifies histamine release from ECL cells. Consequently, the precise state of acid secretion by the parietal cell depends on the overall influence of the positive and negative stimuli.

Fig. 49.6, Central role of the enterochromaffin-like (ECL) cell in regulation of acid secretion by the parietal cell. As shown, ingestion of a meal stimulates vagal fibers to release acetylcholine (cephalic phase). Binding of acetylcholine to M 3 receptors located on the ECL cell, parietal cell, and G cell results in the release of histamine, hydrochloric acid, and gastrin. Binding of acetylcholine to M 3 receptors on D cells results in the inhibition of somatostatin release. After a meal, G cells are also stimulated to release gastrin, which interacts with receptors located on ECL cells and parietal cells to cause the release of histamine and hydrochloric acid (gastric phase). Release of somatostatin from D cells decreases histamine release and gastrin release from ECL cells and G cells. In addition, somatostatin inhibits parietal cell acid secretion (not shown). The principal stimulus for the activation of D cells is antral luminal acidification (not shown). The principal stimulus for the activation of D cells is antral luminal acidification (not shown).

In the absence of food, there is always a basal level of acid secretion that is approximately 10% of maximal acid output (1 to 5 mmol/hr). This is reduced after vagotomy or H 2 -receptor blockade. Thus, it appears likely that basal acid secretion is caused by a combination of cholinergic and histaminergic input.

Stimulated Acid Secretion

Ingestion of food is the physiologic stimulus for acid secretion. Three phases of the acid secretory response to a meal have been described—cephalic, gastric, and intestinal. These three phases are interrelated and occur concurrently.

Cephalic phase

The cephalic phase originates with the sight, smell, thought, or taste of food, which stimulates neural centers in the hypothalamus. Although the exact mechanisms whereby senses stimulate acid secretion are not yet fully elucidated, it is hypothesized that several sites are stimulated in the brain. These higher centers transmit signals to the stomach via the vagus nerves, which release acetylcholine that activates muscarinic receptors located on target cells. Acetylcholine directly increases acid secretion by the parietal cells as well as stimulating ECL and G cells and inhibiting D cells. Although the intensity of the acid secretory response in the cephalic phase surpasses that of the other phases, it accounts for only 20% to 30% of the total volume of gastric acid produced in response to a meal because of its short duration.

Gastric phase

The gastric phase of acid secretion begins when food enters the gastric lumen. Protein products of ingested food interact with microvilli of antral G cells to stimulate gastrin release. Food also stimulates acid secretion by causing mechanical distention of the stomach. Gastric distention activates stretch receptors in the stomach to elicit the vagovagal reflex arc as well as local enteric nervous system acetylcholine release. The vasovagal reflex is abolished by proximal gastric vagotomy and is, at least in part, independent of changes in serum gastrin levels. Antral distention also causes gastrin release. The entire gastric phase accounts for most (60%–70%) of meal-stimulated acid output.

Intestinal phase

The intestinal phase of gastric secretion is initiated by entry of chyme into the duodenum, which initially stimulates gastrin release and suppresses gastric motility. It occurs after gastric emptying and lasts as long as partially digested food components remain in the proximal small bowel. It accounts for only 5% to 10% of the acid secretory response to a meal and does not appear to be mediated by serum gastrin levels. Chyme also stimulates release of CCK and secretin in the duodenum.

Activation and Secretion by the Parietal Cell

The two second messengers principally involved in stimulation of acid secretion by parietal cells are intracellular cAMP and calcium. These two messengers activate protein kinases and phosphorylation cascades. The intracellular events following ligand binding to receptors on the parietal cell are shown in Fig. 49.7 . Histamine causes an increase in intracellular cAMP, which initiates a cascade of phosphorylation events that culminates in activation of proton pump (H + , K + -ATPase). In contrast, acetylcholine and gastrin stimulate phospholipase C, which converts membrane-bound phospholipids into inositol triphosphate to mobilize calcium from intracellular stores. Increased intracellular calcium activates other protein kinases that ultimately activate H + , K + -ATPase in a similar fashion to initiate the secretion of hydrochloric acid.

Fig. 49.7, Intracellular signaling events in a parietal cell. As shown, histamine binds to H 2 receptors, stimulating adenylate cyclase through a G protein–linked mechanism. Adenylate cyclase activation causes an increase in intracellular cAMP levels, which activates protein kinases. Activated protein kinases stimulate a phosphorylation cascade, with a resultant increase in levels of phosphoproteins that activate the proton pump. Activation of the proton pump leads to extrusion of cytosolic hydrogen in exchange for extracytoplasmic potassium. In addition, chloride is secreted through a chloride channel located on the luminal side of the membrane. Gastrin binds to type B CCK receptors, and acetylcholine binds to M 3 receptors. Following the interaction of gastrin and acetylcholine with their receptors, phospholipase C is stimulated through a G protein–linked mechanism to convert membrane-bound phospholipids into inositol triphosphate (IP 3 ). IP 3 stimulates the release of calcium from intracellular calcium stores, leading to an increase in intracellular calcium that activates protein kinases, which activate the H + /K + -ATPase. (From Yeo C, Dempsey DT, Klein AS, et al, eds. Shackelford’s Surgery of the Alimentary Tract . 6th ed. Philadelphia: Saunders; 2007.) ATP , Adenosine triphosphate; ATPase , adenosine triphosphatase; CCK , cholecystokinin; G i , inhibitory guanine nucleotide protein; G s , stimulatory guanine nucleotide protein; PIP 2 , phosphatidylinositol 4,5-diphosphate; PLC , phospholipase C.

H + , K + -ATPase is the final common pathway for gastric acid secretion by the parietal cell. It is composed of two subunits, a catalytic α-subunit (100 kDa) and a glycoprotein β-subunit (60 kDa). During the resting state, gastric parietal cells store H + , K + -ATPase intracellularly. Cellular relocation of the proton pump subunits through cytoskeletal rearrangements must occur for acid secretion to increase in response to stimulatory factors. The subsequent heterodimer assembly of the H + , K + -ATPase subunits and insertion into the microvilli of the secretory canaliculus causes an increase in gastric acid secretion. A potassium chloride efflux pathway must exist to supply potassium to the extracytoplasmic side of the pump. Cytosolic hydrogen is secreted by H + , K + -ATPase in exchange for extracytoplasmic potassium (see Fig. 49.7 ), which is an electroneutral exchange and does not contribute to the transmembrane potential difference across the parietal cell. Secretion of chloride is accomplished through a chloride channel moving chloride from the parietal cell cytoplasm to the gastric lumen. The exchange of hydrogen for potassium requires energy in the form of adenosine triphosphate because hydrogen is being secreted against a gradient of more than a million fold. Because of this large energy requirement, the parietal cell has a mitochondrial compartment representing about one third of its cellular volume. In response to a secretagogue, the parietal cell undergoes a conformational change, and a several-fold increase in the canalicular surface area occurs ( Fig. 49.8 ). In contrast to stimulated acid secretion, cessation of acid secretion requires endocytosis of H + , K + -ATPase, with regeneration of cytoplasmic tubulovesicles containing the subunits, and this occurs through a tyrosine-based signal. The tyrosine-containing sequence is located on the cytoplasmic tail of the β-subunit and is highly homologous to the motif responsible for internalization of the transferrin receptor.

Fig. 49.8, Diagrammatic representation of resting and stimulated parietal cells. Note the morphologic transformation between the nonsecreting parietal cell and stimulated parietal cell, with increases in secretory canalicular membrane surface area.

More than 1 billion parietal cells are found in the normal human stomach and are responsible for secreting approximately 20 mmol/hr of hydrochloric acid in response to a protein meal. There is a linear relationship between maximal acid output and parietal cell number. However, gastric acid secretory rates may be altered in patients with upper gastrointestinal (GI) disease. For example, gastric acid is often increased in patients with duodenal ulcer or gastrinoma, whereas it is decreased in patients with pernicious anemia or gastric atrophy. Patients with proximal gastric ulcers have lower secretory rates, whereas distal, antral, or prepyloric ulcers are associated with acid secretory rates similar to rates in patients with duodenal ulcers.

Gastric acid plays a critical role in the digestion of a meal. It is required to convert pepsinogen into pepsin, elicits the release of secretin from the duodenum, and limits colonization of the upper GI tract with bacteria.

Pharmacologic Regulation

Given the role stomach acid plays in many disease pathologies and the diversity of mechanisms that stimulate acid secretion, there has been much interest in the development of many site-specific drugs aimed at decreasing acid output by the parietal cell. The best-known site-specific antagonists are the group collectively known as the H 2 - receptor antagonists , which inhibit the H 2 -receptor on the parietal cell. The most potent of the H 2 -receptor antagonists is famotidine, followed by ranitidine, nizatidine, and cimetidine.

PPIs block acid secretion more completely than H 2 -receptor antagonists because of their irreversible inhibition of the H + , K + -ATPase proton pump. After oral administration, these agents are absorbed into the bloodstream as prodrugs and selectively concentrate in the secretory canaliculus of the parietal cell. At low pH, they become ionized and activated, with the formation of an active sulfur group. The cysteine residues on the α-subunit of the H + , K + -ATPase form a covalent disulfide bond with activated PPIs, which irreversibly inhibits the proton pump. Because of the covalent nature of this bond, these PPIs have more prolonged inhibition of gastric acid secretion than H 2 blockers. For recovery of acid secretion to occur, new protein pumps must be synthesized.

Antacids and sucralfate are two other medications with effects on gastric acid; however, they are both less potent. Antacids typically contain aluminum hydroxide, calcium carbonate, or magnesium trisilicate and can be used to neutralize gastric acid and decrease acid delivery to the duodenum, although the exact mechanism is unclear. Sucralfate is a sucrose octasulfate complexed with aluminum hydroxide and has been shown to bind to injured gastric tissue and stimulate angiogenesis and granulation tissue formation. Given the lower potency, the indications for antacids and sucralfate are limited to treatment of mild GE reflux disease.

Other Gastric Secretory Products

Gastric juice

Gastric juice is the combined result of secretion by the parietal cells, chief cells, and mucous cells, in addition to swallowed saliva and duodenal refluxate. The electrolyte composition varies with the rate of gastric secretion. Parietal cells secrete an electrolyte solution that is isotonic with plasma and contains 160 mmol/L. The pH of this solution is 0.8. The lowest intraluminal pH commonly measured in the stomach is 2 because of dilution of the parietal cell secretion by other gastric secretions, which also contain sodium, potassium, and bicarbonate.

Intrinsic factor

Intrinsic factor is a 50-kDa glycoprotein produced by the parietal cells that is essential for the absorption of vitamin B 12 in the terminal ileum. It is secreted in amounts that far exceed the amounts necessary for vitamin B 12 absorption. In general, secretion of intrinsic factor parallels gastric acid secretion, yet the secretory response is not linked to acid secretion. For example, PPIs do not block intrinsic factor secretion nor do not alter the absorption of vitamin B 12 . Intrinsic factor deficiency can develop in the patients with pernicious anemia or in patients undergoing total gastrectomy; both groups of patients require vitamin B 12 supplementation. Treatment typically consists of vitamin B 12 supplementation via intramuscular injection of either cyanocobalamin or hydroxocobalamin, although oral administration may be similarly effective.

Pepsin

Pepsinogens are proteolytic proenzymes that are secreted by the glands of the gastroduodenal mucosa. Two types of pepsinogens are produced. Group 1 pepsinogens are secreted by chief cells and by mucous neck cells located in the glands of the acid-secreting portion of the stomach. Group 2 pepsinogens are produced by surface epithelial cells throughout the acid-secreting portion of the stomach, antrum, and proximal duodenum. In the presence of acid, both forms of pepsinogen are converted to pepsin by removal of a short amino-terminal peptide. Pepsin is an endopeptidase that preferentially hydrolyzes peptide linkages where one of the amino acids is aromatic and accounts for approximately 20% of the protein digestion in the GI tract. Pepsins have optimal function at a pH of 1.5 to 2.0 and become inactive at higher pH.

Mucus and bicarbonate

Mucus and bicarbonate combine to neutralize gastric acid at the gastric mucosal surface. They are secreted by the surface mucous cells and mucous neck cells located in the acid-secreting and antral portions of the stomach. Gastric mucin is a large glycoprotein and the mucus is a viscoelastic gel containing approximately 85% water and 15% mucin. It provides a mechanical barrier to injury, is relatively impermeable to pepsins, and also acts as an impediment to ion movement from the gastric lumen to the apical cell membrane. Mucus is in a constant state of flux because it is secreted continuously by mucosal cells while simultaneously being solubilized by luminal pepsin in the stomach. Mucus production is stimulated by vagal stimulation, cholinergic agonists, prostaglandins, and some bacterial toxins. In contrast, anticholinergic drugs and nonsteroidal antiinflammatory drugs (NSAIDs) inhibit mucus secretion. Helicobacter pylori secretes various proteases and lipases that break down mucin and impair the protective function of the mucous layer.

In the acid-secreting portion of the stomach, bicarbonate secretion is an active process, whereas in the antrum, active and passive secretion of bicarbonate occurs. However, the magnitude of bicarbonate secretion is considerably less than acid secretion. Although the luminal pH is 2, the pH observed at the surface epithelial cell is usually 7. The pH gradient found at the epithelial surface is a result of the unstirred layer of water in the mucus gel and of the continuous secretion of bicarbonate by the surface epithelial cells.

Gastric Motility

Gastric motility is regulated on three main levels: extrinsic neural control, intrinsic neural control, and myogenic control. The extrinsic neural controls are mediated through parasympathetic (vagus) and sympathetic pathways, whereas the intrinsic controls involve the enteric nervous system and interstitial cells of Cajal. In contrast, myogenic control resides in the excitatory membranes of the gastric smooth muscle cells.

Fasting Gastric Motility

The electrical basis of gastric motility begins with the depolarization of pacemaker cells located in the mid-body of the stomach along the greater curvature. Once initiated by the interstitial of Cajal, slow waves travel at 3 cycles/minute in a circumferential and antegrade fashion toward the pylorus. In addition to these slow waves, gastric smooth muscle cells are capable of producing action potentials, which are associated with larger changes in membrane potential than slow waves. Compared with slow waves, which are not associated with gastric contractions, action potentials are associated with actual muscle contractions. During fasting, the stomach goes through a cyclical pattern of electrical activity composed of slow waves and electrical spikes, which has been termed the migrating myoelectric complex. Each cycle of the migrating myoelectric complex lasts 90 to 120 minutes. The net effects of the myoelectric migrating complex are frequent clearance of gastric contents during periods of fasting. The exact regulatory mechanisms of myoelectric migrating complex activities are unknown, but these activities remain intact after vagal denervation.

Postprandial Gastric Motility

Ingestion of a meal results in a decrease in the resting tone of the proximal stomach and fundus, referred to as receptive relaxation and gastric accommodation, respectively. Because these reflexes are mediated by the vagus nerves, interruption of vagal innervation to the proximal stomach, such as by truncal vagotomy or proximal gastric vagotomy, can eliminate these reflexes, with resultant early satiety and rapid emptying of ingested liquids. In addition to its storage function, the stomach is responsible for the mechanical mixing of ingested solid food particles. This activity involves repetitive forceful contractions of the midportion and antral portion of the stomach, causing food particles to be propelled against a closed pylorus, with subsequent retropulsion of solids and liquids. The net effect is a thorough mixing of solids and liquids and sequential shearing of solid food particles to smaller than 1 mm prior to passage through the pylorus to the proximal duodenum.

The emptying of gastric contents is influenced by coordinated neural and hormonal mediators. Additionally, the chemical and mechanical properties and temperature of the intraluminal contents can influence the rate of gastric emptying. In general, liquids empty more rapidly than solids, and carbohydrates empty more readily than fats. In addition, hot and cold liquids tend to empty at a slower rate than ambient temperature fluids. These responses to luminal stimuli are regulated by the enteric nervous system. Osmoreceptors and pH-sensitive receptors in the proximal small bowel have also been shown to be involved in the activation of feedback inhibition of gastric emptying. Inhibitory peptides proposed to be active in this setting include CCK, vasoactive intestinal peptide, glucagon, and gastric inhibitory polypeptide.

Abnormal Gastric Motility

Symptoms of abnormal gastric motility are nausea/vomiting, postprandial fullness, early satiety, abdominal pain, and bloating. In the most severe cases, patients can suffer weight loss. The first steps in evaluating patients with suspected abnormal gastric motility after history and physical examination should be to exclude a mechanical obstruction with upper endoscopy and either a computed tomographic (CT) or magnetic resonance (MR) enterography or barium follow-through. The most common cause of gastroparesis is idiopathic, accounting for approximately half of patients. Other common causes include diabetes mellitus, viral infection (i.e., cytomegalovirus and Epstein-Barr virus), neurologic disease (i.e., multiple sclerosis, stroke, Parkinson disease), autoimmune disease, and certain medications such as tricyclic antidepressants, calcium channel blockers, and cyclosporine. Additionally, gastric motility can also be impaired after surgery due to either intentional or accidental vagotomy. Vagotomy results in loss of receptive relaxation and gastric accommodation in response to meal ingestion, with resultant early satiety, postprandial bloating, accelerated emptying of liquids, and delay in emptying of solids. Delayed gastric emptying is also seen after pancreatic surgery, most notably pancreaticoduodenectomy, where it is reported in 10% to 40% of patients postoperatively.

Clinical manifestations of diabetic gastropathy, which can occur in insulin-dependent or non–insulin-dependent patients, are thought to be related to a variety of factors. Impaired neural control via the vagus nerve, myenteric nervous system, interstitial cells of Cajal, and the underlying smooth muscle have been implicated. Additionally, hyperglycemia itself has been shown to cause a decrease in contractility of the gastric antrum, increase in pyloric contractility, relax the proximal stomach, and suppress the migrating myoelectric complexes. Hyperinsulinemia, which is often associated with non–insulin-dependent diabetes, may play a role in the gastroparesis seen in non–insulin-dependent diabetes because it also leads to suppression of migrating myoelectric complex activity.

Gastric-Emptying Studies

There are numerous ways to assess gastric emptying. The most common is a radionucleotide scan. This nuclear scintigraphy study is performed using a meal of radiolabeled food. Scans are obtained immediately after ingestion of the meal and at 1, 2, and 4 hours after the meal. Measurement of residual gastric contents at 4 hours provides the most sensitive means for diagnosing gastroparesis. At 4 hours, retention of 10% to 15% signifies mild gastroparesis, 15% to 35% is moderate, and greater than 35% is severe. More recently, some institutions include a clear liquid gastric-emptying study to detect patients with normal solid food emptying but delayed liquid emptying. There are multiple other diagnostic options; however, they are utilized less frequently. An upper abdominal x-ray series after barium swallow can provide information on gastric emptying and may reveal mechanical causes that could contribute to a delay, such as gastric outlet obstruction. An electrogastrogram or antroduodenal motility study assesses for nerve and muscle abnormalities; however, neither evaluates the actual functional significance. Wireless motility capsules have also been investigated as an alternative to scintigraphy, which can measure the pH, temperature, and pressure during its transit.

Treatment

Regardless of the cause of gastroparesis, initial first-line treatment of mild gastroparesis involves dietary modification. Patients should be encouraged to eat multiple, small meals with little fat or insoluble fiber. Acidic and spicy foods should also be avoided. Medications that affect gastric motility, such as opioids, calcium channel blockers, tricyclic antidepressants, and dopamine agonists, should be stopped when possible. Glycemic control should be optimized in diabetic patients.

Pharmacologic therapy is necessary for persistent symptoms despite the above nonpharmacologic modifications. First-line medical therapy is metoclopramide (Reglan), a dopamine 2 receptor antagonist that stimulates antral contractions and decreases postprandial relaxation of the fundus. Outside of the United States, domperidone, another dopamine 2 antagonist, is available for patients who do not respond to metoclopramide or who experience side effects. Macrolide antibiotics (erythromycin and azithromycin) are motilin agonists that act by stimulating fundal contraction and have also shown benefit; they can be considered for second-line therapy.

Surgery for gastroparesis is rarely required and only indicated for refractory symptoms despite maximal medical therapy, partly because poor improvement in symptoms was observed historically after traditional open operations including gastrojejunostomy and subtotal gastrectomy. Surgical venting gastrostomy tubes can be placed (if unable to place endoscopically) for venting and jejunostomy tubes can be placed if needed for nutrition. Pyloromyotomy and pyloroplasty are options for the surgical management of gastroparesis that function by lowering outflow resistance at the pylorus and enhancing any remaining gastric contractility. Surgical implantation of gastric electrostimulators has also been used as a treatment for refractory idiopathic and diabetic gastroparesis. In this technique, electrical leads are placed onto the antrum laparoscopically and connected to a subcutaneously positioned simulator that delivers high-frequency, low-energy current. A recent systematic review reported that pyloric surgery improved nausea and abdominal pain more than gastric electrical stimulation ; however, robust comparative trials are lacking. Endoscopic therapies, such as the gastric peroral endoscopic myotomy, are also being explored. Retrospective series have shown durable improvement up to 1 year; however, prospective clinical trials are needed. Endoscopic pyloric dilation and stenting can also be considered in select settings.

Gastric Barrier Function

The stomach’s barrier function depends on multiple physiologic and anatomic factors. Blood flow plays a critical role in gastric mucosal defense by providing nutrients and delivering oxygen to ensure that the intracellular processes that underlie mucosal resistance to injury can proceed unabated. Decreased gastric mucosal blood flow has minimal effects on ulcer formation until it approaches 50% of normal. When blood flow is reduced by more than 75%, marked mucosal injury results, which is exacerbated in the presence of luminal acid. After damage occurs, injured surface epithelial cells are replaced rapidly by the migration of surface mucous cells located along the basement membranes. This process is referred to as restitution or reconstitution.

Exposure of the stomach to noxious agents causes a reduction in the potential difference across the gastric mucosa. In normal gastric mucosa, the potential difference across the mucosa is −30 to −50 mV and results from the active transport of chloride into the lumen and sodium into the blood by the activity of Na + , K + -ATPase. Disruption of the tight junctions between mucosal cells causes the epithelium to become leaky to ions (i.e., Na + and Cl ) and a resultant loss of the high transepithelial electrical resistance normally found in gastric mucosa. In addition, agents such as NSAIDs or aspirin possess carboxyl groups that are nonionized at a low intragastric pH because they are weak acids. Consequently, they readily enter the cell membranes of gastric mucosal cells, whereas they will not penetrate the cell membranes at neutral pH because they are ionized. On entry into the neutral pH environment found in the cytosol, they become reionized, do not exit the cell membrane, and are toxic to the mucosal cells.

Peptic Ulcer Disease

Peptic ulcers are erosions in the GI mucosa that extend through the muscularis mucosae. The most common symptom of peptic ulcer disease (PUD) is dyspepsia, although the majority of patients with peptic ulcers are asymptomatic. PUD can be complicated by bleeding, gastric outlet obstruction, fistulization, and perforation. The two predominant causes of PUD are H. pylori and NSAIDs. Many other less common mechanisms exist as well, including ZES, other medication and infectious exposures, radiation therapy, and gastric bypass surgery.

Epidemiology

The incidence and prevalence of PUD in developed countries, including the United States, have been declining in recent decades, as has the progression to complicated PUD. This change is likely due to a combination of increased detection and eradication of H. pylori infection, more rational NSAID use, and environmental factors. The lifetime prevalence of PUD is estimated to be 5% to 10%, with an annual incidence of 0.1% to 0.3%; however, both of these numbers are likely overestimations currently in developed countries. In addition to declining incidence, epidemiological studies have shown decreased hospitalization and mortality related to PUD in the last two to three decades. Much of this decline in ulcer incidence and the need for hospitalization have stemmed from increased knowledge of ulcer pathogenesis. Specifically, the role of H. pylori has been clearly defined, and the risks of long-term NSAID use have been better elucidated. The need for surgery in the treatment of ulcer disease has also decreased primarily as a result of a marked decline in elective surgical therapy for chronic disease.

Pathogenesis

Peptic ulcers are caused by decreased defensive factors, increased aggressive factors, or both. Protective (or defensive) factors include mucosal bicarbonate secretion, mucus production, adequate blood flow, growth factors, cell renewal, and endogenous prostaglandins. Damaging (or aggressive) factors include hydrochloric acid secretion, pepsins, ethanol ingestion, smoking, duodenal reflux of bile, ischemia, NSAIDs, hypoxia, and, most notably, H. pylori infection. Although it is now clear that most ulcers are caused by H. pylori infection or NSAID use, it is still important to understand all of the other protective and causative factors to optimize treatment and ulcer healing and prevent disease recurrence.

Helicobacter pylori Infection

It is estimated that half of the world’s population is affected by H. pylori. While, previously, 80% to 95% of duodenal ulcers and approximately 75% of gastric ulcers were associated with H. pylori infection, its prevalence in peptic ulcers has fallen to 50% to 75% in developed countries more recently with improved diagnosis, treatment, and prevention. Infection with H. pylori has been shown to temporally precede ulcer formation, and when this organism is eradicated as part of ulcer treatment, ulcer recurrence is extremely rare. These observations have secured the place of H. pylori as a primary causative factor in the pathogenesis of PUD.

The interplay between bacterial and host factors determines the clinical outcome of H. pylori infection. H. pylori is a spiral-shaped, flagellate, gram-negative bacteria that resides in gastric-type epithelium within or beneath the mucus layer. Its shape and flagella aid its movement through the mucous layer, and it produces enzymes that help it adapt to this hostile environment. Mucolytic enzymes both facilitate passage through the mucus layer and protect the bacteria from mucin’s antibiotic effects. Most notably, H. pylori is a potent producer of urease, which is capable of splitting urea into ammonia and bicarbonate, creating an alkaline microenvironment in the setting of an acidic gastric milieu, allowing for the bacteria’s survival in the stomach. The bacteria attach to the gastric epithelial cells by binding to surface adhesions. H. pylori organisms are microaerophilic and can live only in gastric epithelium. Thus, H. pylori can also be found in heterotopic gastric mucosa in the proximal esophagus, in Barrett esophagus, in gastric metaplasia in the duodenum, within a Meckel diverticulum, and in heterotopic gastric mucosa in the rectum. The host response to H. pylori is at least partially determined genetically, with associations shown with interleukin 1β and toll-like receptors, both components of the inflammatory response.

The exact mechanisms responsible for H. pylori –induced GI injury are still not fully understood, but the following four potential mechanisms have been proposed (and likely interact) to cause a derangement of normal gastric and duodenal physiology that leads to subsequent ulcer formation:

  • 1.

    Production of toxic products that cause local tissue injury. Locally produced toxic mediators include breakdown products from urease activity (e.g., ammonia), cytotoxins, mucinase (which degrades mucus and glycoproteins), phospholipases that damage both epithelial and mucus cells, and platelet-activating factor (which is known to cause mucosal injury and thrombosis in the microcirculation).

  • 2.

    Induction of a local mucosal immune response. H. pylori can cause a local inflammatory reaction in the gastric mucosa, attracting neutrophils and monocytes, which then produce numerous proinflammatory cytokines and reactive oxygen metabolites.

  • 3.

    Increased gastrin levels and changes in acid secretion. In patients with antral H. pylori infection, basal and stimulated gastrin levels are significantly increased, presumably secondary to a reduction in somatostatin release from antral D cells because of infection with H. pylori. During the acute phase of H. pylori infection, acid secretion is decreased. With chronic infection, H. pylori has trophic effects on ECL and G cells, which can result in acid hypersecretion. A decrease in serum levels of somatostatin could also contribute to the gastric hyperacidity. However, if oxyntic glands are destroyed by the chronic infection, hypoacidity will result.

  • 4.

    Gastric metaplasia occurring in the duodenum. Metaplastic replacement of areas of duodenal mucosa with gastric epithelium likely occurs as a protective response to decreased duodenal pH, resulting from the above-described acid hypersecretion; this allows for H. pylori to colonize these areas of the duodenum, which causes duodenitis and likely predisposes to duodenal ulcer formation. The presence of H. pylori in the duodenum is more common in patients with ulcer formation compared with patients with asymptomatic infections isolated to the stomach.

Peptic ulcers are also strongly associated with antral gastritis. Studies performed before the H. pylori era demonstrated that almost all patients with peptic ulcers have histologic evidence of antral gastritis. It is now known that most cases of histologic gastritis are caused by H. pylori infection. Of patients with NSAID-associated ulcers, 25% have evidence of a histologic antral gastritis compared with 95% of patients with non–NSAID-associated ulcers. In most cases, the infection tends to be confined initially to the antrum and results in antral inflammation. The causative role of H. pylori infection in the pathogenesis of gastritis and PUD was first elucidated by Marshall and Warren in Australia in 1984. To prove this connection, Marshall himself ingested inocula of H. pylori after first confirming that he had normal gross and microscopic gastric mucosa. Within days, he developed abdominal pain, nausea, and halitosis as well as histologically confirmed presence of gastric H. pylori infection. Acute inflammation was observed histologically on days 5 and 10. By 2 weeks, acute inflammation had been replaced by chronic inflammation with evidence of a mononuclear cell infiltration. For their pioneering work, Marshall and Warren were jointly awarded the Nobel Prize in Medicine in 2005.

Evidence of infection is seen in childhood in developing countries and in adulthood in developed countries. Spontaneous remission is rare. There is an inverse relationship between infection rates and socioeconomic status. The reasons for this relationship are poorly understood, but it seems to be the result of factors such as sanitary conditions, familial clustering, lack of running water, and overcrowding. Such factors likely also explain why developing countries have a comparatively higher rate of H. pylori infection, especially in children. In the United States, H. pylori prevalence is higher in African Americans and Hispanics.

H. pylori infection is associated with many common upper GI disorders, but most infected individuals are asymptomatic. Healthy U.S. blood donors have an overall prevalence anywhere from 20% to 55%. H. pylori infection is almost always present in the setting of active chronic gastritis. In addition, most patients with gastric cancer have current or past H. pylori infection. Although the association between H. pylori and gastric cancer is strong, no causal relationship has been proven. However, H. pylori –induced chronic gastritis and intestinal metaplasia are thought to play a role. A meta analysis of case-control studies comparing H. pylori –positive and H. pylori –negative individuals found that infection was associated with a twofold increased risk of developing gastric cancer. There is also a strong association between mucosa-associated lymphoid tissue (MALT) lymphoma and H. pylori infection. Regression of these lymphomas has been demonstrated after eradication of H. pylori .

Invasive Tests

Urease assay

Endoscopic biopsy specimens should be taken from the gastric body and the antrum and are then tested for urease. Sensitivity in diagnosing infection is greater than 90%, and specificity is 95% to 100%, meaning there are almost never false-positive results. However, the sensitivity of the test is lowered in patients who are taking PPIs, H 2 -receptor antagonists, or antibiotics. Rapid urease test kits are commercially available and can detect urease in gastric biopsy specimens within 1 hour with a similar level of diagnostic accuracy.

Histology

Endoscopy can also be performed with biopsy samples of gastric mucosa, followed by histologic visualization of H. pylori using either routine hematoxylin-eosin stains or special stains (e.g., silver, Giemsa, Genta stains). Sensitivity is approximately 95% and specificity is 99%, making histology slightly more accurate than the urease assay testing. Similar to the urease assay, the sensitivity of histologic evaluation is lower in patients taking PPIs or H 2 -receptor antagonists, but it remains the most accurate test available even in this setting. Histology additionally affords the ability to assess the severity of gastritis and confirm the presence or absence of the organism; however, it is a more expensive option for evaluation of biopsy samples than the urease assay.

Culture

Culturing of gastric mucosa obtained at endoscopy can also be performed to diagnose H. pylori . The sensitivity is approximately 80%, and specificity is 100%. However, culture requires laboratory expertise, is not widely available, is relatively expensive, and diagnosis requires 3 to 5 days. However, it does provide the opportunity to perform antibiotic sensitivity testing on isolates, if needed.

Noninvasive Tests

Urea breath test

The carbon-labeled urea breath test is based on the ability of H. pylori to hydrolyze urea as a result of its production of urease. Both sensitivity and specificity are greater than 95%. As with other testing modalities, the sensitivity of the urea breath test is reduced in patients taking antisecretory medications and antibiotics. It is recommended that patients discontinue antibiotics for 4 weeks and PPIs for 2 weeks to ensure optimal test accuracy. The urea breath test is less expensive than endoscopy and samples the entire stomach. In evaluating treatment efficacy, false-negative results can occur if the test is performed too soon after treatment, so it is usually best to perform this test 4 weeks after therapy is completed.

Stool antigen

H. pylori bacteria are present in the stool of infected patients, and several assays have been developed that use monoclonal antibodies to H. pylori antigens to test fecal specimens. These tests have demonstrated sensitivities of greater than 90% and sensitivities of 86% to 92%. Several studies have shown that stool antigen testing has an accuracy of greater than 90% in detecting eradication of infection after treatment, on par with invasive histology and noninvasive urea breath testing. Additionally, stool antigen testing is likely the most cost-effective method for assessing treatment efficacy.

Serology

There are various enzyme-linked immunosorbent assay laboratory-based tests available and some rapid office-based immunoassays that are used to test for the presence of IgG antibodies to H. pylori . Serology has a 90% sensitivity but a more variable specificity rate between 76% and 96%, and tests need to be locally validated based on the prevalence of specific bacterial strains. Antibody titers can remain high for 1 year or longer after eradication; consequently, this test cannot be used to assess response to therapy. For these reasons, stool antigen and urea breath tests are the preferred modalities for diagnosis and evaluation of treatment efficacy in patients with PUD and suspected H. pylori infection.

Nonsteroidal Antiinflammatory Drugs

NSAIDs, including aspirin, are absorbed through the stomach and small intestine and function as inhibitors of the cyclooxygenase enzymes. Cyclooxygenase enzymes form the rate-limiting step of prostaglandin synthesis in the GI tract. Prostaglandins (including thromboxane A2) promote gastric and duodenal mucosal protection from luminal acid and pepsin via numerous mechanisms, including increasing mucin and bicarbonate secretion and increasing blood flow to the mucosal endothelium and promoting epithelial cell proliferation and migration to the luminal surface. The presence of NSAIDs disrupts these naturally protective mechanisms, increasing the risk of peptic ulcer formation in the stomach and the duodenum.

The Food and Drug Administration estimates that NSAIDs are associated with a 1% to 4% risk per year of a clinically significant GI event, including bleeding, pyloric obstruction, and perforation. The risk for mucosal injury or ulceration is roughly proportional to the anti-inflammatory effect associated with each NSAID. Compared with H. pylori ulcers, which are more frequently found in the duodenum, NSAID-induced ulcers are more often found in the stomach. H. pylori ulcers are also almost always associated with chronic active gastritis, whereas gastritis is not frequently found with NSAID-induced ulcers. When NSAID use is discontinued, the ulcers usually do not recur.

Gastric Ulcers

The modified Johnson anatomic classification system for gastric ulcers (i.e., types I through V, described in Table 49.3 ) was developed before the modern understanding that most ulcers are the consequence of H. pylori infection or NSAID usage. However, despite having an increased understanding of the mechanisms of how and why most ulcers develop, this historical classification system is still relevant to surgical treatment because it dictates what operation should be performed in the setting of complications of such ulcers.

Table 49.3
Gastric ulcer types.
Type Location Acid Level
I Lesser curve at incisura Low to normal
II Gastric body with duodenal ulcer Increased
III Prepyloric Increased
IV High on lesser curve Normal
V Anywhere Normal, NSAID-induced
NSAID , Nonsteroidal antiinflammatory drug.

Gastric ulcers can occur at any location in the stomach, although they usually manifest on the lesser curvature near the incisura. Approximately 60% of ulcers are in this location and are classified as type I gastric ulcers. These ulcers are generally not associated with excessive acid secretion and may occur with low to normal acid output. Most occur within 1.5 cm of the histologic transition zone between the fundic and antral mucosa and are not associated with duodenal, pyloric, or prepyloric mucosal abnormalities. In contrast, type II gastric ulcers (approximately 15%) are located in the body of the stomach in combination with a duodenal ulcer. These types of ulcers are usually associated with excess acid secretion. Type III gastric ulcers are prepyloric ulcers and account for approximately 20% of the lesions. They also behave similar to duodenal ulcers and are associated with hypersecretion of gastric acid. Type IV gastric ulcers occur high on the lesser curvature, near the GE junction. The incidence of type IV gastric ulcers is less than 10%, and they are not associated with excessive acid secretion. Type V gastric ulcers can occur at any location and are associated with long-term NSAID use. Finally, some ulcers may appear on the greater curvature of the stomach, but the incidence is less than 5%.

Gastric ulcers rarely develop before the age of 40 years, and the peak incidence occurs in individuals 55 to 65 years old. Gastric ulcers are more likely to occur in individuals in a lower socioeconomic class and are slightly more common in the non-white compared to the white population. Some clinical conditions that may predispose to gastric ulceration include chronic alcohol intake, smoking, long-term corticosteroid therapy, infection, and intra arterial therapy. With regard to acid and pepsin secretion, the presence of acid appears to be essential to the production of a gastric ulcer; however, the total secretory output appears to be less important. In contrast to the acidification of the duodenum leading to ulcer formation, patients with gastric ulcers caused by H. pylori can have normal or reduced gastric acid production. Ulcer formation is more likely due to an inflammatory response to the bacterial infection itself. Nevertheless, rapid healing follows antacid therapy, antisecretory therapy, or vagotomy even when the lesion-bearing portion of the stomach is left intact because in the presence of gastric mucosal damage, acid is ulcerogenic, even when present in normal or less than normal amounts.

Clinical Manifestations

One clinical challenge of gastric ulcer management is the differentiation between gastric carcinoma and a benign ulcer. This is in contrast to duodenal ulcers, in which malignancy is extremely rare. Similar to duodenal ulcers, gastric ulcers are also characterized by recurrent episodes of quiescence and relapse. They also cause pain, bleeding, and obstruction and can perforate. Occasionally, benign ulcers have also been found to result in spontaneous gastrocolic fistulas. Surgical intervention is required in patients who develop complications from gastric ulcer disease. Patients who develop significant bleeding from gastric ulcers usually are older, are less likely to stop bleeding spontaneously, and have higher morbidity and mortality rates than patients with bleeding from a duodenal ulcer. The most frequent complication of gastric ulceration is perforation. Most perforations occur along the anterior aspect of the lesser curvature. In general, older patients have increased rates of perforations, and larger ulcers are associated with higher morbidity and mortality. Similar to duodenal ulcers, gastric outlet obstruction can also occur in patients with type II or III gastric ulcers. However, one must carefully differentiate between benign obstruction and obstruction secondary to carcinoma.

Diagnosis and Treatment

The diagnosis and treatment of gastric ulceration generally mirror the diagnosis and treatment of duodenal ulcer disease. The significant difference is the possibility of malignancy in a gastric ulcer. This critical difference demands that cancer be ruled out in acute and chronic presentations of gastric ulcer disease. Acid suppression and H. pylori eradication are important aspects of any treatment.

As with duodenal ulcers, intractable nonhealing ulcers are becoming increasingly less common. It is important to ensure that adequate time has elapsed and appropriate therapy has been administered to allow healing of the ulcer to occur; this includes confirmation that H. pylori has been eradicated and that NSAIDs have been eliminated. The presentation of a nonhealing gastric ulcer in the H. pylori era should raise serious concerns about the presence of an underlying malignancy. These patients should undergo a thorough evaluation with multiple biopsies to exclude malignancy ( Fig. 49.9 ). The approach for a complicated gastric ulcer varies depending on the type of ulcer and its association with pathophysiologic acid levels. Types I and IV ulcers, which are not associated with increased acid levels, do not require acid-reducing vagotomy. Fig. 49.10 is an algorithm for managing complicated gastric ulcers.

Fig. 49.9, Algorithm for evaluation, treatment, and surveillance of a patient with a gastric ulcer.

Fig. 49.10, Algorithm for the management of complicated gastric ulcer disease.

You're Reading a Preview

Become a Clinical Tree membership for Full access and enjoy Unlimited articles

Become membership

If you are a member. Log in here