Chapter 6
Aqueous Humor: Secretion And Dynamics
B'ANN TRUE GABELT, JULIE A. KILAND, BAOHE TIAN and PAUL L. KAUFMAN
Main Menu   Table Of Contents

Search

ANATOMY OF THE CILIARY BODY
FORMATION AND SECRETION OF AQUEOUS HUMOR
DIFFUSION
DIALYSIS AND ULTRAFILTRATION
GIBBS-DONNAN EQUILIBRIUM
SECRETION
BLOOD–AQUEOUS BARRIER
SECONDARY AQUEOUS
AQUEOUS HUMOR COMPOSITION
AQUEOUS FLOW
METHODS OF MEASURING RATES OF AQUEOUS FORMATION
MEASUREMENT OF FORMATION BY THE RATE OF APPEARANCE OR DISAPPEARANCE OF A CHEMICAL SUBSTANCE FROM THE AQUEOUS
MATHEMATICAL DERIVATION OF AQUEOUS FORMATION RATE
FACTORS AFFECTING AQUEOUS FORMATION
PATHWAYS OF AQUEOUS FLOW
AQUEOUS OUTFLOW
METHODS OF MEASUREMENT OF AQUEOUS OUTFLOW
TRABECULAR PATHWAY
OTHER AGENTS
UVEOSCLERAL PATHWAY
ACKNOWLEDGMENTS
REFERENCES

The aqueous humor is a transparent, colorless solution continuously formed from plasma by the epithelial cells of the ciliary processes. It is secreted into the posterior chamber, passes from the posterior chamber through the pupil into the anterior chamber, and is drained at the anterior chamber angle. Most of the aqueous drains into the venous circulation via the trabecular meshwork, Schlemm's canal, scleral collector channels, and aqueous and episcleral veins; the remainder drains into the orbit via the interstices of the ciliary muscle, the suprachoroidal space, and the sclera (Fig. 1). The composition and formation of aqueous resembles that of cerebrospinal fluid.1 Aqueous humor is thought to serve several functions:

Fig. 1. Cross section through the anterior segment of the eye illustrating the chamber angle. Aqueous humor is formed by active secretion from the ciliary processes (A). Drainage occurs via the outflow pathways, principally the trabecular meshwork (B) and Schlemm's canal (C) into the aqueous veins (D). A smaller proportion of the aqueous humor makes its way directly into the ciliary body (uveoscleral pathway) and is drained by way of the ciliary muscle, the suprachoroidal space, and the sclera (E). (From Karnezis TA, Murphy MB: Dopamine receptors and intraocular pressure. Trends Pharmacol Sci 9:389–390, 1988, with permission.)

  1. Aqueous delivers oxygen and nutrients to, and removes waste products, blood, macrophages, inflammatory products, or other debris from the posterior cornea, crystalline lens, and perhaps the anterior vitreous, structures that are necessarily avascular.
  2. Continuous formation and drainage of the aqueous helps maintain the intraocular pressure (IOP), necessary for maintaining the shape and internal alignment of the ocular structures and, consequently, optimal optical properties.
  3. The aqueous maintains a transparent and colorless medium of lower refractive index between the posterior cornea and the lens, and thus constitutes an important component of the eye's optical system.

Circulation of the aqueous in the anterior chamber occurs via hydrostatic phenomena, including mechanical forces caused by eyeball and head movements, thermal currents resulting from the temperature differential between the warmer vascular iris and the cooler avascular cornea, and the pressure gradients between the posterior chamber, anterior chamber, and episcleral veins. As the fluid bathes the anterior lens, iris, and corneal endothelium, its composition is altered as a result of the exchange of nutrients, cellular waste products, and other substances within these structures. The entire volume of the aqueous humor is replaced every 90 to 100 minutes.2

Continuous formation and drainage of the aqueous is essential to the good health of the eye. In the absence of aqueous circulation, the cornea is thickened, the anterior chamber is absent, the iris is partly atrophic, and the lens is cataractous.3

This chapter reviews the anatomy and physiology of aqueous humor circulation from formation to drainage. The ciliary body and its secretory mechanisms, the blood–aqueous barrier, the aqueous humor composition, the methods of measuring the aqueous flow rate and factors affecting it, the pathways of aqueous flow within the eye, and the aqueous outflow system are all discussed.

Back to Top
ANATOMY OF THE CILIARY BODY
The ciliary body forms a ring along the inner wall of the eyeball, and extends anteriorly from the scleral spur and iris to the ora serrata posteriorly. The greater part of the ciliary body mass is accounted for by the ciliary muscle, the bundles of which are arranged in three regional orientations: radial, circular, and longitudinal. When viewed in transverse section, the ciliary body appears as an isosceles triangle (Fig. 2). The base of the triangle faces anteriorly, while one of its sides lies along the sclera, separated from it only by a potential space continuous with the suprachoroidal space. The other side, or inner portion of the ciliary body, is divided anatomically into two parts: the posterior portion (pars plana) and the anterior portion (pars plicata). Projecting inwardly from the pars plicata are approximately 70 villus-like structures: the ciliary processes. Viewed posteriorly, the ciliary processes appear as radial ridges, to which collectively the name corona ciliaris has been given. It is the ciliary processes that are responsible for aqueous formation.

Fig. 2. Transverse section of the human ciliary body. (From Lütjen-Drecoll E, Rohen JW: Pathology of the trabecular meshwork in primary open angle glaucoma. In Kaufman PL, Mittag TW, eds.: Glaucoma. Textbook of Ophthalmology Series, Vol. 7. New York: CV Mosby, 1994:1.1–1.16, with permission.)

The ciliary processes are long and slender in early life but become more blunt in later years. Their dimensions vary but average 2 mm in length anteroposteriorly, 0.5 mm in width, and 1 mm in height.4 The processes are greatly convoluted. Structurally, they consist of capillaries surrounded by a loose connective tissue, encircled by a double epithelial layer. This structure provides a large surface area (in rabbits5 approximately 5.7 cm2) for capillaries to be in close proximity to the double layer of epithelium and for the epithelium to face the posterior chamber. This arrangement maximizes access of the ciliary body secretions into the small space of the posterior chamber.

The blood supply of the ciliary body has a dual origin.4,6–13 The two (medial and lateral) long posterior ciliary arteries, which penetrate the sclera posteriorly and travel anteriorly in the suprachoroid, give rise to the major arterial circle of the iris, the vascular structure that supplies the inner and anterior division of the ciliary muscle, ciliary processes and iris.13 The seven anterior ciliary arteries, which penetrate the sclera anteriorly after supplying the extraocular rectus muscles, also contribute to the major iris arterial circle, and to the outer and posterior areas of the ciliary muscle, and the peripheral (anterior) area of the choroid. The microvasculature of the ciliary processes themselves, arising from the short radial ciliary arteries, which in turn arise from the major arterial circle of the iris, is arranged into three distinct vascular areas. The first of these is located at the anterior end of the major processes, with venous drainage achieved via venules passing the ciliary body. The second and third vascular areas supply the major and minor ciliary processes, and are drained posteriorly by venules located at the margin of the ciliary processes. The system of ciliary process venules in turn drains mainly through the vortex system of the choroid. Each short radial ciliary artery has many branches, providing an extensive capillary network (Fig. 3). The capillaries of the ciliary processes are large, thin-walled, and highly fenestrated. Thus, the capillary network of the ciliary processes provides a large surface area of highly permeable vessels to initiate the process of aqueous formation.

Fig. 3. A: Blood supply to the ciliary processes. LCM, longitudinal ciliary muscle; RCM, radial ciliary muscle; CCM, circular ciliary muscle. B: Vascular architecture in the human ciliary body. (1), Perforating branches of the anterior ciliary arteries; (2), major arterial circle of the iris; (3), first vascular territory. The second vascular territory is depicted in 4a, marginal route and 4b, capillary network in the center of this territory. (5), third vascular territory; (6 and 7), arterioles to the ciliary muscle; (8) recurrent choroidal arteries. Light circles, terminal arterioles; dark circle, efferent venous segment. (A, From Caprioli J: The ciliary epithelia and aqueous humor. In Hart M, ed.: Adler's Physiology of the Eye, 9th ed. St. Louis: Mosby Year-Book, 1992:228–247, with permission; B, From Funk R, Rohen JW: Scanning electron microscopic study on the vasculature of the human anterior eye segment, especially with respect to the ciliary processes. Exp Eye Res 51:651, 1990, with permission.)

The capillaries are surrounded by the stroma, which consists mainly of loose connective tissue and collagen. A basement membrane, which is the thickened anterior continuation of Bruch's membrane separating choroid and retinal pigment epithelium, separates the stroma from the epithelial layers.

The ciliary epithelial cells have been subject to intensive study by light and electron microscopy. A striking feature is the interdigitation of the lateral surfaces of adjacent cells and the basal infoldings (Fig. 4), which are characteristic features of secretory epithelia concerned with fluid transport.14 The relation of the two epithelial cell layers is of importance because as the secreted aqueous is derived from an ultrafiltrate of blood in the stroma of the ciliary body, transport must occur across both layers. The double-layered ciliary epithelium itself is derived from anterior continuations of the retinal pigmented epithelium (forming the pigmented layer), and the neuroepithelium from which the retinal cells are derived (forming the nonpigmented layer). However, during embryogenesis, invagination of the neuroepithelium occurs, with the result that the apical surfaces of each cell layer in the ciliary epithelium face one another, while the basolateral surface of the nonpigmented layer faces directly into the posterior aqueous chamber. Conversely, the basolateral surface of the pigmented layer is tightly bound to the basement membrane. Cells of the pigmented epithelium contain many melanin granules. Gap junctions are present between the lateral interdigitations of the pigmented cells. Desmosomes also occur between the lateral interdigitations of the pigmented and nonpigmented epithelia and between their apical membranes.

Fig. 4. Schematic diagram of nonpigmented and pigmented epithelial cells. Note apices of cells facing each other. Basal infoldings (BI); basement membrane (BM); ciliary channels (CC); desmosomes (DES); fenestrated capillary endothelium (FE); gap junction (GJ); melanosome (MEL); mitochondrion (MIT); red blood cell (RBC); rough endoplasmic reticulum (RER); tight junction (TJ). (From Caprioli J: The ciliary epithelia and aqueous humor. In Hart M (ed): Adler's Physiology of The Eye, 9th ed, pp 228–247. St. Louis, Mosby Year-Book, 1992, with permission.)

The cells of the inner nonpigmented epithelium possess numerous intermediate-sized mitochondria, and the rough endoplasmic reticulum is particularly well developed, indicative of active protein synthesis. Small vacuoles may be present in large numbers near the apex of these cells. Occasionally, a few pigmented granules may be seen. Tight junctions are present in the lateral interdigitations between the nonpigmented cells (Fig. 4), thus forming a barrier for the passage of larger molecules between the cells. This important physiologic barrier, constituting part of the blood-aqueous barrier, is discussed further in a later section.

The internal limiting membrane of the ciliary body is a more complex and thicker structure than its posterior counterpart on the retina, in part because it serves as the basement membrane for the inverted nonpigmented ciliary epithelium, and as the site of insertion of the zonular fibers,4 to which the lens is attached.

Back to Top
FORMATION AND SECRETION OF AQUEOUS HUMOR
Early in the twentieth century, aqueous humor was regarded as a stagnant fluid.15 However, this misconception was revoked after a number of experiments designed to investigate this were carried out, including Seidel's procedure,16 in which a cannula connected to a reservoir of indigo carmine dye was inserted into the anterior chamber of the rabbit eye. The reservoir was raised, thus creating a pressure of 15 mm Hg, and the dye was seen to enter the anterior chamber and subsequently the episcleral veins. From this, it was concluded that aqueous humor is continuously formed and drained, and it is to a large extent from this historic work that the modern study of aqueous humor dynamics has developed.

Other aspects of the anatomy and physiology of aqueous drainage were discovered subsequently. Boerhaave first described the presence of the aqueous veins,17 and Ascher18 observed a clear fluid in veins of the episclera and demonstrated by means of external compression with a glass rod that these veins were interconnected with veins containing blood. Goldmann19 demonstrated that these vessels contained aqueous humor by injecting fluorescein intravenously and observing the dye entering the anterior chamber and subsequently the aqueous veins. Ashton20 identified an aqueous vein in a living human eye, and postmortem examination using a neoprene cast showed that there was a direct passage between the vessel and Schlemm's canal.

Three physiologic processes are known to contribute to the formation and chemical composition of the aqueous. These are diffusion, ultrafiltration (and the related dialysis), and active secretion. The first two are passive and, therefore, require no active cellular participation. Diffusion of solutes across cell membranes occurs down a concentration gradient. Substances with high lipid solubility coefficients that can easily penetrate biological membranes move readily in this way. Ultrafiltration is the term used to describe the bulk flow of blood plasma across the fenestrated ciliary capillary endothelia, into the ciliary stroma, which can be increased by augmentation of the hydrostatic driving force. This process is responsible for the formation of the reservoir of the plasma ultrafiltrate in the stroma, from which the posterior chamber aqueous is derived, via active secretion across the ciliary epithelium. Active secretion requires energy, normally provided via the hydrolysis of adenosine triphosphate (ATP). The energy is used to secrete substances against a concentration gradient.

Of these three processes, active secretion is believed to contribute the most to the chemical composition and volume of the posterior chamber aqueous, accounting for 80% to 90% of total aqueous humor formation.21–29

Back to Top
DIFFUSION
Diffusion arises from the fact that the molecules in a fluid are in constant, random motion. The magnitude and rate of motion vary directly with the temperature. If the molecules in a liquid or gas are not evenly distributed, then simply by the laws of statistical probability the molecules will eventually reach a state of equilibrium whereby they are redistributed equally. For example, if there is initially a cloud of smoke particles on the right side of a closed room without air currents, more particles will move from the right side to the left, than from left to right. This process will continue until there is a relatively even distribution throughout the room, at which time the number of particles going from right to left at any moment will be equal to the number moving in the opposite direction.

A similar process occurs in single solutions and in situations in which two solutions are separated by a membrane, as long as the membrane is permeable to at least some of the constituents of the solution (semipermeable membrane). Most capillary walls are permeable to water, dissolved gases, and many small molecules and ions. In a stagnant system, substances of higher concentration on one side of a semipermeable membrane show a net movement to the side of lower concentration until the concentrations are equal on both sides. When equilibrium is reached, movement across the membrane still occurs, but the number of particles going in one direction equals the number going the other way, thus yielding no net movement.

It should be noted that water (or another solvent) participates in this process. Net movement of water is usually in the opposite direction of solute movement, since a higher concentration of solute means, in effect, a lower concentration of solvent.

Fick's law describes quantitatively the net movement of a substance across a semipermeable membrane where only diffusion is occurring:

Rate of movement = K(C1 − C2)

where

C1 = concentration on side with higher concentration

C2 = concentration on side with lower concentration

K = constant, which depends on nature and permeability of membrane, nature of solute and solvent, and temperature.

The less permeable the membrane to the solute or solvent, the lower the temperature, and the more viscous the fluid medium, then the longer will be the time required for equilibration. Therefore, diffusion tends to occur more rapidly through extracellular fluids than across cells. It should also be remembered that conditions for diffusion are markedly altered in a dynamic system such as the ciliary processes, in which blood is flowing rapidly, aqueous humor is constantly being formed and carried away, and other processes occur, such as those described below.

Back to Top
DIALYSIS AND ULTRAFILTRATION
In most biologic solutions, there exists a combination of salts, sugars, proteins, and other large molecules. Most biological membranes are permeable to water, salt, and some small organic molecules. However, these membranes are relatively impermeable to larger molecules, such as proteins.

If a solution of protein and salt is separated from either water or a less concentrated salt solution by a membrane permeable to the salt and water but not to the protein, then there will be a net movement of water to the protein side by diffusion, and a movement of salt away from the protein side. The protein, of course, cannot move across the membrane. This process is called dialysis (Fig. 5) and is utilized, for example, for removing unwanted salts and other toxic substances from the blood, using the peritoneum as the membrane (peritoneal dialysis), or by using a synthetic membrane such as that found in a dialysis machine (artificial kidney).

Fig. 5. Dialysis. The presence of protein molecules (large circles) induces a net movement of water (small dots) across the semipermeable membrane to the protein side. There is also a net movement of salt molecules (broken circles) away from the protein side. With the exception of protein, movement of all molecules occurs in both directions, but net movement is in the direction as indicated by the solid arrows. (Courtesy of RL Stamper, MD.)

By the addition of a hydrostatic pressure on the protein side of the system, the exchange of salt and water can be accelerated. This process is called ultrafiltration and differs from dialysis only because the hydrostatic pressure changes the rate of movement of ions and slightly changes their final respective concentrations. The fluid formed by the process of dialysis is called a dialysate, and that formed by ultrafiltration is called an ultrafiltrate. Ultrafiltration is the process that occurs across capillary walls due to the higher pressure and higher protein concentration in the plasma as compared with the extracellular space (Fig. 6).

Fig. 6. Ultrafiltration. This process is similar to dialysis, but with the addition of a hydrostatic pressure that increases the rate of net movement of water and salt molecules across the semipermeable membrane. The final equilibrium concentrations on either side of the membrane are the same as in dialysis. The hydrostatic pressure merely increases the rate at which equilibrium is achieved. (Courtesy of RL Stamper, MD.)

Back to Top
GIBBS-DONNAN EQUILIBRIUM
The salt does not distribute itself equally on both sides of the membrane. Since the protein carries an electrical charge (generally an overall negative charge at physiologic pH), the positive ions in solution tend to be bound to the negatively charged residues of the protein molecules. Thus, there is an excess of cations such as Na+ or K+ on the protein side of the membrane.

This unequal distribution is called the Gibbs-Donnan effect, and the quantitative relationships between the various ions (at least in simple systems) is predictable (Fig. 7). In order to maintain electrical balance, the total positive charges (e.g., number of Na+ and K+ ions) on one side must equal the total negative charges (e.g., sum of negative protein charges, Cl ions, and any other negative ions). Furthermore, it has been shown that the final equilibrium concentrations of Na+ and Cl on each side of the membrane are related according to the following equation1,30,31:

[Na+]1/[Na+]2 = [CL]2/[CL]1

Fig. 7. Gibbs-Donnan effect. Because protein molecules carry an overall electrical charge (usually negative) at physiological pH, the distribution of ions at equilibrium is altered slightly to reflect the tendency of the protein molecules to attract oppositely charged and repel like-charged ions. Therefore there is a higher concentration of cations on the protein side and a higher concentration of anions on the nonprotein side of the system. (Courtesy of RL Stamper, MD.)

Thus, if the aqueous humor were like extracellular fluid in most capillary beds of other parts of the body, we should expect to see a protein-free aqueous solution with ionic concentrations like those on side 2 of a Gibbs-Donnan ultrafiltrate of plasma. The Na+ and K+ concentrations should be less than plasma, and the Cl and HCO3 concentrations should be slightly higher. Further, the actual values of the ratios of the concentrations of each respective ion in aqueous to plasma would conform to the values obtained experimentally when plasma is dialyzed against its own ultrafiltrate. In addition, no organic substance should be at higher concentration than in the plasma because diffusion and ultrafiltration can, at most, only equalize the concentrations of organic substances. These processes cannot promote an excess concentration on side 2 of the membrane.

Although aqueous humor resembles a dialysate of plasma in many ways, as will be seen, the ionic concentrations do not quite fit the Gibbs-Donnan predictions, and some nonionic substances have higher concentrations in aqueous than in plasma (Table 1). Such a condition can only occur in the presence of some active metabolic process.1,22

TABLE 1. Aqueous Versus Dialysate

SubstanceConcentration Ratio (aqueous/plasma)Concentration Ratio (dialysate/plasma)
Na+0.960.945
K+0.9550.96
Ca2+0.580.65
Mg2+0.780.80
Cl0.0151.04
HCO31.261.04
H2CO31.291.00
Glucose0.860.97
Urea0.871.00

Values are for the rabbit eye. Although the chemical composition of aqueous humor bears some similarities to a plasma dialysate, the small differences are significant. These differences can only be accounted for theoretically by an active secretory mechanism. (Adapted from Davson H: The Aqueous Humor and the Intraocular Pressure. In: Physiology of the Eye, 5th ed, pp 3–95. New York, Pergamon Press, 1990.)

 

Back to Top
SECRETION
Secretion implies an active process that selectively transports some substances across the cell membrane. Because energy is consumed, substances can be moved across a concentration gradient in a direction opposite to that which would be expected by passive mechanisms alone. One example of this is the ability of the thyroid gland to accumulate iodide at up to 40 times the circulating plasma level.27,32 One would expect the iodide concentration in the plasma and the thyroid gland to be similar if only diffusion and ultrafiltration were operating. Aqueous humor exhibits increased ascorbate, lactate, and certain amino acid concentrations as compared with plasma, as a consequence of active secretion.1

Another way of testing for the presence of an active metabolic process is to apply specific metabolic inhibitors to the ciliary body and observe the effect on aqueous secretion. Ultrafiltration of fluid from the plasma to the posterior aqueous has been suggested to be responsible for approximately 70% of aqueous formation.27,33,34 However, the results of experiments where metabolic inhibitors have been used have shown conclusively that this is not the case. For example, systemic35 or intravitreal injection36 of ouabain (an inhibitor of the enzyme sodium potassium-activated adenosine triphosphatase [ATPase]—Na+/K+ ATPase) results in a decrease of up to 70% in aqueous formation, in a variety of species. The topical administration of vanadate (also a Na+/K+ ATPase inhibitor) lowers aqueous secretion in rabbits37,38 and monkeys.39 Ouabain as well as a number of other ion transport and channel blocking drugs significantly reduce aqueous humor formation in arterially perfused bovine eyes. Bumetanide (a specific inhibitor of Na-K-2Cl cotransport) and furosemide (a nonspecific anion transport inhibitor) reduce aqueous humor formation by 35% and 45% in bovine eyes in vitro. Similarly, 4,4'-diisothiocyanatostilbene-2,2'-disulfonic acid (DIDS, a probable inhibitor of the Cl-HCO3 exchanger, the Na-HCO3 cotransporter, and chloride channel) and 5-nitro-2-(3-phenylpropylamino)-benzoic acid (NPPB, a chloride channel blocker in nonpigmented cells) reduce aqueous humor formation by 55% and 25%, respectively, in bovine eyes in vitro.40 Also, catecholamines such as epinephrine, norepinephrine, isoproterenol and dopamine, that stimulate aqueous humor formation in humans also stimulate Na-K-Cl cotransport.41 If the greatest proportion of aqueous secretion was attributable to ultrafiltration, then this would not occur. In addition, Bill22 noted that the hydrostatic and oncotic forces that exist across the ciliary epithelium–posterior aqueous interface would favor resorption, not secretion, of aqueous humor. The ciliary process stroma has an oncotic pressure of approximately 14 mm Hg, because of its protein content. Because IOP in the healthy eye is maintained at approximately 15 mm Hg, a capillary hydrostatic pressure of greater than 29 mm Hg would be required to drive an ultrafiltrate. Capillary hydrostatic pressure in the ciliary stroma has been estimated to be 25 to 33 mm Hg.14 It has been calculated27 that a capillary pressure of greater than 50 mm Hg would be necessary to promote ultrafiltration as the major mechanism for the secretion of aqueous. The values of capillary hydrostatic pressure in the ciliary processes and a consideration of the hydrostatic and oncotic forces involved do not favor ultrafiltration as an important mechanism for aqueous humor secretion. Acetazolamide, and other specific inhibitors of the enzyme carbonic anhydrase, decrease the formation of aqueous by 40% to 60%.42–44 A reduction in temperature, which inhibits most active metabolic processes, also results in a decrease in aqueous formation, to a greater extent than would be expected if only diffusion were operating.1,45

Active transport systems usually exhibit a limit beyond which an increase in substrate concentration produces no further increase in transport. When this limit is reached, the system is said to be saturated. Thus, the fact that a transport mechanism for a substance can be saturated provides evidence of an active system. It has been demonstrated that by increasing the ascorbate concentration in plasma, a level of plasma ascorbate is reached above which no further increase in aqueous ascorbate concentration will occur. This provides evidence that the ascorbate transport system in the eye is saturable.

Several membrane active transport systems have been identified in the ciliary epithelium that are known to pump various substances against a concentration gradient, including Na+/K+ ATPase and amino acid membrane transporters (Fig. 8). There are also passive transport proteins specific for HCO3 and Cl. Membrane-bound Na+/K+ ATPase is one important system involved in Na+ and K+ transport and is present mainly along the lateral cellular interdigitations of the nonpigmented ciliary epithelium.35,46–51 The transmembrane Na+/K+ transport protein is energized by the Gibbs free energy of hydrolysis of ATP, mediated by the ATPase enzyme intimately associated with it in the membrane. The ATP required for this process is derived predominantly from oxidative metabolism of glucose via the Krebs' citric acid cycle. It is most likely that Na+ is pumped across the cell membrane and Cl is passively carried with it in order to maintain electrical neutrality, although the relative rate of transport of Cl is unclear.52 The Na+ pump induces a potential difference across the ciliary body, ranging from 6 to 10 mV. Measurements of the potential difference across the ciliary epithelia indicate that the aqueous is positive with respect to the stroma. The magnitude of this potential difference is reduced after poisoning with ouabain.53 These data are consistent with the hypothesis that Na+ is the primary mover and that active transport of Cl is probably small in comparison to that of Na+.54 Interspecies differences in the aqueous-plasma ratios of Cl (e.g., high in human, low in rabbits) might be explained by the relative proportions of Cl actively transported.55 Transepithelial electrical measurements in the isolated rabbit iris-ciliary body indicate that Na+/K+ ATPase and HCO3 are required for active ion transport.56

Fig. 8. Diagram of possible secretory pathways in the ciliary processes. AA, ascorbic acid; CA carbonic anhydrase. (From Wiederholt M, Helbig H, Korbmacher C. Ion transport across the ciliary epithelium: Lessons from cultured cells and proposed role of the carbonic anhydrase. In Carbonic Anhydrase. Basel: Verlag-Chemie, 1991:232–244, with permission.)

However, recent studies suggest that active transport of Cl ions plays an equally important role 40,41,57 providing the driving force for aqueous humor formation.57 Shahidullah et al40 found that in the isolated bovine eye, aqueous humor is formed mostly by processes involving active secretion and chloride transport.

Recently, it has been discovered that water transport across many membranous barriers, including those in the eye, are facilitated by aquaporin (AQP) water channels.58 The AQPs are a family of water channels expressed in animals, plants, and lower organisms.58 There are at least 11 mammalian AQPs (AQP0 to AQP10) each of which is a small membrane protein approximately 30 kd in size. The eye expresses several AQPs at sites of fluid transport. AQP158,59 and AQP460 are expressed in the ciliary epithelium. AQP1-null mice have decreased IOP and aqueous humor production compared to normals.58

Histochemical studies have demonstrated a number of active enzyme systems in the ciliary epithelia. These include nucleotide phosphatases (especially ATPase, as mentioned previously), adenylate cyclase, and carbonic anhydrase, the enzyme that forms HCO3 (the body's alkaline ion) from CO2 and H2O61–65 by the following reaction:

CO2 + H2O → H2CO3 → H+ + HCO3

                I           II

Carbonic anhydrase catalyzes step I of the reaction and step II is an almost instantaneous ionic dissociation.66,67 The enzyme is localized in both the pigmented and the nonpigmented epithelial cells of the rabbit ciliary processes,68,69 and monkey and human pigmented and nonpigmented epithelial cells,70 most prominently in the basal and lateral membranes, but also in the cytoplasm. A variety of carbonic anhydrase inhibitors will reduce aqueous secretion, including acetazolamide,71 methazolamide, methoxazolamide, ethoxzolamide, dichlorphenamide,72,73 aminozolamide,74 trifluormethazolamide,75 6-hydroxyethoxzolamide,76 dorzolamide,77–79 brinzolamide,80–82 and several biscarbonylamidothiadiazole compounds.44 The site of action of these drugs is intraocular rather than systemic, because the unilateral injection of acetazolamide into the carotid artery lowers the IOP of only the ipsilateral feline eye.83 Precisely how carbonic anhydrase helps mediate aqueous secretion has been a long-standing debate. However, in recent years many advances in understanding of this system have been achieved. In the past, it was suggested that acetazolamide caused constriction of afferent iris or ciliary arteries, thereby decreasing ultrafiltration and hence aqueous humor formation.84,85 However, this view is no longer held. For example, Bill,86 using radiolabeled microspheres, concluded that acetazolamide had no significant effect on the blood flow in the intraocular tissues in the rabbit eye, and that it was unlikely that the drug had any effect on blood flow in the anterior uvea. It has now been irrefutably demonstrated that acetazolamide exerts its ocular effects via a reduction in aqueous secretion, mediated by a direct effect on the ciliary epithelial cells.26 The ciliary carbonic anhydrase system is now thought to function as described by Maren.87 Inhibition of the enzyme decreases the rate of Na+ and HCO3 transport into the posterior chamber by equimolar amounts, indicating a linkage of the accession of these two solutes into the posterior chamber of dogs and monkeys.88 Furthermore, inhibition of carbonic anhydrase lowers the aqueous Cl concentration in primates and the HCO3 concentration in rabbits.24,89 The mechanism by which carbonic anhydrase activity is coupled to Na+ and HCO3 movement into the posterior chamber is still subject to much study. However, it is now known that the primary reaction, which occurs within the cytosol of the nonpigmented ciliary epithelial cells, is the proteolysis of water to yield OH and H+ ions (Fig. 9). The hydroxide ions so formed react with CO2 catalytically (or noncatalytically when the enzyme is inhibited) to form HCO3, which is passively transported to the aqueous humor, simultaneously with the active transport of Na+. The protons liberated from water pass into the blood circulation where they are buffered by proteins. The theoretical rates of flow of HCO3, based on known constants and some assumptions regarding cell volume and pH of the secretory region, are shown at the bottom of Figure 9. The catalytic rate (i.e., when active carbonic anhydrase is present) is 7 μmol/min, and the uncatalyzed is 0.07 μmol/min.

Fig. 9. Model for HCO3 formation and its linkage to Na+ transport in the ciliary processes. Observed rates are for monkey. Chemical rate constants (37°C) taken from the work of Maren. (From Maren TH: The kinetics of HCO3 synthesis related to fluid secretion, pH control, and CO2 elimination. Annu Rev Physiol 50:695, 1988, with permission.)

The uncatalyzed rate is found experimentally by inhibiting the enzyme in vivo by a large dose of acetazolamide or similar compound. Such inhibition, when complete, should yield the uncatalyzed rate of HCO3 formation. Measurements obtained experimentally indicate that the observed inhibited rate (Fig. 9; 0.024 μmol/min) is not different from the calculated uncatalyzed rate of 0.07 μmol/min. Considering the assumptions about pH and cell volume, the difference between the calculated and experimentally measured inhibited rates of bicarbonate flow lies within acceptable limits of error. Inhibition of the flow of bicarbonate also leads to an inhibition of the flow of Na+. Several hypotheses have been put forward to explain this: (i) inhibition of carbonic anhydrase causes a decrease in HCO3 available for movement with Na+ to the aqueous side to maintain electroneutrality; (ii) a reduction in intracellular pH may inhibit Na+-K+ ATPase; and (iii) decreased availability of H+ produced by the reaction catalyzed by carbonic anhydrase decreases H+/Na+ exchange and reduces the availability of intracellular Na+ for transport into the intercellular channel.

A large excess of carbonic anhydrase has been found in the ciliary processes of all species studied. This has pharmacologic implications when one hopes to reduce aqueous secretion by inhibition of the enzyme. In order to achieve a clinically useful reduction in secretion and hence IOP, more than 99% of the enzyme must be inhibited. The sulfonamide carbonic anhydrase inhibitors have been in clinical use for glaucoma since 1955, when their action on the eye was discovered as an offshoot of their development as diuretics.73 They have also been extremely useful in research on aqueous humor formation, because they are highly specific, having no other actions at concentrations below 1 μM. One major compound is acetazolamide but many ophthalmologists have used methazolamide or related compounds, which have a greater diffusibility than acetazolamide, and that act on the eye in lower doses, thereby achieving the desired ocular effect but having lesser action on other systems in the body where carbonic anhydrase is present, such as the secretory epithelia of the kidney.42 Nonetheless, the systemic dose of sulfonamide carbonic anhydrase inhibitors necessary to significantly suppress aqueous secretion leads to undesirable side effects, even with methazolamide.90

In recent years topically applied carbonic anhydrase inhibitors have been developed for clinical use. Dorzolamide has been used since 199580,91,92 and brinzolamide since 1998.80 As a rule, the topically active carbonic anhydrase inhibitors have high lipid and water solubility, as well as high inhibitory potency and efficacy. Sufficient concentration is achieved in the ciliary processes via transcorneal, aqueous humor and iris absorption after instillation of a single drop.43,92,93 This can result in essentially as large a decrease in IOP as with the sulfonamides91–93 without the adverse side effects of oral acetazolamide. Dorzolamide has a high affinity for carbonic anhydrase and penetrates the eye well. However, some studies have found it to be a less effective at lowering IOP than oral acetazolamide.94 Brinzolamide's effectiveness equals that of dorzolamide.80,81,95,96 However, both produce localized, albeit mild, side effects such as burning and stinging upon instillation and blurred vision.95,96 Brinzolamide produces less ocular discomfort than dorzolamide80,81,95,96 possibly due to the more physiologic pH of the preparation.80

In summary, aqueous humor is mainly a product of active cellular secretion requiring metabolic energy. Na+ and K+ are transported from plasma to the posterior chamber, with HCO3, Cl and water following passively to maintain electrical and osmotic balance. Other substances such as certain amino acids are also actively transported. Some other small molecules probably appear in the aqueous via diffusion or ultrafiltration.

The process begins with an ultrafiltrate of plasma in the extravascular spaces of the ciliary stroma. The ciliary epithelium (probably nonpigmented) then incorporates certain molecules and ions selectively for concentration and direct active secretion into the posterior chamber, while other substances are secreted passively.

Back to Top
BLOOD–AQUEOUS BARRIER
Large molecules such as proteins are present in the aqueous in only small quantities, even if their respective concentrations in the plasma are raised to high levels. In humans, normal plasma total protein levels are 6 g per 100 mL, compared to less than 20 mg per 100 mL in the aqueous, less than 0.5% that of plasma.97,98 Thus, a restraint in the free passage of many solutes from the blood vessels of the ciliary stroma into the aqueous humor exists. This constitutes one component of the blood–aqueous barrier,1,99 the anatomic correlate of which are the tight junctions between the interdigitating surfaces of the nonpigmented ciliary epithelial cells. Bill100,101 demonstrated that albumin and other proteins pass through the ciliary capillary walls at a much faster rate than they enter the aqueous humor and concluded that a barrier to these substances existed at a site other than the vessel walls. The intravenous injection of horseradish peroxidase into the monkey leads to a rapid filling of the ciliary stroma by the enzyme.102 The enzyme passes through the many fenestrations present in the blood capillaries of the ciliary body, since these have few tight junctions (similar to choroidal capillaries), and also fills the spaces between the pigmented cells. However, it is prevented from entering the aqueous by the tight junctions of the nonpigmented layer, between the cell apices. Comparable results have been reported in the mouse103,104 and in the rabbit.105,106 Numerous other studies have tested the blood–aqueous barrier using a number of different intravascular tracers designed to mimic the behavior of plasma-proteins.2 The conclusion of these studies is that tight junctions between the apicolateral surfaces of the nonpigmented epithelium of the ciliary body, and between the endothelial cells of the iris vasculature, prevent the passage of plasma proteins into the aqueous humor.102–105,107–111

The blood-aqueous barrier is not absolute. Water-soluble substances of medium molecular weight such as urea, creatinine, and certain sugars may penetrate at varying rates, but all are slower than their transit across capillary walls. Generally, the greater the lipid solubility coefficient of a substance, the greater its ability to penetrate the blood-aqueous barrier and pass to the posterior aqueous chamber.1,112–114 In addition, regional differences of permeability in the ciliary body have been noted. For instance, the epithelia of the anterior portions of the ciliary processes of the rabbit are less permeable than the epithelia of the pars plana.105

Substances such as mannitol are used clinically to reduce IOP. They function as hyperosmotic agents, by exploitation of the fact that they penetrate the blood–aqueous barrier only poorly but are distributed widely within the extracellular spaces of the body. Water is drawn from cells and the ocular fluids, balancing the high osmotic pressure induced in the extracellular space via the high concentrations of mannitol. The resulting loss of water from the eye leads to a reduction in the IOP.1 The effect of hyperosmotic agents is most pronounced in eyes exhibiting pathologically elevated IOP.115

Other examples of hyperosmotic agents include glycerin, urea, isosorbide, and ethanol. Urea was the first substance to be used for this purpose, however, it will slowly penetrate the blood–aqueous barrier, and hence the hyperosmotic effect of urea on the eye is shorter lived than that of mannitol. Ethanol is not used clinically because it penetrates the eye even more rapidly than urea and in the required doses has undesirable effects on the sensorium.

Certain antibiotics (e.g., chloramphenicol, cephalothin, and ampicillin) are known to penetrate the blood-aqueous barrier well, whereas others do so only poorly (e.g., penicillin, methicillin, erythromycin, and gentamicin).

A similar barrier to the passage of solutes exists in the retinal pigment epithelium, thus forming a blood–vitreous barrier between the vitreous and the blood capillaries of the choroid. Blood capillaries with tight junctions in their endothelia form further physiologic barriers (functionally similar to the blood–brain barrier) in the iris and retina.

Back to Top
SECONDARY AQUEOUS
The blood–aqueous barrier is fragile and may be disturbed by a variety of noxious stimuli. A corneal abrasion, paracentesis (withdrawal of a small volume of aqueous via a needle inserted into the anterior chamber),116–120 intraocular infection, uveal inflammation, intraocular surgery, and certain drugs, applied topically (such as nitrogen mustard or an anticholinesterase, e.g., echothiopate, diisopropyl flurophosphate [DFP], demecarium, neostigmine, or physostigmine), or delivered by intracarotid infusion, such as hyperosmotic agents (causing separation of the ciliary epithelial layers, opening of the blood–aqueous barrier, and severe, permanent damage to the pigmented epithelial cells121), are all capable of breaking down the blood-aqueous barrier and inducing changes in the aqueous humor composition (Table 2). Disruption of the blood–aqueous barrier has also been reported in the contralateral eyes of patients who have had cataract extraction and lens implantation surgery,122 and after argon laser trabeculoplasty.123 Severe damage to the blood-aqueous barrier occurs with cyclodestructive procedures used to treat advanced glaucoma and is evidenced by the prolonged or chronic presence of flare (the scattering of light upon slit-lamp examination by increased levels of protein in the anterior chamber).

TABLE 2. Factors Interrupting the Blood-Aqueous Barrier

  1. Traumatic
    1. Mechanical
      1. Paracentesis
      2. Corneal abrasion
      3. Blunt trauma
      4. Intraocular surgery
      5. Stroking of the iris
    2. Physical
      1. X-ray
      2. Nuclear radiation
    3. Chemical
      1. Alkali
      2. Irritants (e.g., nitrogen mustard)
  2. Pathophysiologic
    1. Vasodilation
      1. Histamine
      2. Sympathectomy
    2. Corneal and intraocular infections
    3. Intraocular inflammation
    4. Prostaglandins
    5. Anterior segment ischemia
  3. Pharmacologic
    1. Melanocyte-stimulating hormone
    2. Nitrogen mustard
    3. Cholinergic drugs, especially cholinesterase inhibitors
    4. Plasma hyperosmolality
(Courtesy of RL Stamper, MD)

 

After breakdown of the blood–aqueous barrier, the resultant aqueous produced is known as secondary or plasmoid aqueous.1,24,30,124 The most notable change is a marked increase in protein concentration. In this situation, the ionic composition of the aqueous approaches that of a simple dialysate of plasma, and substances that are normally barred from entering the aqueous now do so with ease. The unusually rapid rate of entry of substances such as fluorescein, Evan's blue dye, albumin, or fibrinogen (which actually may allow the aqueous to coagulate) can be used as a diagnostic indicator of barrier breakdown.

Many of the processes that disrupt the blood–aqueous barrier also lead to vasodilation. Vasodilation can occur by means of an axon reflex (as may be seen after abrasion of the cornea), a sudden drop in IOP, or inflammation.125,126 In any case, vasodilation may be associated with loss of some of the tight junctions in the iridial vessels, which may help explain the breakdown of the barrier.125

In addition, it has been suggested that plasma proteins and other constituents can leak in a retrograde manner through Schlemm's canal, and, therefore, account for the breakdown of the barrier after paracentesis.127 In this study, no changes in the permeability of the ciliary epithelium or iris to horseradish peroxidase were found after paracentesis, contradicting previous studies and conventional views of the mechanism of barrier breakdown.

Release of prostaglandins causes vasodilation and many other findings associated with inflammation. Evidence points to the possible involvement of this class of compounds with breakdown of the blood-aqueous barrier.124,128–134 Prostaglandins have been implicated in the irritative response after mechanical trauma to the eye, and cause miosis, vasodilation, release of protein into the aqueous, and elevated IOP. Prostaglandins applied topically to the eye in sufficiently high concentration cause breakdown of the tight junctions of the nonpigmented ciliary epithelium and increase the protein content of the aqueous humor, the highest levels of protein being found in the posterior chamber.135,136

Pretreatment with inhibitors of prostaglandin synthesis such as indomethacin or aspirin will inhibit breakdown of the blood-aqueous barrier, as well as some of the other manifestations of inflammation ordinarily induced by the aforementioned process.137–141 It may well be that prostaglandin release represents a final common pathway for the action of many different kinds of trauma and irritants. However, small doses of particular prostaglandins (especially prostaglandin F [PGF] and certain congeners) applied topically to the eyes of cynomolgus monkeys and humans result in a large increase in uveoscleral outflow, with a concomitantly large decrease in IOP, without clinically evident ocular inflammation.132,133,142–147 In the past decade, PGF analogues such as latanoprost,148–153 travoprost, and the prostamide, bimatoprost, have been approved for clinical use and are now some of the most commonly used drugs in the treatment of primary open-angle glaucoma (POAG) and ocular hypertension (see uveoscleral outflow section for greater detail).154,155

Back to Top
AQUEOUS HUMOR COMPOSITION
In the healthy eye, the aqueous humor has a refractive index taken to be constant at a value of 1.336.156 Because this index of refraction is lower than that of the cornea, there is a slight divergence of light rays as they pass the cornea–aqueous interface. Both the viscosity and density of aqueous humor are slightly higher than that of pure water, while the osmolality is slightly higher than that of plasma.30,157–160

The volume of the human anterior chamber is approximately 200 μL,3 while that of the posterior chamber is approximately 60 μL.1 This makes chemical analysis of the aqueous difficult because of the tiny volume of fluid as well as the relatively poor accessibility of the posterior chamber. Furthermore, it may account for the differences in values for the concentrations of many substances obtained by different investigators.

The greatest difference between aqueous and plasma resides in the very low protein concentration in the aqueous (Table 3), which is in the region of 0.5% that of plasma.97,98,157–160 However, the composition of protein in the aqueous is also different from that in plasma. The ratio of the levels of the lower molecular weight plasma proteins (such as albumin and the γ-globulins) to the higher molecular weight proteins (such as the α-lipoproteins and the heavy immunoglobulins) is much higher in aqueous in the normal healthy eye than in plasma.161 However, when the blood–aqueous barrier breaks down (as in uveitis), the composition and concentrations of protein in the aqueous are similar to that of plasma.98 The levels of immunoglobulin in the aqueous have been determined, both in the normal eye and in eyes in which the barrier has broken down.162–164 In the healthy eye, immunoglobulin (Ig) G is present at a concentration of approximately 3 mg per 100 mL,164 while IgM, IgD, and IgA are absent, presumably because of their larger molecular structure. In eyes with uveitis, the concentration of IgG increases, and IgM and IgA appear also. In the normal aqueous, trace concentrations of active complement C2, C6, and C7 globulins are also present.165

TABLE 3. Aqueous and Serum Protein Concentration


ProteinHumanRabbit
Aqueous
(g/100 mL)
Serum
(g/100 mL)
Aqueous
(g/100 mL)
Serum
(g/100 mL)
Total protein0.0137.50.0505.6
Globulin0.0032.50.0252.5
Albumin0.01050.0253.1

(Adapted from Davson H: The Eye. In: Vegetative Physiology and Biochemistry, Vol 1, 2nd ed, New York, Academic Press, 1969; and Duke-Elder S: The Aqueous Humour. The Physiology of the Eye and of Vision, Vol 4, pp 104–200. In Duke-Elder S (ed): System of Ophthalmology. St. Louis, CV Mosby, 1968)

 

There are also trace quantities of several components of the fibrinolytic and coagulation system present in the aqueous, with the exception of plasminogen and plasminogen proactivator, which are present at more significant concentration. Only trace quantities of the inhibitors of plasminogen activation are present in the aqueous,166 thereby ensuring that the aqueous outflow pathways remain free of fibrin. The aqueous humor of diseased eyes incorporates significant quantities of all of the major components of coagulation and fibrinolysis, giving rise to formation of intracameral clots; however, the composition of these clots is different from those that occur in blood vessels.24

The α and γ lens crystallins are also present in only small amounts in the aqueous humor of healthy eyes, although the concentration of these proteins increases in cataract.167,168

It has been calculated that in the normal healthy eye, the blood–aqueous barrier behaves as a semiporous membrane with a pore radius of approximately 10.4 nm. The protein concentration in the peripheral portion of the anterior chamber, close to the meshwork, may be much higher than in the more central region because of protein entry directly from the peripheral iris, as demonstrated in monkey and human eyes.169–172 Inclusion of serum in the perfusand of monkey173 or bovine174 eyes decreases resistance washout. The reduction in washout may be the result of interactions of particular serum proteins and not due to the general level of serum proteins.175

The concentration of amino acids, on the other hand, is frequently higher in the aqueous than in the plasma.176 In the canine eye, however, there is a lower concentration of amino acids in the aqueous than in the plasma.177 At least three transport systems for amino acids have been proposed in the eye,55 one each for the acidic, basic, and neutral groups. A statistical study of the covariation of the concentration of amino acids and related compounds in human aqueous suggested the existence of six transport systems in the ciliary epithelia: three independent mechanisms for neutral amino acids, and independent mechanisms for basic amino acids, acidic amino acids, and urea.178

The distribution of ions varies greatly amongst different species. For example, the monkey has a higher concentration of H+ and Cl and a lower concentration of HCO3 compared to plasma. On the other hand, rabbit aqueous has a lower concentration of Cl and H+ and a higher concentration of HCO3 compared to plasma.179 Active transport of Cl across the feline isolated ciliary epithelium has been reported.52 The concentration of Na+ in the aqueous is almost the same as in the plasma in many species.1,30,180 However, the osmotic pressure of aqueous is slightly higher than that of plasma with respect to Na+, because of Gibbs-Donnan equilibrium.

Most species tested have very high concentrations of ascorbate and lactate in the aqueous. Ascorbate is actively secreted into the posterior chamber, and the secretion mechanism will only function in the presence of ATP and a Na+ gradient. The physiologic function of ascorbate remains to be elucidated; however, it is known to be concentrated by the lens epithelium181 and has been shown to have a protective effect against UV-induced DNA damage to lens epithelium.182 Ascorbate may function as an antioxidant, regulate the sol–gel balance of mucopolysaccharides in the trabecular meshwork, or partially absorb UV radiation183,184 because diurnal mammals have approximately 35 times the concentration of aqueous ascorbate than nocturnal mammals.185

The key oxidant in the aqueous humor, hydrogen peroxide, is normally present186 as a result of reactions of ascorbic acid and trace metals.187 Additional hydrogen peroxide and reactive oxygen species are generated by light-catalyzed reactions, metabolic pathways, and phagocytic or inflammatory processes.186,187 Hydrogen peroxide affects aqueous outflow in the calf perfusion system.188 Human trabecular meshwork cells exposed to 1 mmol of hydrogen peroxide show reduced adhesiveness to the extracellular matrix proteins fibronectin, laminin, and collagen types I and IV.189 Extensive and repeated oxidative stress in vivo may result in reduced TM cell adhesion, leading to cell loss that is identified as one of the major culprits in glaucomatous conditions.190–193

Lactate is produced as a result of the glycolytic degradation of glucose by both the ciliary body and the retina.194 It diffuses into the posterior chamber but is present at only marginally higher concentration than in the plasma at this site. However, it accumulates in the anterior chamber at considerably higher concentration than in the plasma.

Glucose, urea, and nonprotein nitrogen concentrations are slightly less than in plasma.1,30 Glucose is thought to diffuse into the aqueous, where its concentration is approximately 80% that of plasma. Glucose also diffuses into the cornea. Its concentration within the corneal endothelium is approximately half that in the aqueous. In diabetes mellitus, the aqueous concentration of glucose is increased. High glucose levels in the aqueous humor may increase fibronectin synthesis and accumulation in the trabecular meshwork and accelerate the depletion of trabecular meshwork cells. In bovine trabecular meshwork cells grown in high-glucose medium, fibronectin mRNA is significantly upregulated, fibronectin immunofluorescence is more intense, and relative amounts of fibronectin protein are significantly increased.195 Also the chemoattractant potential of fibronectin in aqueous humor is reported to play a role in trabecular meshwork cell loss in glaucoma.196

Oxygen is also present in the aqueous humor, at a tension determined to lie between 13 to 80 mm Hg, depending upon the method of measurement.197–201 The tension of oxygen in the aqueous can be decreased by topical epinephrine, possibly as a result of uveal vasoconstriction198 or by the wearing of polymethylmethacrylate (perspex) contact lenses, which by restricting the normal passage of oxygen across the corneal epithelium, cause corneal hypoxia and thus an increase in movement of oxygen from the aqueous, across the corneal endothelium.198

Berzelius202 first demonstrated the presence of lipids in bovine aqueous. Since then, sphingomyelin, phosphatidyl choline, and lysophosphatidyl choline have all been shown to be present in the aqueous,203 although at a concentration of less than 1 mg per 100 mL, because lipids are largely barred from entry to the aqueous by the blood–aqueous barrier.204

Other substances, such as corticosteroids,205 monoamine metabolites,206 Cr3+ ions,207 vitamin B12,208 sialic acid,209 and hyaluronic acid210 have all been found to exist in the aqueous of a number of different species. Hyaluronic acid covering the surfaces of the outflow pathways might prevent adherence of molecules to extracellular matrix components within the cribriform region and thereby prevent clogging of the outflow pathways.211 Knepper et al212 hypothesize that POAG is characterized by a decreased concentration of hyaluronic acid and increased turnover and downregulation of the hyaluronic acid receptor CD44 in the eye, which, in turn, may influence cell survival of TM and retinal ganglion cells. Corticosteroid regulation of IOP is proposed to occur via 11β-hydroxysteroid dehydrogenase (HSD)-1 expression that has been localized in the nonpigmented ciliary epithelium of human eyes.213–215 This enzyme catalyzes the conversion of cortisone to cortisol which, in turn, induces sodium and concomitant water transport into the posterior chamber, through epithelial sodium channels, including Na+-K+-ATPase,216,217 resulting in aqueous production. Levels of cortisol compared to cortisone in the aqueous humor are normally much greater than in the systemic circulation.218 However, long-term interactions of cortisol in the aqueous with glucocorticoid receptors in the trabecular meshwork, could contribute to increasing outflow resistance in individuals susceptible to steroid induced glaucoma (see section on trabecular outflow). Ingestion of the 11β-HSD antagonist carbenoxolone, decreased IOP in normal213 and ocular hypertensive214 human eyes. The role of the other compounds within the aqueous is not understood.

Transforming growth factor (TGF) β2 is a component of normal aqueous humor detected in many mammalian eyes219–222 and may play a role in glaucoma pathogenesis. The intrinsic activity of TGFβ2 is considered to be an important factor for the maintenance of the anterior chamber-associated immune deviation (ACAID).219,223 (see section on Aqueous Outflow: Other Agents for more details).

As the aqueous flows from the posterior chamber to the anterior chamber, changes occur. Nutrients diffuse into the lens, iris, and vitreous.224 Diffusion from the posterior aqueous into the vitreous is a major contributor to the existence of concentration gradients of low molecular weight substances in the vitreous.225 Waste products, such as lactate diffuse from the lens, iris, and corneal endothelium into the aqueous. Some exchanges of small molecules occur across the iridial vessels. Predictably, the chemical compositions of posterior chamber aqueous and anterior chamber aqueous are different.30 To complicate the matter further, some substances appear to be actively transported out of the eye. Paraaminohippurate (PAH), diodrast, and penicillin are examples of large anions that are actively transported out of the eye. The system appears to be similar to that occurring in the renal tubules. This active transport system can be saturated and can also be inhibited by low temperatures and probenecid.1,30 The nonpigmented ciliary epithelium has been implicated as the site of this active transport system within the anterior eye. Another independent transport system actively excretes injected iodide from the aqueous. This latter system resembles iodide transport mechanisms in the thyroid and salivary glands.226 The significance to the normal physiology of the eye of these outward-directed transport systems has not been established. With the discovery that prostaglandins may be actively transported out of the eye,227,228 some have suggested that such mechanisms may be useful to rid the eye of biologically active substances no longer needed, or which may even be detrimental.45 Their removal from the eye to the blood facilitates their excretion via the hepatic route. There are other outwardly directed ion-uptake mechanisms present in the eye. The anterior uvea of the rabbit eye, for example, accumulates the anions cholate, glycocholate, deoxycholate, chenodeoxy-cholate, iodipamide, and o-iodohippurate.229,230 At least one inwardly directed anion pump mechanism also operates—ascorbate is accumulated in the anterior chamber and lens, its concentration in the aqueous humor being approximately 20 times that in plasma.181 This may help protect the anterior ocular structures from oxidative damage. Bárány,231 making an analogy to the ion pump located at the renal peritubular cell border adjacent to the blood, which pumps simple cations from the blood to the kidney tubule, investigated whether there are any inwardly directed cation pump mechanisms from blood to aqueous. He concluded, however, that such mechanisms probably do not exist, at least within the rabbit eye. But outwardly directed cationic pump mechanisms have been reported. For example, iris–ciliary body preparations have been shown to accumulate the cation emepronium,232 although a later report233 questioned whether any other cations are actively eliminated from the eye.

Back to Top
AQUEOUS FLOW
The result of all of the biochemical processes previously discussed is the production of a quantity of fluid that circulates continuously. Many of the biochemical changes caused by diffusion occur as the fluid moves from the posterior chamber, through the pupil, around the anterior chamber, and into the outflow system. Superimposed on this bulk flow is the anterior chamber thermal circulation, which causes the aqueous closer to the cooler avascular posterior cornea (cooled additionally by evaporation of tears from the corneal epithelium) to move downward, whilst the aqueous closer to the warmer vascular anterior iris moves upward. In addition, movements of the eyes and head modify these flow parameters.

In the human eye, the rate of aqueous formation is approximately 2.5 μL/min, while that in the rabbit is approximately 3 to 4 μL/min. Because of the formation rate and aqueous chamber volumes in the human eye, approximately 3% of the posterior chamber volume and 1% of the anterior chamber volume are replaced per minute1,30 with the entire volume of the aqueous humor being replaced every 90 to 100 minutes.2 The more rapidly the aqueous is formed, the less is the potential for diffusional exchange with the ciliary processes, lens, anterior iris, and posterior cornea. However, within physiologic limits, changes in the rate of formation probably do not significantly affect the diffusional exchange. Furthermore, the rate of aqueous formation contributes to the regulation of IOP.

Back to Top
METHODS OF MEASURING RATES OF AQUEOUS FORMATION
In 1951, Goldmann234 first described a technique by which rate of flow of aqueous humor in the eye could be quantified. This technique was based on the measurement of the kinetics of unbound fluorescein in the plasma and concomitant fluorescence in the anterior chamber after intravenous injection.3 Thereafter, other investigators devised techniques for the determination of aqueous flow, generally involving a cannulation of the anterior chamber with a needle, thus permitting drainage of aqueous or infusion of a fluid at a known rate.235–241 For example, one can inject into the anterior chamber a known amount of dye, radioactively labeled substance, or large molecule, the concentration of which is easily measured. If the substance mixes rapidly with all of the aqueous, then repeated sampling of the fluid and measurement of the concentration provides an indirect measurement of bulk flow. The underlying assumption with this technique is that no change occurs in aqueous dynamics as a result of the injection or sampling technique. In practice such an assumption does not hold true, because the blood–aqueous barrier is breached by simple paracentesis.242 Furthermore, this technique assumes that none of the substance leaves the eye by diffusion or any other way except by direct bulk flow (which includes outflow via the uveoscleral pathway).

Later still, other techniques more applicable to the measurement of aqueous formation in the human eye were devised, which did not involve any invasive procedure.243–249 These methods, discussed in the following section, fall essentially into one of two categories: (i) measurement of the rate of appearance or disappearance of a chemical substance from the aqueous or (ii) derivation of the aqueous formation rate from a mathematical formula (to be discussed) after obtaining measurements of the IOP, episcleral venous pressure, and resistance to aqueous outflow.1,30 Each method incorporates intrinsic advantages and disadvantages, and specific sources of error. However, in spite of this, there exists good agreement for the values of aqueous formation obtained, and in several different species.

Back to Top
MEASUREMENT OF FORMATION BY THE RATE OF APPEARANCE OR DISAPPEARANCE OF A CHEMICAL SUBSTANCE FROM THE AQUEOUS

It is possible to inject a substance such as PAH or fluorescein into the bloodstream and maintain a high concentration there. Over a period of several hours, a small but measurable amount diffuses into the aqueous humor and reaches an equilibrated concentration. The injection is then stopped, and the level of the fluorescein or PAH in the blood subsequently declines rapidly, owing to renal clearance. The concentration of the substance in the aqueous, however, remains relatively constant over a short period of time, after which it starts to decrease, as fresh aqueous is formed not incorporating any PAH or fluorescein, because of low plasma levels. The rate of fresh aqueous formation can be calculated by measuring the declining concentration either by sampling the aqueous first in one eye, then later in the second eye, as in the case of PAH, or by optical means using fluorophotometry, as in the case of fluorescein (Fig. 10). Bárány and Kinsey250 developed this technique for PAH, and it was extended for use with fluorescein.248,251–253 The technique using fluorescein has been modified by dropping the fluorescein topically onto the eye rather than by injecting it.254,255 The dye in this case gains access to the anterior chamber via corneal absorption. There are, however, inherent problems with this approach, specifically, the problem of measurement of fluorescence in the anterior chamber and cornea, and the deduction of aqueous flow from the change in fluorescence over time.3 The first problem was solved by the construction of a slit-lamp fluorometer.256,257 The second was addressed by a number of investigators who devised several new experimental approaches.55,250,258 Thus, despite the drawbacks of fluorophotometry using topically applied fluorescein, the technique has become the gold standard in studies involving the human eye. Jones and Maurice248 realized that the corneal stroma could serve as a depot from which fluorescein could be introduced slowly into the anterior chamber. This method clarified the important role of the cornea in affecting the kinetics of topically applied drugs and tracers. Maurice's technique is now used most frequently for the measurement of the rate of aqueous formation in the human eye. The method involves the application of fluorescein topically to the eye, which subsequently penetrates the corneal epithelium and enters the stroma. Fluorescein is not metabolized by the eye and thus can disappear in only three ways: (i) rediffusion through the corneal epithelium and loss with the tears, (ii) lateral diffusion into the limbal tissue, or (iii) penetration of the endothelium and entrance into the aqueous humor, from where it is washed away by flowing aqueous, or is lost by diffusion into the iris (approximately 10% of fluorescein in the aqueous is lost by this route in the human eye).259 The third pathway of movement of fluorescein from the corneal stroma offers the least resistance, and thus is the major pathway of loss of fluorescein from the stroma.

Fig. 10. Principles of measurement of aqueous flow by ocular fluorophotometry. A: Optical axis of eye is scanned for background fluorescence with a scanning ocular fluorophotometer. B: Topical application of drops of fluorophore (2% fluorescein) applied to cornea. C: After a suitable delay (approximately 15 hours), to allow fluorescein to diffuse from the corneal depot to the aqueous humor, the eye is scanned once again. D: Repeated scans at 30-minute to 1-hour intervals over a 3- to 6-hour period facilitate monitoring of decline in fluorescence of aqueous humor with time. This can be related mathematically to aqueous flow rate (a calculation often performed by computer) after subtraction of background fluorescence and derivation of anterior chamber volume from keratometry and pachymetry determinations. The graphs to the right of the diagrams indicate typical fluorescence patterns obtained along the optical axis at each stage in the procedure. C, cornea; AC, anterior chamber; L, lens; V, vitreous.

The advantages of Maurice's technique are that it is safe, repeatable, and objective. Studies of the human eye can span 18 to 24 hours after application of a single dose of fluorescein. The procedure disturbs the eye minimally, and the subject need not be constrained during the interval of measurement. Furthermore, the technique is reliable even when the rate of flow is not constant.3

In one variation of this technique, fluorescein is introduced into the stroma simply by applying drops topically to the inferior fornix of the conjunctiva.254,260 A waiting period of 6 hours or more allows the dye to become more uniformly distributed within the stroma. In another variation, fluorescein is applied by iontophoresis.248 This procedure involves the introduction of fluorescein into the anterior chamber by forcing it through the cornea via the application of a small electric current. Fluorescence is measured in the stroma and in the anterior chamber at the beginning and end of an interval. Flow is the clearance of the dye during the interval minus the diffusional loss. The technique, however, is not applicable to eyes that lack an iridolenticular barrier between the anterior and posterior chambers. Furthermore, the technique measures only that portion of secreted aqueous that passes into the anterior chamber.3

McLaren261 developed a technique of measurement of aqueous flow based on flare. Flare (pathologic scattering of light resulting from the presence of protein in the aqueous resulting from inflammation and breakdown of the blood–aqueous barrier) was induced by argon laser photocoagulation of the iris in rabbits, and a scanning ocular spectrofluorophotometer was used to measure scattering in the anterior chamber. This method was used to study changes in aqueous flow over the diurnal cycle. A technique for the measurement of aqueous flow using corneal and vitreous depots of fluorescein in the rabbit eye has also been described.262,263 It may be concluded that (i) movement of water into or out of the vitreous can cause large changes in the rate of movement of dye from the vitreous to the anterior chamber and can make interpretation of the vitreous method ambiguous and (ii) the vitreous method is probably superior for measuring sustained changes of the rate of aqueous flow over at least 10 hours, or perhaps several days, but it cannot be reliably used for measuring changes over shorter periods.

Similar types of studies have been performed using iodide.264 In all cases, it is assumed that the amount of substance leaving the eye by alternative routes is negligible compared with the dilution by fresh aqueous.

By using the method of Maurice and introducing fluorescein by iontophoresis, a value of 2.48 ± 0.17 μL/min (mean ± standard error of the mean [SEM]) was calculated for aqueous flow in the human eye.248 Holm245 carried this method one step further. After forcing fluorescein into the anterior chamber, the fluorescein is carefully mixed by rapid eye movements. Then, as fresh aqueous humor not containing fluoresecein is produced, it forms a small, clear, growing bubble at the pupillary border. Multiple slit images are projected onto the bubble and photographed. The volume of the bubble can be estimated by measuring the deviation of each slit as it passes over the bubble and integrating the area of all the deviations. Two sets of photographs are taken 15 seconds apart in order to calculate the rate of change. However, the pupil must be miotic (constricted) for this technique to work, and a miosis is induced by topical application of pilocarpine, which may stimulate aqueous flow slightly.265 Furthermore, the production and maintenance of the bubble of clear aqueous is technically difficult. For these reasons, Holm's technique of iontophoresis has not gained widespread popularity in studies of aqueous humor secretion, either in the human or animal eye, although the results that have been obtained using this method agree closely with those obtained using Maurice's technique of fluorophotometry.3

In all the techniques whereby a fluorescent dye is used to measure aqueous flow, there are other potential confounding factors, such as binding of a proportion of the dye to ocular and/or plasma proteins, that must be considered.266 However, under normal circumstances these factors do not compromise the measurements.

Back to Top
MATHEMATICAL DERIVATION OF AQUEOUS FORMATION RATE
The other major types of investigational tools used to estimate aqueous formation involve physical measurements. These methods depend on the following algebraic manipulation of the modified Goldmann equation describing IOP in terms of episcleral venous pressure (Pe), aqueous flow (Fin), trabecular outflow facility (Ctrab), and uveoscleral outflow (Fu):
IOP = Pe + ((Fin − Fu)/Ctrab))

thus

IOP − Pe = ((Fin − Fu)/Ctrab))

and

Fin − Fu = Ctrab (IOP − Pe)

therefore

Fin = Ctrab (IOP − Pe) + Fu

In order to determine Ctrab and Fu, the eye in a living anesthetized animal may be cannulated with a needle and perfused with a solution of mock aqueous humor267 incorporating radiolabeled albumin (see later in this chapter).268,269 A value for IOP can be obtained by tonometry, or (in experimental animals) cannulation of the anterior chamber with connection via tubing to a pressure transducer. An estimate of Pe can be obtained in experimental animals by direct cannulation of an episcleral vein, or by measurement of the pressure necessary to completely collapse an episcleral vein.270 (In practice these procedures would be repeated several times and a mean value calculated, because Pe is not constant throughout the entire episcleral venous system). By substitution into the above equation, Fin may be determined. Fin can also be calculated by determination of the difference in radioactivity of a perfused radiolabeled solution before entering and after leaving the anterior chamber, the difference being related proportionally to total aqueous flow. These approaches, however, are applicable only to experimental animals; their invasive nature makes them unsuitable for use in the human eye.

Another method for estimating the rate of aqueous flow involves the use of a perilimbal suction cup.85,271 The perilimbal suction, theoretically, occludes the outflow channels of the eye, and thus IOP rises. The rate of increase in IOP depends not only on the rate of formation of aqueous humor but also on the elasticity or distensibility of the eye. Unfortunately, there is a question as to whether this method may also alter the blood flow to the ciliary body, thus possibly having an effect on aqueous formation. Furthermore, as the pressure rises in the eye, the rate of inflow of aqueous humor declines, a phenomenon known as pseudofacility. In addition, the suction cup may not totally close off the drainage of aqueous. Nevertheless, Galin,271 using this method, calculated rates of aqueous formation that compare well with other methods.

It is apparent that each of the foregoing methods has inherent errors and assumptions, yet despite this, the results indicate reasonable agreement among the various methods. The study of these methods has added much to the understanding of the factors that influence aqueous formation. In fact, it is not so much the absolute values themselves that are important but, rather the ability to compare changes in values for aqueous formation under different conditions that has been most informative.

Of all of these methods, however, noninvasive fluorophotometry has become the method of choice for the determination of the rate of flow of aqueous.

Back to Top
FACTORS AFFECTING AQUEOUS FORMATION
The rate of formation of aqueous is not constant. Variations in the formation rate occur hourly and daily.3 Knowledge concerning the cause(s) of this variation is particularly sparse. It is likely that many diverse physiologic systems, including the central nervous, endocrine, and cardiovascular systems, as well as changes in metabolic activity, all influence the production of aqueous.

Aqueous secretion also exhibits a diurnal cycle.1,24,272 It is hypothesized that the diurnal cycle of aqueous humor formation is regulated in part by a factors such as circulating catecholamines, epinephrine and norepinephrine, that have a circadian rhythm and partly by a factor that depends on the activity of the subject.272 Further, the magnitude of the fluctuation in flow rate differs among different individuals.

The ciliary body is innervated by nerves arising from the long posterior and short ciliary nerves, which run parallel to the long posterior and short ciliary arteries. These nerve fibers are of both the myelinated and nonmyelinated variety. Parasympathetic fibers originate in the Edinger-Westphal nucleus of the third cranial nerve, run with the inferior division of this nerve in the orbit, and synapse in the ciliary ganglion.225 Sympathetic fibers synapse in the superior cervical ganglion and are distributed to the muscles and blood vessels of the ciliary body. Numerous unmyelinated nerve fibers surround the stromal vessels of the ciliary processes; these are most likely noradrenergic and subserve vasomotion.24 Sensory fibers arise from the ophthalmic division of the trigeminal nerve and enter the ciliary body, but their distribution and function have not been well studied. However, despite the distribution of nerves to the ciliary body, little evidence for innervation of the ciliary epithelium itself has yet been found.4 Wetzel and Eldred273 found that some of the dendritic processes from peptidergic amacrine cells in the retina of the turtle formed a dense circumferentially oriented nerve fiber plexus and that collaterals from this plexus projected into and innervated the nonpigmented ciliary epithelium in the pars plana region of the ciliary body. The authors suggested that these peptidergic amacrine cells may be involved in the control of aqueous inflow.273 It has also been found that stimulation of the ciliary ganglion or application of cholinergic agents, in nonprimates, will lead to an increase in aqueous formation, even in the isolated, perfused feline eye.85 Additionally, in this model, sympathetic activity was reported to suppress aqueous secretion. Also, in the bovine perfused eye model,274 delivery of the adrenergic agonist terbutaline via the intra-arterial route to the ciliary body was found to decrease the rate of aqueous secretion.275

Results obtained with isolated perfused eyes of nonprimate animals, however, may bear little relationship to the in vivo physiologic situation reported in humans and subhuman primates. For example, adrenergic agonists are reported almost exclusively to increase aqueous secretion in primates. Topical epinephrine, norepinephrine, or isoproterenol stimulate secretion in the monkey eye,276,277 and topical epinephrine or terbutaline stimulate aqueous secretion in the human eye.278,279 Whether these processes work at the vascular level to change vascular tone or permeability or their effect is dependent on ciliary epithelial cell surface neurohumoral receptors associated with adenylate cyclase and/or guanylate cyclase activity has not yet been completely elucidated; however an amassing body of evidence points to the latter.275,280–296

The α2-adrenoceptor agonists apraclonidine and brimonidine are reported to lower aqueous secretion and IOP in cats, rabbits, monkeys,297–299 and humans.300 The effect is thought to be mediated via a postjunctional or nonjunctional α2 receptor present on the membrane of the nonpigmented ciliary epithelium299 but may in part be centrally mediated via imidazoline receptors in some species.301 The effect in monkeys is not abolished by superior cervical ganglionectomy,299 indicating that intact sympathetic innervation is not required for the drugs to lower secretion and IOP. However, intact sympathetic innervation is required for a response to the α2-adrenoceptor agonists clonidine in rabbits302 or brimonidine in cats and rabbits.297 The lack of aqueous flow suppression by brimonidine in pentobarbital-anesthetized monkeys may reflect the suppression of neural and humoral catecholaminergic tone by this anesthetic,303 especially relative to ketamine, which elevates such tone.304 Timolol, a nonselective β-adrenergic antagonist, suppresses aqueous flow in both eyes of ketamine305 but not pentobarbital-anesthetized306 monkeys that have undergone unilateral superior cervical ganglionectomy, suggesting that the relevant β-receptors may be nonjunctional rather than postjunctional and that sympathetic neural tone to the ciliary body may play relatively little role in regulating aqueous humor formation in the monkey. It is thus not entirely clear to what extent aqueous secretion is under neuronal versus humoral adrenergic control.

The effects of cholinergic drugs on aqueous humor formation and composition, and on the blood–aqueous barrier, are unclear, with conflicting results arising from various studies. In general, cholinergic drugs cause vasodilation307–309 resulting in increased blood flow to the iris, ciliary processes and ciliary muscle.309,310 The presence of flare (Tyndall effect, indicating increased protein concentration) and/or the detection of cells in the aqueous humor by biomicroscopy indicates that these agents can also cause breakdown in the blood-aqueous barrier.311 Pilocarpine increases barrier permeability to iodide264 and inulin.312 Cholinergic drugs may alter the aqueous humor concentration of inorganic ions313 and the movement of certain amino acids from the blood into the aqueous humor, and may also influence the outward-directed transport systems of the ciliary processes.314,315 Under certain conditions, pilocarpine may increase pseudofacility.316,317 Cholinergic agents or parasympathetic nerve stimulation have been reported to increase, decrease, or not alter aqueous humor formation rate and to slightly increase the episcleral venous pressure.265,317–324 The minimal effect on the rate of aqueous humor formation and episcleral venous pressure are clearly not responsible for the drug-induced decrease in IOP that results from cholinergic therapy.

Aqueous formation varies directly with the blood pressure in the internal carotid-ophthalmic arterial system in the primate277 but only when blood pressure is altered artificially to a physiologically abnormal extent. For example, ligation of the internal carotid artery causes a profound drop in aqueous secretion. Formation is not significantly altered however by changes in blood pressure within the normal physiologic range for any given species. Aqueous formation rate diminishes slightly with age. After the age of approximately 10 years, formation declines by 3.2% to 3.5% per decade.3 The mechanism underlying this decrease is unknown. An age-dependent loss of ciliary epithelial cells has not been described. Some authors have demonstrated a 5.8% loss of trabecular cells per decade in humans,190,325,326 while others have reported a loss of corneal endothelial cells at the rate of 3.5% per decade.327 Brubaker3 suggested that the age dependency of the population of ciliary epithelial cells should be studied, because if it is found that aqueous formation parallels the number of secreting cells, it would suggest that although aqueous formation may depend on neuronal or hormonal stimulation, the normal rate of formation may depend on cell count. Alternatively, the decline in aqueous formation could be a result of the changes observed in the fine structure of aging ciliary epithelial cells.328 Hypothermia leads to a decrease in aqueous formation, for example, a drop in body temperature of 7°C (about 19%) leads to approximately a 50% reduction in secretion, reflecting the deactivation of metabolic processes necessary to maintain active secretion.45

The ultrafiltration component of aqueous humor formation is pressure-sensitive, decreasing with increasing IOP. This phenomenon is quantifiable and is termed pseudofacility, because a pressure-sensitive decrease in inflow appears as an increase in outflow facility when techniques such as tonography and constant pressure perfusion are used to measure outflow facility.237,238,316,329–333 Although some sensory nerve endings exist in the ciliary body, they do not appear to be of the pressure-sensitive variety.4 The initiating event of this pressure-induced response of the ciliary processes remains obscure. Bill and Bárány237 reported that an artificially induced rise in IOP caused a reduction in secretion. This has been confirmed in the monkey.334 An intracameral injection of erythrocytes yielded a partial blockade of the trabecular meshwork, and hence an elevation in IOP. This led to a suppression of aqueous secretion, corresponding to 0.06 μL/mm Hg per minute increase in pressure. Therefore with a normal secretion rate in the pentobarbital anesthetized cynomolgus monkey of approximately 1 μL/min,335 an increase in IOP of 20 mm Hg should theoretically suppress aqueous secretion altogether. However, this does not in fact occur. It is well established that secretion continues, even against very high pressures. Increased IOP does decrease the blood flow in the ciliary body and may decrease secretion in this manner.315,330,336 In addition to the acute effect of elevated IOP, some patients in the late stages of glaucoma may show hyposecretion and even have normal pressures despite almost totally occluded outflow channels.337 It is not known if this is due to the same mechanism as the acute type of pressure-related hyposecretion. There is evidence, however, that much of the perceived reduction in aqueous secretion in response to increased IOP is caused by measurement artifact237,238,268,315,334,338 and that the real magnitude of pseudofacility in the monkey is less than 0.02 μL/mm Hg per minute, or less than 5% of total facility.238,339

A decrease in aqueous secretion occurs in association with uveitis, especially uveitis involving the ciliary body epithelium (iridocyclitis). A reduction in IOP is also seen clinically and experimentally in the monkey eye, perhaps mediated by prostaglandin release,340 and probably caused mainly by an increase in uveoscleral outflow rather than to a reduction in aqueous secretion.340,341 Other clinical conditions associated with decreased aqueous production include retinal, choroidal, or ciliary body detachment (Table 4).

TABLE 4. Factors Causing Reduced Aqueous Secretion

  1. General
    1. Age
    2. Diurnal cycle
    3. Exercise
  2. Systemic
    1. Reduction in blood pressure
    2. Artificial reduction in internal carotid arterial blood flow
    3. Diencephalic stimulation
    4. Hypothermia
    5. Acidosis
    6. General anesthesia
  3. Local
    1. Increased IOP (pseudofacility)
    2. Uveitis (especially iridocyclitis)
    3. Retinal detachment
    4. Retrobulbar anesthesia
    5. Choroidal detachment
  4. Pharmacologic
    1. β-adrenoreceptor antagonists (e.g., timolol, betaxolol, levobunolol, carteolol, metipranolol)
    2. Carbonic anhydrase inhibitors
    3. Nitrovasodilators; atrial natriuretic factor
    4. Calcium channel antagonists
    5. 5-HT1A antagonists (e.g., ketanserin)
    6. DA2 antagonists (e.g., pergolide, lergotrile, bromocriptine)
    7. α1-adrenoceptor antagonists (e.g., prazosin, phentolamine)
    8. α2-adrenoceptor agonists (e.g., apraclonidine, brimonidine)
    9. ACE inhibitors
    10. H1 receptor antagonists (e.g., antazoline, pyrilamine)
    11. Δ9-tetrahydrocannabinol (Δ9-THC)
    12. Metabolic inhibitors (e.g., DNP, fluoracetamide)
    13. Cardiac glycosides (e.g., ouabain, digoxin)
    14. Spironolactone
    15. Plasma hyperosmolality
    16. Cyclic GMP
  5. Surgical
    1. Cyclodialysis
    2. Cyclocryothermy
    3. Cyclodiathermy
    4. Cyclophotocoagulation
(Courtesy of RL Stamper, MD)

 

There are hormonal influences on aqueous secretion. Adrenalectomy in experimental animals leads to a decrease in aqueous formation, probably due to the resultant sharp decrease in the levels of circulating glucocorticoids or epinephrine.1,30 Furthermore, spironolactone, an aldosterone inhibitor, also decreases aqueous formation.342 The magnitude of this effect is small, however, and this, coupled with considerable toxicity, prevents spironolactone from being clinically useful in glaucoma therapy. However, in a recent study of patients who had undergone adrenalectomy,343 it was concluded that both the circadian rhythm of aqueous flow and the daytime response to timolol persist in the absence of the adrenal glands.

There was no correlation between endogenous progesterone levels and aqueous humor flow or IOP during the menstrual cycle of 20 healthy, nonpregnant women.344

Decreased plasma osmolality causes an increase in IOP and aqueous formation.345,346 Although aqueous formation is probably increased by an increase in water carried across the ciliary epithelium by the Na+ pump, the IOP increase is primarily the result of water gain to the eye by diffusion from the blood to the vitreous, and probably the aqueous as well.1 This test was utilized at one time for diagnosis of open-angle glaucoma (the water drinking test). However, the effect was not specifically diagnostic for glaucoma, and has been abandoned by most ophthalmologists.337 Similarly, increased plasma osmolality (as obtained by administering the hyperosmotic agents mannitol, isosorbide, or glycerine) causes a profound decrease in IOP, primarily because of water movement from vitreous to blood and only secondarily to a decrease in aqueous formation.345,347,348 Some authors report little effect of hemodialysis-reduced plasma osmolality on IOP.349

Numerous and varied pharmacologic agents are known to decrease aqueous formation and thus IOP. Several are known to increase formation.

Drugs that are known to reduce aqueous secretion include many β-adrenoceptor antagonists350 such as timolol,351–353 betaxolol,354 carteolol,355 levobunolol,356 metipranolol,357 oxprenolol, metoprolol, and befunolol,358 bupranolol,359 pindolol,360 labetalol,361 falintolol,362 and spirendolol.291

Carbonic anhydrase inhibitors reduce aqueous formation by approximately 50% at maximal dosage.44,68,69,72–76,91,363–366

Certain vasodilator substances also reduce aqueous secretion, including the cardiac peptide atrial natriuretic factor (ANF),367 which also reportedly stimulates an intracellular particulate guanylate cyclase and raises the levels of cyclic guanine monophosphate (cGMP) in the ciliary epithelium.292,368–370

In addition, certain of the nitrovasodilators, which are also known to mediate an increase in intracellular cGMP (via activation of a soluble guanylate cyclase), reduce aqueous secretion, including sodium nitroprusside,368,371 sodium azide,368 and nitroglycerin.294 However, Krupin and colleagues372 found that topical sodium nitroprusside or sodium azide increased aqueous secretion in human volunteers. This discrepancy may be dose-related.90 Nitrovasodilators may also lower intraocular pressure by enhancing aqueous humor outflow (see later section).373–375 cGMP itself also affects aqueous humor dynamics, for example, intravitreal injection of 8-bromo cGMP in the monkey promotes a reduction in aqueous flow of 15% to 20% and, at higher doses, an increase in outflow facility by 25% to 30%.376 A reduction in aqueous flow in rabbits in response to topical 8-bromo cGMP has also been described.377

Other vasoactive drugs, such as the calcium channel antagonists verapamil and nifedipine,378 reportedly reduce aqueous secretion in rabbits; however Beatty and co-workers379 reported that topical diltiazem and verapamil increased secretion and IOP in human volunteers.

The serotonergic antagonist ketanserin reduces the rate of secretion of aqueous in rabbits, cats, and monkeys.380,381 The serotonergic agonist serotonin sometimes reduces the rate of secretion. For example, topical application of serotonin in the rabbit eye results in a decrease in aqueous secretion, although intracameral injection results in an increased rate of secretion.381 Serotonergic receptors of a 5-HT1A–like subtype have been reported to exist in the iris/ciliary body of rabbits and humans.382,383 The serotonergic agonist flesinoxan has been found to decrease IOP and possibly aqueous secretion in rabbits.384 However, a significant increase in aqueous secretion, but no effect on IOP, was found in monkeys that had been treated with the 5-HT agonist 8-hydroxy-2(di-n-propylaminotetralin) (8-OH-DPAT), indicating the possible presence of a secretion-stimulating 5-HT1A receptor in monkey ciliary epithelium.385 It has been suggested that these receptors may be effectively antagonized by timolol and other β-blockers. Species differences may well account for the conflicting nature of these data, and furthermore, the precise nature of the putative 5–HT1A–like receptor subtype is still in question.382,383

Agonists at the 5-HT2 receptors have been identified as effective ocular hypotensive agents in the primate experimental glaucoma model.386,387 The ocular hypotensive response observed with 5-HT2 receptor antagonists in monkeys380 and humans388,389 may be mediated via pathways not related to their 5-HT2 antagonist activity. It was reported that neither 5-HT1A agonists nor 5-HT2 antagonists decreased IOP in the monkey with laser-induced ocular hypertension.386 R-DOI [R(-)-2-(4-iodo-2,5-dimethoxyphenyl)-2-aminopropane], a selective 5-HT2 agonist, causes a small but significant increase in aqueous humor formation and lowers IOP in normotensive monkeys primarily by increasing uveoscleral outflow.390 The magnitude of the ocular hypotensive response and increase in uveoscleral outflow as well as the slight increase in AHF produced by R-DOI are similar to the responses observed with PGF, i.e., in monkeys.143,145,391 The fact that functional 5-HT2 receptors have been found in human ciliary muscle and trabecular meshwork cells offers a plausible explanation for the IOP lowering action of R-DOI.392

Studies suggest that dopamine (DA2 and DA3) receptors play a role in aqueous humor secretion. The presence of such receptors in ocular structures remains to be shown definitively, and specific binding of [3H]-spiroperidol to membrane fractions of rabbit iris-ciliary body (ICB) has not been demonstrated.393 However, there is evidence based on adenylate cyclase activation studies to suggest that there are DA1-receptors in the ciliary body of the bovine and human eye but not in the rodent or rabbit eye.394

Pharmacologic evidence suggests the involvement of dopamine receptors in control of aqueous secretion. The topical administration of the dopamine DA2 agonists R(-)-2,10,11-trihydroxy-N-propyl-noraporphine hydrobromide (TNPA), bromocriptine, lergotrile, pergolide and the DA3 agonist 7-hydroxy-2-dipropylaminotetralin (7-OH-DPAT) to rabbits reduces secretion.395–397 Similarly, IOP decreases after topical application of the DA2/DA3 receptor agonist PD128,907 and aqueous humor secretion is reduced after intravitreal injection of PD128,907 in rabbits.398 However, this may not represent a local action of the drugs, because IOP is reduced in both eyes after unilateral administration of PD128,907, 7-OH-DPAT, bromocriptine, lergotrile, and pergolide. In contrast to lergotrile and pergolide, bromocriptine does not reduce secretion in normal monkeys.399 The ability of these drugs to reduce secretion in the rabbit is blocked by superior cervical ganglionectomy or by pretreatment with a DA2 or DA3 receptor antagonist such as domperidone (DA2), raclopride (DA2/DA3), UH232 (DA3) or U-99194 (DA3). These findings suggest that the most likely sites of action are the DA2 and DA3 receptors located on sympathetic nerve endings or ganglia.400 The activation of these receptors has an inhibitory effect on the release of norepinephrine.401 The acute unilateral instillation of 0.1% pergolide lowers the elevated IOP of glaucomatous monkeys,402 again via a reduction in aqueous secretion. Both orally and topically administered bromocriptine lowers the IOP of human volunteers.403–405 The DA2 antagonist haloperidol administered topically lowers the IOP of normotensive rabbits by suppression of aqueous formation.406,407 α1-Adrenoceptor antagonists such as prazosin,408 dibenamine, phentolamine, phenoxy-benzamine, thymoxamine,409 and yohimbine410 reportedly reduce secretion rate in rabbits but only by a very modest amount in humans. Unilateral instillation of 5% coryanthine (a stereoisomer of yohimbine) an antagonist with α1-adrenoceptor selectivity411 reduces IOP in normal rabbits and monkeys, but no change in either outflow facility or aqueous flow was observed in monkeys, and it was proposed that topical coryanthine may act by increasing uveoscleral outflow.412 Certain α2-adrenoceptor agonists are much more effective at reducing secretion rate and are used clinically, such as apraclonidine413–419 and brimonidine.297,298,419,420

The circulating renin-angiotensin system is an important determinant in the maintenance of adequate systemic blood pressure and may also be involved in organ-specific blood flow. All recognized renin-angiotensin system components have been identified in the eye of the human and other species. Angiotensin-converting enzyme (ACE) is present in human and rabbit aqueous humor421–423 and also in the feline ciliary body.424 Furthermore, angiotensinogen (a precursor of angiotensin I) is present in the nonpigmented ciliary epithelium of the human eye, more in the pars plana than in the pars plicata.425 Angiotensinogen was also found in the blood vessel lumina of the uvea and retina.425 Both the pigmented and nonpigmented layers of the human ciliary epithelium contain prorenin, the prohormone of renin,426 more so in the pars plicata than in the pars plana region.427 ACE inhibitors, such as captopril, enalapril, enalaprilic acid, and SCH 33861 reduce aqueous secretion when applied topically to rabbit eyes.428 Topical application of 0.1% SCH 33861 reduces aqueous secretion in glaucoma patients, although it is less effective than 0.5% timolol.429 H1-antihistamines such as antazoline and pyrilamine, when applied topically to normotensive albino rabbits in 4% solution,430 decrease aqueous secretion. The effect, however, may be unrelated to H1-receptor antagonism, being blocked by the α-adrenoceptor antagonist phentolamine; also other H1-receptor antagonists from the same chemical class (such as pyrilamine and tripelennamine) are ineffective.

A component of marijuana (cannabis), Δ9-tetrahydrocannabinol, reportedly reduces secretion of aqueous in human volunteers431 when injected intravenously or when inhaled via marijuana smoking. Topical Δ9-tetrahydrocannabinol also reduces secretion of aqueous in the rabbit.432 In contrast, however, topical Δ9-tetrahydrocannabinol has no effect on the human eye.433,434 Various cannabinoids such as HU-210 [(-)-7-hydroxy-Δ6-tetrahydrocannabinol dimethylheptyl],435 WIN 55-212-2 [R(+)-[2,3 dihydro-5-methyl-3-[(morpholinyl) methyl] pyrrolo-[1,2,3-de]-1,4-benzoxazinlyl](1-napthalenyl) methanone mesylate],436,437 and noladin ether [2-arachidonyl glycerol ether]438 reduce IOP in rabbits,438,439 rats,437 monkeys,436 and glaucomatous humans resistant to other forms of treatment.440 The mechanism by which cannabinoids lower IOP is not yet known. Studies have demonstrated the presence of functional CB1 receptors in the ciliary processes and trabecular meshwork of human and animal tissue.440–442 This suggests that cannabinoids may activate the CB1 receptors in the ciliary processes to decrease aqueous humor formation or in the trabecular meshwork to increase outflow facility or both, thus reducing IOP. Topical application of WIN-55-212-2 significantly reduces aqueous humor formation in normal and glaucomatous cynomolgus monkeys, but no increase in outflow facility was found. The decrease in aqueous humor formation after a single application was insufficient to explain the reduction in IOP, suggesting other mechanisms may be involved.436 Aqueous humor dynamics have not been determined after multiple doses of these cannabinoid compounds.

Opioid receptors can modulate various functions in the eye. The κ-opioid receptor agonists, bremazocine and dynorphin A, lower IOP bilaterally after unilateral topical administration by suppressing aqueous humor formation in rabbits.443,444 This is accompanied by an increase in atrial natriuretic peptide release via activation of ATP-sensitive K+ channels.444,445 The IOP and aqueous flow suppression were antagonized by the relatively selective κ-opioid receptor antagonist, nor-binaltorphimine (nor-BNI).443–445 A role for protein kinase C dependent activity is also suggested by the attenuation of bremazocine-induced C-type natriuretic peptide by chelerythrine.446 Inhibition of norepinephrine release and cyclic adenosine monophosphate (cAMP) accumulation in iris ciliary body in vitro by the κ-opioid agonists ICI 204 448 and spiradoline mesylate suggests that there are both prejunctional and postjuctional sites of action of κ- agonists.447 However, species differences may exist because the IOP-lowering response in monkeys after topical administration of bremazocine could be completely blocked by maintaining the mean arterial pressure by simultaneous intravenous infusion of angiotensin II.448

Metabolic inhibitors such as dinitrophenol (DNP) and fluoracetamide decrease aqueous formation, presumably via inhibition of the sodium pump, as well as other metabolically active pathways.30 The cardiac glycosides, such as ouabain and digoxin, specifically inhibit Na+/K+ ATPase in the ciliary body and thus decrease aqueous formation. From a clinical viewpoint, the dose of digoxin necessary to yield a significant reduction in aqueous secretion causes cardiac toxicity.365

There are few compounds known to increase the rate of secretion of aqueous. There are convincing reports of increased secretion in response to the β-adrenergic agonists epinephrine,278,449 salbutamol,450 isoproterenol,451 and terbutaline.452 The effect is small in the conscious state, but more pronounced during sleep. Such agents also stimulate aqueous flow in monkeys under pentobarbital anesthesia.277,306 As described earlier, some serotonogeric agents have been shown to increase aqueous humor flow in monkeys.385,390 Conversely, β-adrenoceptor antagonists, such as timolol, reduce the formation of aqueous humor during the day but not during sleep in humans.279 In monkeys, timolol does not reduce flow under pentobarbital (but does under ketamine) anesthesia305 but will block the increase in flow induced by terbutaline in the pentobarbital-anesthetized state.306 Certain other compounds, however, are effective at reducing aqueous formation during sleep, including the carbonic anhydrase inhibitor acetazolamide, and the α2-adrenoceptor agonists clonidine and apraclonidine.300,453 Therefore, the adrenergic agonist epinephrine (which acts on both α and β-adrenoceptor types) may have a dual effect on aqueous humor flow: stimulation via β-adrenoceptors, and inhibition via α2-adrenoceptors. These observations suggest that during the day aqueous humor formation is increased because of increased activity in the sympathetic nerves running into the ciliary body or to increased levels of circulating catecholamines, with resultant β-adrenoceptor activation. However, at night, aqueous flow falls to a basal, unstimulated level, consistent with diminished activity in the sympathetic nervous system or decreased humoral catecholamine secretion. The secretory epithelium of the ciliary body thus exhibits aroused and sleeping states.

Long-term use of drugs that decrease IOP by decreasing aqueous humor formation could have a negative effect on the eye. Some patients who had well-controlled IOP with timolol show evidence of reduced pressure control with continued administration.454 Studies on rabbits receiving acetozolamide show a restoration of IOP despite the fact that aqueous humor formation is suppressed. This is consequent to a reduction of outflow facility455 Similarly, humans receiving oral acetazolamide over a several week period demonstrate restoration of IOP but reduction in tonographic outflow facility.455 Underperfusion of the trabecular meshwork can lead to detrimental morphologic effects456 (see the Trabecular Outflow: Other Agents section for additional details).

Back to Top
PATHWAYS OF AQUEOUS FLOW
Aqueous humor is formed in the posterior chamber, and while a small amount undoubtedly finds its way into the vitreous, most of the newly formed aqueous passes from the posterior chamber to the anterior chamber, via the pupil.1,30 While in the posterior chamber, the aqueous undergoes some chemical changes. The lens, and possibly anterior vitreous, extract oxygen, amino acids, glucose, ascorbate, and other small molecules, and discharge metabolic wastes such as CO2 and lactate.1,224

Ordinarily, there is little or no hindrance of the flow of aqueous through the pupil and thus the pressures in the posterior and anterior chambers are approximately equal, although the posterior chamber pressure must be slightly higher for flow to occur. Under certain circumstances, such as adhesions between the posterior surface of the iris and the anterior lens capsule (posterior synechiae) or the anterior surface of the iris and the corneal endothelium (anterior synechiae), or unusually tight juxtaposition of the pupillary iris and the anterior lens capsule, resistance to the flow of aqueous through the pupil (pupillary block) occurs (Fig. 11).457 If the resistance to flow is great enough, then the pressure in the posterior chamber increases relative to the anterior chamber. The pressure difference may be small, but even so may be sufficient to cause the peripheral iris to balloon forward. If the anterior chamber is at all shallow, as occurs by heredity in a small percentage of people, the peripheral iris may make contact with the trabecular meshwork, shutting off outflow of aqueous from the eye. The result is a rapid increase in IOP,458 as occurs clinically in acute congestive angle-closure glaucoma. Such a situation represents a medicosurgical emergency and must be treated rapidly with initial drug therapy (usually a combination of a miotic (pilocarpine), a β-blocker (timolol), a carbonic anhydrase inhibitor (acetazolamide), and a hyperosmotic agent (glycerine)), followed shortly afterward by iridectomy or laser iridotomy (the surgical formation of a hole [fistula] in the peripheral iris in order to allow unimpeded access of posterior chamber aqueous to the anterior chamber337) in order to save the retina and optic nerve from the damaging effects of the high IOP. Other causes of pupillary block include swelling of the lens or ciliary body, posterior direction of aqueous (ciliary block), choroidal or retinal tumor, retrolental fibroplasia, or persistence of hyperplastic primary vitreous.

Fig. 11. Mechanism of angle-closure glaucoma. A: Relative pupil block. B: Iris bombé. C: Iridotrabecular contact. (From Kanski JJ: Glaucoma. In Kanski JJ, ed.: Clinical Ophthalmology, 2nd ed. London: Butterworth-Heinemann, 1989:182–231, with permission.)

In the normal situation, once through the pupil and into the anterior chamber, the aqueous humor flows peripherally toward the anterior chamber angle where the drainage system lies. Superimposed on this peripherally directed and hydrostatically generated flow is the thermal circulation of the anterior chamber (Fig. 12), evidenced clinically in uveitis by the movement of cells in the anterior chamber upward near the warmer iris and lens and downward near the cooler cornea. Frequently, cells, pigment, and other debris may be deposited on the corneal endothelium in a spindle-shaped pattern reflecting the thermal current.1,457 By gonioscopy, the inferior trabecular meshwork may frequently be seen to contain more pigment and detritus than other areas of the meshwork, probably as a result of the effects of gravity.

Fig. 12. Thermal circulation in the anterior chamber. Arrows indicate direction of aqueous flow.

Back to Top
AQUEOUS OUTFLOW
Aqueous humor leaves the eye through the angle of the anterior chamber, mostly by way of the trabecular meshwork, Schlemm's canal, the intrascleral collector channels, the aqueous veins, and the episcleral venous plexus, whence aqueous joins the general venous drainage.1 In addition, an alternate uveoscleral pathway exists (Fig. 1) that accounts for nearly half of total aqueous drainage in young human eyes459 as is detailed later.

A minimal amount of fluid leaves the anterior and posterior chambers through the iris and through the vitreous to the optic nerve and retinal vessels.1,30

Back to Top
METHODS OF MEASUREMENT OF AQUEOUS OUTFLOW
There are several techniques available for the measurement of aqueous outflow in the living eye. Of these, Bárány's two-level constant pressure perfusion technique267 is main the method used in experimental animals, when measurement of total outflow facility is required. This method involves cannulation of the anterior chamber with a steel needle and connection to a reservoir containing mock aqueous humor solution. The reservoir is maintained at a particular height above the eye, thereby creating an artificially elevated IOP, which is measured via a second steel needle in the anterior chamber, connected to a microsensitive pressure transducer. The reservoir is mounted on a sensitive strain gauge that indicates via a pen recorder the weight of the reservoir. As the experiment proceeds, solution flows from the reservoir to the anterior chamber, and from there flows through the outflow pathways of the eye. The rate at which the reservoir empties is calculated from its change in weight with time. The procedure is run thus for a period of 4 minutes, whereupon the reservoir is raised to a new, higher level, creating a higher IOP. The procedure is then run for 4 minutes more, and then the reservoir lowered to its original level; the procedure is run for a further 4 minutes, and so on. Then, the mean rate at which the reservoir empties at each pressure is calculated. Total outflow facility is calculated as the difference in the mean rates of emptying of the reservoir at the two different pressures, divided by the pressure difference, and is expressed in microliters per minute per millimeter of mercury (μL/min/mm Hg). All of these calculations may be performed instantaneously on a computer, linked to the experimental apparatus.

It is also possible to resolve trabecular and uveoscleral outflow, and their respective facilities, by perfusion of the anterior chamber with a solution of radiolabeled albumin. In this procedure, if both eyes are to be studied, one eye is perfused with 125I-albumin, and the other with 131I-albumin. The radiolabeled albumin is continuously circulated through the anterior chamber by a pump. As secreted aqueous humor drains through the trabecular pathway, it takes some of the radiolabeled albumin with it and passes it directly to the general circulation almost immediately. The proportion of aqueous that flows via the uveoscleral route, however, takes approximately 2 hours to reach the general circulation, and so the radiolabel flowing with this proportion of the aqueous does not appear in the blood for this time, after commencement of the perfusion.145,268,460 The experiment is run for 1 hour and 40 minutes. During this time, blood samples are withdrawn from the experimental animal every 5 minutes, and counted on a γ-counter. This yields an estimate of the rate of flow of the radiolabel to the general circulation via the trabecular meshwork, equated with trabecular outflow. One can modify the technique so that an estimate of trabecular outflow facility may be calculated by determining the difference between such flow measurements made at two different intraocular pressure.268,269 Uveoscleral outflow can be determined directly by subsequent euthanization of the animal, dissection of the eye, and counting of the tissues involved in the uveoscleral outflow pathway. Again, if perfusion at two different pressures is undertaken, uveoscleral outflow facility may also be calculated. High molecular weight fluorescinated dextrans have also been used and have yielded values comparable to those with labeled proteins.143,461 As an alternative to killing the experimental animal, aqueous flow rate (assumed to be equal to the rate of secretion) can be estimated by monitoring the rate of loss of radioactivity circulating through the anterior chamber using a multichannel analyzer.460 From this, uveoscleral outflow can be calculated indirectly as the algebraic difference between aqueous flow rate and trabecular outflow.

Although the above techniques are the most accurate known for the determination of rate of aqueous humor outflow, their invasive nature makes them unsuitable for clinical use. Furthermore, the techniques for resolution of trabecular and uveoscleral outflow using radiolabeled albumin are very complex, time consuming, costly, and difficult to perform in practice.

However, in the human, the less accurate but noninvasive technique of tonography243 can be used to determine total outflow facility. The technique is based on the theoretic work of Friedenwald.462,463 Tonography estimates the decrease in IOP continuously over a period of 4 minutes while a known weight is applied to the eye. From the rate of change of pressure (the slope of the pressure recording), the outflow facility may be determined. The assumptions underlying this method are that episcleral venous pressure maintains a constant value, and that little or no change in aqueous formation is induced by the instrument itself. The values of flow, IOP, and an assumed value for Pe can be substituted into the equation in order to obtain a value for outflow facility. There are many sources of error in tonography, but, as is discussed later, many of these errors appear to cancel each other out because the figures agree quite well with those obtained by the perfusion method, even in the same eye. Prijot and Weekers464 computed an average flow of 1.63 μL/min in 46 human subjects, using this method. Others have also found values between 1.5 to 2 μL/min for normal human subjects,1,30 although these values are significantly lower than the value of 2.5 μL/min found by fluorophotometry266,465 However, this difference may be accounted for by uveoscleral outflow which, because of its relative pressure-independence, is not measurable by tonography. The obvious advantages of tonography are that it can be used readily in the human eye, and can be repeated several times in the same eye, for comparative purposes. Variations of tonography, such as the perilimbal suction cup technique466 or tonography performed at constant pressure and whereby IOP is equal to Pe (Pv tonography),467 were devised, but the original form devised by Grant has become the standard for the measurement of outflow facility and is the most convenient (albeit not the most accurate) method of estimating aqueous humor flow in humans.3

An indirect technique for measuring outflow facility and uveoscleral outflow in humans as well as monkeys is based on calculations done with fluorophotometry measurements before and after aqueous flow suppression with acetazolamide and timolol. Fluorophotometric outflow facility is calculated from the change in flow and IOP measurements before and after acetazolamide and timolol intervention. Uvoescleral outflow is then calculated.152,468–472 This technique is currently the method of choice for obtaining uveoscleral outflow estimates in humans and has been used extensively to elucidate changes in aqueous humor dynamics with age and in glaucoma as well as the mechanisms of action and interactions of the newer glaucoma experimental and therapeutic agents.152,459,469,470,473–476 Normal values for young and old ocular normotensive humans obtained with this method range from 0.21 to 0.25 μL/mm Hg per minute for outflow facility; 2.4 to 2.9 μL/min for aqueous humor flow; and 1.16 to 1.64 μL/min for uveoscleral outflow.459

Back to Top
TRABECULAR PATHWAY
The angle of the anterior chamber is bounded anteriorly by the corneal endothelium, and posteriorly by the root of the iris and ciliary body. At the apex of the angle lies the trabecular meshwork, suspended between Descemet's membrane and the anterior portion of the ciliary muscle (Figs. 1 and 13). The trabecular meshwork commences just posterior to the point where Descemet's membrane terminates. This transition zone is identified gonioscopically as Schwalbe's line but is seen less easily histologically. The trabecular meshwork continues posteriorly until it joins the scleral spur and ciliary muscle. The inner portion of the meshwork (that closest to the anterior chamber) is called the uveal meshwork, and the outer portion closest to Schlemm's canal constitutes the corneoscleral meshwork, which is itself separated from the endothelium lining Schlemm's canal by the juxtacanalicular tissue, or endothelial meshwork.1 There is, however, no sharp dividing line between the portions. A portion of the meridional ciliary muscle fibers insert into the trabecular meshwork.4,477–479

Fig. 13. Cutaway diagram of layers of the trabecular meshwork in the aqueous outflow system. (From Shields MB: Aqueous humor dynamics. In: Kist K, ed.: Textbook of Glaucoma. Baltimore: Williams & Wilkins, 1987:5–44, with permission.)

The uveal meshwork consists of small, round criss-crossing, interlocking, and branching bands (trabeculae) approximately 4 mm in diameter (Figs. 1 and 13), which tend to be radially oriented. The bands or chords consist of a single layer of endothelial cells surrounding a core of collagen. The spaces between the chords are irregular and range from 25 to 75 μm across. The chords are arranged in layers, or lamellae, but the layers are interconnected. Occasionally, some of the lamellae bridge across the ciliary body face in the angle recess to attach to the root of the iris and form an iris process.4,477,480

The trabeculae of the corneoscleral meshwork are more like broad, flat, endothelium-lined sheets, which tend to be progressively oriented in a circumferential pattern (Figs. 1 and 13). They are 3 to 20 μm in length, and are arranged in bands running around the angle. The spaces between the trabeculae are much smaller than in the uveal meshwork (10 to 30 μm) and are more elliptical. The spaces of adjacent lamellae are not behind one another but are offset, providing a rather convoluted passageway from internal to more external lamellae. Furthermore, as the lamellae approach Schlemm's canal, the spaces become 1 to 2 μm.481,482

The juxtacanalicular or endothelial meshwork, which separates the corneoscleral meshwork from Schlemm's canal is composed of a ground substance of connective tissue, incorporating glycoproteins and glycosaminoglycans.

The trabecular meshwork forms the major site of resistance to aqueous outflow, that is, 60% to 80%.483 Removal of the trabecular meshwork from enucleated perfused human eyes results in a 75% or greater decrease in outflow resistance.270,484–490 The question of the exact site of resistance to aqueous humor outflow has significance for delineating the pathophysiology of POAG. Unfortunately, to date, the exact mechanism of aqueous humor outflow and site of resistance remain elusive. It is assumed that the aqueous easily percolates through the large spaces of the uveal meshwork and through the larger openings of the inner corneoscleral meshwork. The resistance to aqueous outflow is believed by most investigators to be located largely in the juxtacanalicular connective tissue, which has a covering of a single layer of endothelial cells on the side facing Schlemm's canal (the juxtacanalicular cells).491,492 Some investigators feel that the main resistance may lie slightly proximal to the juxtacanalicular tissue.493–495 Separation of the basal lamina cells lining of the inner wall of Schlemm's canal from the underlying juxtacanalicular tissue as seen using quick-freeze/deep-etch electron microscopy may represent flow pathways.496 Under normal conditions, the inner wall of Schlemm's canal and juxtacanalicular cells may be in a contracted state, limiting the routes available for fluid flow as demonstrated with gold particle infusion studies in nonhuman primates.497,498 Expansion of the area available for fluid drainage can increase the rate of fluid outflow.497,498 The accompanying loss of extracellular material may not be responsible for the decrease in resistance to fluid outflow.496–499

Aqueous humor enters Schlemm's canal inner wall endothelial cells via 1.5-μm passages.500,501 Vesicles and large vacuoles are frequently seen in the endothelium of the inner wall of Schlemm's canal (Fig. 14).491,501–504 The vacuoles may be as large as 5 by 14 μm. Their function is as yet unknown, but there is much speculation on the subject. Some have felt them to be an artifact, whereas others have suggested that these vacuoles serve a role in the exit, by bulk flow, of fluid and large molecules from the trabecular meshwork into the canal.505,506 Such vacuoles are also found in the arachnoid villi, and are involved in the drainage of cerebrospinal fluid.507 Tripathi508,509 postulated that these giant vacuoles alone serve as a transcellular pathway for aqueous flow across the inner wall of Schlemm's canal. Based on electron microscopic studies using ferritin as a tracer, Cole and Tripathi510 calculated that endothelial vacuoles constantly discharging their contents into Schlemm's canal have enough volume to account for the entire aqueous outflow. While this view is not universally held,511 these vacuoles undoubtedly provide a channel across which some aqueous and large particulate matter flow. This pathway has also been elegantly demonstrated by others,491 showing that red and white blood cells, as well as latex particles up to 1 μm in diameter, entered Schlemm's canal by way of the vacuoles (Fig. 15). Grierson and Lee512 showed that as pressure increased within the physiologic range, vacuolization of the endothelium of Schlemm's canal increased, but that pressures above the physiologic range produced actual occlusion of Schlemm's canal by distension of the corneoscleral meshwork and prolapse of the endothelium. Most giant vacuoles have short survival time once perfusion pressure decreases to zero, suggesting they can respond rapidly to IOP changes.513 Giant vacuoles may not be real intracellular vacuoles but rather dilatations of paracellular spaces.511

Fig. 14. Electron micrograph of the inner wall of Schlemm's canal in a cynomolgus monkey after anterior chamber perfusion with cationized ferritin (black) (21,000×). (From Epstein DL, Rohen JW: Morphology of the trabecular meshwork and inner-wall endothelium after cationized ferritin perfusion in the monkey eye. Invest Ophthalmol Vis Sci 32:160, 1991, with permission.)

Fig. 15. A red blood cell (RBC) is seen passing through a pore in the flat portion of the endothelium (En) to the lumen of Schlemm's canal. The endothelium extends a funnel-like process toward the meshwork side. ER, endoplasmic reticulum; Th, thorotrast (12,000×). (From Inomata H, Bill A, Smelser GK: Aqueous humor pathways through the trabecular meshwork and into Schlemm's canal in the cynomolgus monkey (Macaca irus). An electron microscopic study. Am J Ophthalmol 73:760, 1972, with permission.)

Pores of 0.8 to 1.8 μm were also present in nonvacuolated endothelial cells.491 Why endothelial cells should have tight junctions between them but pores within them is difficult to reconcile. Bill and Svedbergh514 have shown that the endothelial cells have a total of 20,000 pores with diameters of up to 3 μm, and that calculations based on this reveal that the pores could be responsible for all the bulk aqueous flow. Pores of the inner wall endothelium are thought to generate at most 10% of the resistance to aqueous humor outflow in humans. Pores of the inner wall cause a funneling effect in which aqueous humor flows preferentially through those regions of the juxtacanalicular tissue.515 However, the possibility that inner wall pores are fixation-induced artifacts cannot be excluded.516 Decreasing the temperature does not appreciably affect aqueous outflow, suggesting that an active metabolic process is not directly involved. Although the cells of the inner canal wall and the juxtacanalicular tissue are biologically active and control flow in various ways, the flow itself is passive.517

Water channels (APQs), described earlier in the ciliary epithelium, are also found in trabecular meshwork and Schlemm's canal cells.58,518–520 They may be involved in transcellular water and movement in these tissues. Adenovirus overexpression of APQ 1 in trabecular meshwork cells in vitro results in an increase in cell volume and a reduction in paracellular permeability, suggesting that APQ 1 may affect outflow facility in vivo.521

Another mechanism for regulation of trabecular paracellular pathways may be via Na-K-2Cl cotransport.522 Inhibition of this transport by bumetanide and other agents alters trabecular cell volume and the permeability of trabecular cells monolayers.523 Exposure to hyperosmotic or Cl-free medium or medium containing bumetanide transiently increases outflow facility in human and calf eyes in vitro.524 However, contradictory results were found in another study where outflow facility in monkeys eyes in vivo and in human anterior segments in vitro was unaffected by bumetanide.525 Trabecular meshwork cell volume is greater in glaucoma compared to normal eyes despite reduced Na-K-2Cl cotransport activity, suggesting that other volume-regulatory ion flux pathways may be involved in the reduced outflow of glaucoma.526 These include K+ and Cl- channels, Na+/H+ antiports, and possibly K+ - Cl symports.527 (Fig. 16).

Fig. 16. Schematic representation of a trabecular meshwork cell showing the large number of transporters, channels, and receptors that have been identified in these cells. cAMP, cyclic adenosine monophosphate; cGMP, cyclic guanosine monophosphate; EP2, prostaglandin E2; TP, thromboxane A2. (From Wiederholt M, Stumpff F: The trabecular meshwork and aqueous humor reabsorption. In Civan MM, ed.: Current Topics in Membranes. San Diego: Academic Press, pp.1998:163–202, with permission.)

The morphology of the monkey trabecular meshwork after perfusion of the anterior chamber with cationized ferritin (CF) has also been studied.511 CF, which binds negatively charged sites, was found to adhere to the free surfaces of trabecular cell membranes and to accumulate in the cribriform layer underlying the endothelial lining of Schlemm's canal. Tangential sections of the inner-wall endothelium of Schlemm's canal demonstrated that separations of the adjacent cell membranes occur between the tight junctions, forming lacunae and bent, tunnellike channels that represent continuous paracellular pathways, through which the perfusing fluid had completely passed. These paracellular pathways appeared enlarged and were more easily identified at elevated IOP. In general, intracytoplasmic vacuoles demonstrated heavy staining with CF on their luminal surface but only faint staining on their adluminal (juxtacanalicular tissue-facing) surface. The study indicated that there are paracellular routes through the inner-wall endothelium by which high molecular-weight substances such as ferritin and macrophages can exit the anterior chamber. In monkeys eyes in vivo perfused with cationized and noncationized gold, gold particles reached the basal aspects of intercellular junctions but never crossed them toward the lumen of Schlemm's canal. Instead they were transported via vesicles from the basal aspect of the inner wall into the canal lumen.497 Both transcytoplasmic and paracellular mechanisms of aqueous outflow may exist depending on the different conditions of pressure or flow.

An alternative to the bulk flow model of aqueous outflow is one that characterizes it as an active phenomenon driven by means of a mechanical pump. Johnstone528 reviews decades of research supporting the hypothesis that the aqueous outflow pump receives power from transient increases in IOP such as occur in systole of the cardiac cycle, during blinking and during eye movement. These transient pressure spikes cause microscopic deformation in the elastic structural elements of the trabecular meshwork, juxtacanalicular cells, and Schlemm's canal inner wall endothelium. During systole, Schlemm's canal endothelium moves outward into Schlemm's canal forcing aqueous into collector channel ostia and aqueous veins. At the same time, the IOP increase of systole forces aqueous into one-way collecting vessels or valves that span Schlemm's canal. When the pressure decreases, the elastic elements return to their original configuration causing a relative pressure reduction in Schlemm's canal that induces aqueous to flow from the valves into the canal. Mechanotransduction mechanisms, which are well characterized for the vascular system to regulate pressure and flow, also provide a means of regulatory feedback to control IOP and aqueous flow. Although the majority of current evidence favors the bulk flow model of outflow, the mechanical pump theory provides an intriguing area for future research.

Nerve endings and neuropeptides have been demonstrated in the trabecular meshwork and scleral spur. Tamm and colleagues529 described the presence of putative afferent mechanoreceptors in the scleral spur region that measure stress or strain in the connective tissue elements of the scleral spur, which might be induced by ciliary muscle contraction or changes in IOP. Ruskell530 identified trigeminal nerve fibers in the scleral spur and trabecular meshwork and also postulated a role in recording tension produced by ciliary muscle contraction. The close association of varicose axons with the myofibroblastlike scleral spur cells indicates that nervous signals modulate scleral spur cell tone.531 Sympathetic scleral spur cell innervation is present only in cynomolgus monkeys but seems to be absent in humans. Conversely, scleral spur axons of presumably parasympathetic origin are absent in the cynomolgus monkeys but present in humans. Cholinergic innervation of the scleral spur cells seems to be rare or absent.531 Holland et al532 suggested that not only sensory functions but also sympathetic and parasympathetic functions are represented. Calcitonin gene-related peptide (CGRP)533 and substance P534-containing peripheral fibers have been found in the trabecular meshwork and inner and outer wall of Schlemm's canal of the human and rhesus monkey eye. Stone and colleagues535 also located vasoactive intestinal polypeptide-containing neurons in the limbal blood vessels and trabecular meshwork of the human eye. Neuropeptide Y-containing nerves have been reported in the aqueous outflow apparatus and limbal and uveal blood vessels of the rat, guinea pig, cat, and monkey536 but not in the human eye to a significant degree.536 Cholinergic and nitrergic nerve terminals that could induce contraction and relaxation of the trabecular meshwork and scleral spur cells were demonstrated in monkey and human eyes.537 Terminals in contact with the elastic-like network of the trabecular meshwork and containing substance P immunoreactivity resemble afferent mechanoreceptor-like terminals in other parts of the body. These findings raise the possibility that the trabecular meshwork may have some ability to self-regulate aqueous humor outflow. Sympathetic denervation of monkey eyes suggests there is little outflow resistance-relevant resting sympathetic tone on the ciliary muscle or trabecular meshwork.538 Ocular sympathetic innervation has minimal function in regulating IOP in humans539 and subhuman primates.540

Nitric oxide mimicking nitrovasodilators can decrease IOP in monkeys by altering outflow resistance. In human eyes, the trabecular meshwork and ciliary muscle are enriched sites of nitric oxide systhesis.374 One role for nitric oxide in the anterior segment may be to modulate outflow resistance either directly at the level of the trabecular meshwork, Schlemm's canal and collecting channels, or indirectly through alteration in the tone of the longitudinal ciliary muscle (see Uveoscleral Outflow). Nitrovasodilators were shown to relax trabecular meshwork373 and ciliary muscle373,375 strips precontracted with carbachol in vitro.

Once aqueous humor has percolated through the trabecular meshwork, it finds its way into Schlemm's canal (see Fig. 14), a ringlike channel of irregular diameter, which has the structure of a venous channel with a thin connective tissue wall surrounding an endothelium-lined lumen.1,4 The canal often divides in places into two channels that reunite. Many diverticula, as well as blind tubules, may also sprout from the canal.

Interposed between the endothelium of the innermost trabecular layer and the endothelium of Schlemm's canal is amorphous basement-membrane-like material made up of major structural proteins including collagen type IV, laminin, and fibronectin.541–544 A poorly defined, discontinuous endothelial basement membrane has been described.496 The endothelial cells are joined together by tight junctions.4 Tight junctions are more complex in human eyes than in monkey eyes492,545 and become less complex with increasing pressure.546

Phagocytosis by trabecular macrophages of particulate matter and red blood cells does occur (Fig. 17).547 Although this process may be important in clearing the anterior chamber of some inflammatory and other debris, it is probably not significant for bulk outflow. However, phagocytosis results in a loss of adhesiveness of trabecular cells to their substrate.548,549 Loss of trabecular cells as a result of excess or prolonged phagocytosis could alter trabecular outflow. Decreased trabecular meshwork cellularity with age190–192 could alter the synthetic and catabolic control of the extracellular environment.191 Alternatively, the accumulation of debris could be toxic to trabecular meshwork cells or sequester them from the aqueous humor necessary to supply them with nutrients and remove toxic metabolites.550

Fig. 17. Removal of pigment from the trabecular meshwork by macrophages in pigmentary glaucoma. A: Light micrograph of Schlemm's canal. A pigment-filled cell (black arrow) is partially inside the canal lumen and another is close to the lumen (white arrow) (600×). B: An enormously enlarged, pigment-filled cell is found entirely within the canal lumen (600×). C: Electron micrograph of the same cells as in (B). This presumptive macrophage is completely filled with melanin granules. (From Alvarado JA, Murphy CG: Outflow obstruction in pigmentary and primary open-angle glaucoma. Arch Ophthalmol 110:1769, 1992, with permission.)

Once in the lumen of Schlemm's canal, aqueous humor then flows into the 25 to 35 endothelial tubules that sprout irregularly from the circumference of the outer wall of Schlemm's canal (Fig. 18). Some of these external collector channels connect directly to the deep scleral venous plexus. However, most penetrate through the sclera and join the episcleral venous plexus, where aqueous humor is drained into the venous system.4,551

Fig. 18. Schematic drawing of Schlemm's canal and collector channels. (From Hogan MJ, Alvarado JA, Weddell JE: Histology of the Human Eye. An Atlas and Textbook. Philadelphia: WB Saunders, 1971:136–153, 260–319, with permission.)

Ascher552 initially suggested that a small change in shape of the intrascleral collector channels could cause a large increase in the resistance to aqueous outflow. This view was supported by Sears, who felt that a significant percentage of the outflow could be attributed to this site.553 Krasnov554 felt that the intrascleral channels were the site of blockage in some cases of glaucoma, and he designed an operation to bypass just this site. His unconfirmed reports indicate substantial early success with this technique.

Ascher18 identified veins in the living human limbus by slit-lamp microscopy that could be seen to contain aqueous humor. These aqueous veins would then join an episcleral vein where the aqueous could be seen joining the bloodstream, but instead of mixing immediately, the aqueous stream could be seen as a clear lamina flanked by blood (Fig. 19). Further downstream, the lamina of aqueous does mix with blood. That the clear fluid in these aqueous veins is aqueous humor was confirmed by fluorescein injection and later by neoprene and silicon injections.1,20,552

Fig. 19. Aqueous veins. Diagram of scleral collector channel (S) joining episcleral vein (V). Aqueous laminar flow occurs for a short distance prior to mixing of aqueous and blood. (From Kolker AE, Hetherington J Jr: Gonioscopic and microscopic anatomy of the angle of the anterior chamber of the eye. In Becker-Schaffer's Diagnosis and Therapy of the Glaucomas, 4th ed. St. Louis: Mosby-Year Book, 1976:21–41, with permission.)

The outer trabecular meshwork or inner wall of Schlemm's canal is subject to many influences: physiologic, pharmacologic, and pathologic. Removing, disrupting, bypassing, or deforming the outer trabecular meshwork (such as in trabeculectomy or laser trabeculoplasty) leads to marked facilitation of aqueous outflow. Partial internal trabeculotomy causes an increase in aqueous outflow facility almost proportional to the amount of trabecular meshwork bypassed. However, the response of aqueous outflow to partial trabeculotomy suggests that there is little circumferential flow in Schlemm's canal. Despite the fact that, after trabeculotomy, aqueous humor has unrestricted access to up to 30% of the circumference of Schlemm's canal and, although there is an increase in aqueous outflow facility in response to partial trabeculotomy, there is little increase in aqueous outflow in enucleated monkey and human eyes beyond that of normal healthy eyes.485,555 If there were flow of aqueous circumferentially through the canal of Schlemm in the normal eye, then unrestricted access to even a small part of the canal by trabeculotomy in enucleated monkey and human eyes would produce a marked increase in total outflow and outflow facility, to greater than normal physiologic values.

The ciliary muscle plays a key role in the regulation of trabecular outflow facility.479 Many of the trabecular sheets appear histologically to be the point of insertion of the meridional fibers of the ciliary muscle. The tendons of the muscle spread anteriorly and interconnect with the elasticlike tissues of the trabecular meshwork. Additionally, the tendons of the ciliary muscle bundles join with the elastic fibers of the scleral spur. The anterior anatomic relationship between the ciliary muscle and the trabecular meshwork is extremely intimate and of far greater functional significance than merely representing an anterior anchor or insertion point for the muscle. Some of the muscle tendons pass uninterrupted completely through the meshwork to insert into the peripheral cornea at Schwalbe's line, which thus presumably does serve an anchoring function. However, other tendons splay out within the mesh and intermingle with its elastic network. Via connecting fibrils, this network is attached to the juxtacanalicular region and the inner wall endothelium of Schlemm's canal. The latter contain specialized cell surface and cytoplasmic cytoskeletal complexes subserving this attachment at discrete points.478,479,556–559 When the ciliary muscle contracts, the trabecular lamellae are thought to be separated, the scleral spur is displaced posteriorly and internally,560 and Schlemm's canal is dilated, yielding a decrease in outflow resistance.339,561–563 Thus, accommodation stimulated parasympathetically, pharmacologically, and centrally, leads to a decrease in outflow resistance.1,564,565 Kaufman and Bárány562 produced further evidence in support of the hypothesis that the ciliary muscle affects outflow resistance when they were able to disinsert the ciliary muscle from the scleral spur in cynomolgus monkeys without obvious damage to the trabecular meshwork and Schlemm's canal. When their monkeys had recovered from the operative procedure, pilocarpine had no effect on outflow resistance. This suggests that the mechanism of action of cholinergic activity is through the stimulation of the ciliary musculature that in turn, causes a mechanical opening of the trabecular meshwork. Because the longitudinal fibers of the ciliary muscle are putatively more facility-specific and the circular fibers are putatively more accommodation-specific, and because the enzyme histochemistry and cellular ultrastructure of these regions differ such that there appear to be two types of contractile fibers present, namely, twitch (longitudinal) and tonic (circular) fibers,566 perhaps a specific pharmacologic agent, selective for the longitudinal but not the circular fibers of the ciliary muscle, may be able to improve outflow in glaucoma but without a concomitant undesirable accommodative effect. Although the accommodative and outflow facility-increasing effects of cholinergic agonists such as pilocarpine567 and aceclidine568 may be dissociated under some conditions, the facility, accommodative, and miotic responses to both drugs in the monkey eye all seem to be mediated by the M3 muscarinic receptor subtype.569,570 Analagously, contractile responses in the longitudinal and circular vectors of the monkey ciliary muscle to pilocarpine, carbachol, aceclidine, and oxotremorine are not dissociable and seem to be mediated by M3 receptors.571,572

The ciliary muscle is sympathetically innervated,573 and there appears to be a weak β-adrenergic relaxant response of at least some portions of the muscle.574–576 In order to define any possible role of the ciliary muscle in the modulation of outflow facility responses, ciliary muscle disinserted and nondisinserted eyes were given equal intracameral doses of epinephrine or norepinephrine.577 On a percentage basis, the outflow facility increase and dose-response relationships were identical in both types of eyes for each drug, respectively. Thus the ciliary muscle was not involved in the responses, and the disinsertion procedure itself had not altered the responses.578 Sympathetic innervation is not necessary for the outflow facility response to epinephrine in monkeys.538

Rather, the facility-increasing effect is mediated by β2-adrenergic receptors on the trabecular endothelial cells, and the subsequent G-protein–adenylate cyclase–cAMP cascade.579 β-Adrenergic receptors present on cultured human trabecular cells and trabecular meshwork cells obtained from human eyes post mortem have been characterized.580 Competition studies were carried out with a series of agonists and antagonists, and it was concluded that human trabecular cells possess a single class of β-adrenergic receptor of the β2 subtype. Consistent with this finding are the reports of the increase in trabecular outflow facility in response to β-adrenergic agonists.277,278,449,581 Trabecular meshwork cells synthesize cAMP in response to stimulation with β-adrenoceptor–selective agonists, such as epinephrine.553,582 The increase in cAMP synthesis by trabecular meshwork cells in response to epinephrine can be blocked by timolol.582 Furthermore, the facility-increasing effect of epinephrine is blocked by timolol,583,584 but not betaxolol,585,586 consistent with the hypothesis that there are no β1-adrenoceptors present in the trabecular meshwork. Intracameral injection of cAMP results in increased conventional outflow facility in rabbit587 and monkey eyes.588,589 Epinephrine increases facility and perfusate cAMP levels in the organ-cultured perfused human anterior segment, effects that are blocked by timolol and the selective β2 antagonist ICI 118,551.590 In monkey ciliary muscle, ciliary process, trabecular meshwork and iris tissue in vitro, epinephrine stimulates cAMP production. Part of the facility response in vivo and cAMP production in vitro is inhibited by indomethacin, suggesting that part of epinephrine's mechanism of action may be via prostanoid production or release.591 However, the epinephrine-induced facility increase in human anterior segments in vitro was attributed to protein synthesis and not to prostaglandin production.592

The nature of the physiologic change in the meshwork responsible for the decreased flow resistance remains uncertain. One hypothesis involves epinephrine-induced disruption of actin filaments within the trabecular meshwork cells, consequent alteration in cell shape, and increased hydraulic conductivity across the meshwork.593 Relaxation of the trabecular meshwork also could play a role in the outflow facility response to epinephrine.594 Thus, cytochalasin B (a disruptor of actin filament formation) potentiates the facility-increasing effect of epinephrine,595 while phalloidin (a stabilizer of actin filaments), inhibits it.596 Continuous exposure to epinephrine at a concentration of 10 μmol produces arrest of normal cytokinetic cell movements, inhibition of mitotic and phagocytic activity, marked cell retraction, and separation from the substrate and cellular degeneration after 4 to 5 days in cultured human trabecular cells.597 Similarly, the hydraulic conductivity of trabecular cell monolayer cultures grown on filters is increased by epinephrine and is associated with changes in cell shape and with separation between cells.581 These actions of epinephrine are partially blocked by pretreatment with timolol.

The juxtacanalicular, corneoscleral, and uveoscleral regions of the meshwork have been considered as possible sites of resistance to outflow in glaucoma.493,494,598 Cell number diminishes at these sites in POAG and pigmentary glaucoma,190,192,193,599–601 but it is difficult to correlate cell loss with increased resistance to outflow. Alvarado and Murphy547 evaluated the trabecular meshwork tissues internal to the juxtacanalicular apparatus in glaucomatous and normal human eyes to determine how they may participate in the development of pigmentary and POAG. They reported that trabecular cell loss in pigmentary glaucoma is comparable with that in POAG. The cell loss appears to be associated with obliteration of the highly conducting intertrabecular spaces or aqueous channels as they course through the corneoscleral meshwork. The aqueous channels are also reduced at their terminations near the juxtacanalicular apparatus where they form what the authors term cul-de-sacs. Measurements were made of the extent of aqueous channels in the trabecular meshwork and the area occupied by these cul-de-sacs. In normal eyes, it was reported that 94% of the surface area of the cul-de-sacs is lined by trabecular cells. The measurements were used to calculate the resistance to aqueous outflow offered by the cul-de-sacs. Three new concepts were subsequently advanced, specifically: (i) the cul-de-sacs provide a major portion of the normal outflow resistance; (ii) the cul-de-sac area is markedly reduced in pigmentary glaucoma and POAG, accounting for a major portion of the increase in resistance in these conditions; and (iii) macrophages are the major cell type responsible for trabecular meshwork clearance of pigment and debris. Thus, these authors propose a common pathophysiologic sequence of events for the development of pigmentary and POAG. However, most investigators still believe that the major site of resistance to outflow in both normal and POAG eyes lies in the juxtacanalicular region.491,492 Trabecular meshwork cellularity reportedly decreases with age,190,192 and trabecular cellularity in glaucoma is lower than in age-matched normals.193 In addition, the extracellular material present in the juxtacanalicular region increases with age.602 Numerous aqueous humor factors and metabolic products have been identified that could influence the synthesis of extracellular matrix components and adversely influence trabecular cell adhesion and proliferation in glaucoma.603

Cytoskeletal and junctional proteins may be especially important in the maintenance and modification of outflow resistance.604 Agents that interfere with dynamics of the actin cytoskeleton alter the cell shape, contractility and adhesion to neighboring cells and to the extracellular matrix in culture,325,326,605–614 and decrease trabecular outflow resistance in the living monkey eye.615–617 The lowered resistance is accompanied and perhaps caused by a pulling apart of cells in the region of the juxtacanalicular region and the separation of cells in the inner wall of Schlemm's canal, with subsequent ruptures in the inner wall and washout of extracellular material.618,619 Cytochalasin B, a fungal metabolite that inhibits actin polymerization and thereby destabilizes actin filaments,595 increases outflow facility in living monkeys595,619–622 when infused intracamerally. The facility increase is associated with distension of the meshwork, separation of meshwork cells from one another, breaks in the inner wall endothelium, and washout of extracellular material.618 Human trabecular cells in culture separate from one another when exposed to cytochalasin B614; when trabecular cells are grown on filters, the hydraulic conductivity of the preparation increases.614 Calcium chelating agents such as ethylenediaminetetraacetic acid (EDTA) or ethylene glycol-bis(β-aminoethyl ether)N,N,N',N'-tetraacetic acid (EGTA), which primarily affect cell junctions, produce similar cellular separation and physiologic consequences, including an increase in outflow facility in the living monkey.623–625

Recent studies have revealed a number of novel cytoskeletal agents that reduce outflow resistance in the living monkey or rabbit eye and/or the enucleated porcine eye probably by cytoskeleton-related mechanisms.615–617,626–636 With some agents, the lowered resistance is accompanied and perhaps caused by changes in cellular contractility in the trabecular meshwork (e.g., cellular relaxation) without apparent cell–cell separations. H-7, a serine-threonine kinase inhibitor, inhibits actomyosin-driven contractility and induces general cellular relaxation by inhibiting myosin light chain kinase or ρ kinase.605,637 Although H-7 does not directly affect actin polymerization, the inhibition of contractility leads to deterioration of the actin microfilament bundles and perturbation of its membrane anchorage at matrix adhesion sites in human trabecular meshwork and other cultured cells.604–608,638 In living monkeys, H-7 administered intracamerally or topically increases outflow facility (Fig. 20) and decreases intraocular pressure.615,616,639 Morphologic studies in the living monkey eye have shown that H-7 expands the intercellular spaces in the juxtacanalicular meshwork, accompanied by removal of extracellular material. The inner wall cells of Schlemm's canal become highly extended, yet cell-cell junctions are maintained (Fig. 21).497,498

Fig. 20. Effect of intracameral exchange (Ex) infusion with 10 to 500 μM H-7 on outflow facility in monkeys. BL, baseline; Res, reservoir. Data are mean ± standard error of the mean (SEM) μL/min/mm Hg for n animals, each contributing one H-7– and one vehicle-treated eye. Percentages show the increases of overall facility in H-7–treated eyes within 45 minutes post-drug perfusion, compared to contralateral vehicle-treated eyes and corrected for corresponding baselines (in B, item (a) represents the increase for the first 30 minutes, item (b) represents the increase for the second 30 minutes). *p < 0.05, **p < 0.005, ***p < 0.001 for ratios different from 1.0 by the two-tailed paired t test. (From Kaufman PL, Tian B, Gabelt BT, et al: Outflow enhancing drugs and gene therapy in glaucoma. In Weinreb R, Krieglstein G, Kitazawa Y, eds.: Glaucoma in the 21st Century. London: Harcourt-Mosby, 2000:117–128, with permission.)

Fig. 21. H-7 inhibits myosin light chain kinase and ρ kinase to block cellular actomyosin-driven contractility; leading to rapid deterioration of actin-containing stress fibers and focal contacts. A and B: Light micrographs (bars indicate 50 μm) of trabecular meshwork and Schlemm's canal in monkey eyes treated with H-7 ((1-[5-isoquinoline sulfonyl]-2-methyl piperazine), 300 μmol/L (B) or vehicle (A). The juxtacanalicular area (arrow in B) and intercellular spaces are extended, and extracellular material is lost. C and D: Schematic drawings depicting 15-cell stretches (cell–cell junctions marked by arrows) along the Schlemm's canal (SC) of control (C) and H-7–treated (D) eyes showing distribution of perfused gold particles the in juxtacanalicular area. The location of individual gold particles is represented by dots. Expanded areas are available for fluid drainage in H-7 treated eyes. (From Sabanay I, Gabelt BT, Tian B, et al: H-7 effects on structure and fluid conductance of monkey trabecular meshwork. Arch Ophthalmol 118:955, 2000, with permission.)

Compounds isolated from marine sponge macrolides, such as latrunculins A and B, alter cell shape and disrupt microfilament organization by sequestering G-actin, leading to disassembly of actin filaments.611–613 Latrunculins A and B increase outflow facility and decrease intraocular pressure in living monkeys.617,629,630 A preliminary morphological study in the living monkey eye has shown that latrunculin B induces massive ballooning of the juxtacanalicular region, leading to a substantial expansion of the space between the inner wall of Schlemm's canal and the trabecular collagen beams.640 No detrimental effects on tight junctions and cell–cell and cell–extracellular matrix adhesions are observed in the trabecular meshwork,640 although latrunculins interfere with cell–cell adhesions in cultured cells.611–613

Back to Top
OTHER AGENTS
Other agents that are known to affect trabecular outflow facility include α-chymotrypsin,623 ergotamine, angiotensin,641 chondroitinase ABC,642 and testicular643,644 or Streptomyces645 hyaluronidase.

Pharmacologic trabeculocanulotomy utilizing agents such as the ones discussed (cytochalasins, chelators, ethacrynic acid, α-chymotrypsin, chondroitinase ABC, hyaluronic acid, H-7, and latruncluin A/B) has been proposed as a possible strategy for the treatment of glaucoma since the 1970s.646 The sequence of events by which these agents decrease outflow resistance is being clarified as described for H-7 and latrunculin A/B above. Another approach to increase aqueous humor outflow is to inhibit or enhance the molecular pathways involved in regulating trabecular cell contractility by using gene therapy647–650 or to block cellular interactions with the extracellular environment that enhance actomyosin contractility and the formation of actin stress fibers.651 Long-term expression of genes that alter aqueous outflow has not yet been possible in the nonhuman primate eye.

Certain glucocorticoids administered topically or systemically to humans cause an elevation of IOP.337,652–654 Ocular hypertension can be induced in rabbits via the chronic systemic administration of glucocorticoids.655–657 The ability of glucocorticoids to alter IOP has been ascribed to their effect on the trabecular meshwork and aqueous humor outflow.658 Glucocorticoid receptors have been identified in the cells of the outflow pathways,659–661 but the biochemical and consequent physical processes causing the decrease in facility are poorly understood.

Studies suggest that glucocorticoids may play a major role in the normal physiologic regulation of outflow facility and IOP, perhaps by modulation of macromolecular metabolism or prostaglandin/adrenergic interactions involving the outflow system. As mentioned earlier (aqueous secretion section), the aqueous humor of control and POAG patients has levels of cortisol in excess of what is in the general circulation as a result of the activity of 11β-hydroxysteroid dehydrogenase 1 in the ciliary epithelium where it is involved in regulating aqueous secretion.213,214 The cortisol levels could also reduce aqueous outflow facility to a level that is detrimental in susceptible individuals.218 In the normal population, 34% to 42% of patients treated with topical or systemic corticosteroids are termed steroid responders, and develop markedly elevated IOP after several weeks.653,662 A similar percentage of nonhuman primates also develop ocular hypertension after topical dexamethasone treatment.663 This contrasts to patients with POAG, 90% of whom are considered steroid responders.653,662,664 The oral administration of the glucocorticoid biosynthesis inhibitor metyrapone to patients with glaucoma665 or the 11β-hydroxysteroid dehydrogenase inhibitor carbenoxolone to ocular hyperpertensive patients214 elicit small, transient reductions in IOP. A generalized cellular hypersensitivity to glucocorticoids is not intrinsic to POAG.666 Patients with glaucoma have increased plasma levels of cortisol compared to normal individuals.667,668

The glucocorticoid effect on the physiology of the outflow pathways may be more rapid than previously believed—certainly less than the 3 to 6 weeks classically described after topical eye drops.669 In cultured trabecular meshwork cells obtained from patients with POAG, cortisol metabolism is altered. These cells accumulate 5β-dihydrocortisol and, to a lesser extent, 5α-dihydrocortisol, metabolites that are not found to be present in cells derived from normal individuals. This difference is due to a marked increase in Δ4-reductase activity and to a decrease in 3-oxidoreductase activity not found in all cortisol-metabolizing cells.670,671 Southren672 demonstrated that topically applied 3α, 5β-tetrahydrocortisol (3α, 5β-THF), an intermediate metabolite of cortisol, decreases IOP and increases outflow facility in glaucomatous human eyes, and that 3β, 5β-THF antagonizes dexamethasone-induced cytoskeletal reorganization in normal human cultured trabecular meshwork cells.672,673 Interestingly, cultured trabecular meshwork cells from patients with POAG metabolized cortisol predominantly to 5β-dihydrocortisol (5β-DHF), which potentiates the facility-decreasing and IOP-increasing effects of dexamethasone; these cells produce relatively little 3α, 5β-THF from cortisol.672

Possible mechanisms for steroid-induced elevation of IOP have been proposed and include: accumulation or deposition of extracellular matrix material,674–679 decreased protease and stromelysin activities,680,681 reorganization of the trabecular meshwork cytoskeleton,682,683 increased nuclear size and DNA content,684 decreased phagocytic capacity,685 and changes in the synthesis of specific proteins.686 The progressive induction of one major steroid product in human trabecular meshwork cells matches the time course of clinical steroid effects on IOP and outflow facility. This molecule, known as TIGR or myocilin (MYOC), appears to be a secreted glycoprotein with aggregation- and extracellular matrix–binding groups687,688 interacting with extracellular components such as fibronectin.544,689,690 MYOC has also been localized intracellularly in the trabecular meshwork.689,691,692 Initially MYOC has been linked directly to both juvenile- and adult-onset open-angle glaucoma.693,694 MYOC mRNA and protein are present in a variety of ocular and nonocular tissues.695 Mutations in TIGR/MYOC, however, are not implicated in causing or increasing susceptibility to steroid-induced glaucoma.663 It has been suggested that MYOC is a stress protein.696,697 MYOC has an antiadhesive effect on trabecular meshwork cells in culture resulting in a loss of actin stress fibers and focal adhesions. Counter migratory effects on trabecular meshwork cells are also observed.698 These properties could enhance cell loss which occurs during phagocytic activities549 or oxidative stress.189 Conversely, mutations in the MYOC gene could reduce levels of secretion699,700 resulting in a tightening of the matrix attachments of trabecular meshwork cells that could diminish the flexibility needed for maintenance of normal outflow function.698

Other genes differentially upregulated in dexamethasone-treated HTM cells include a protease inhibitor (α1-antichymotrypsin), a neuroprotective factor (pigment epithelium-derived factor), and antiangiogenesis factor (cornea-derived transcript 6), and a prostaglandin synthase (prostaglandin D2 synthase).701 Microarray analysis of gene expression changes induced by dexamethasone in HTM cells show increases in five genes including MYOC, decorin, insulin-like growth factor binding protein 2, ferritin L chain, and fibulin-1C. Downregulated genes include nitric oxide synthase.702 Some of the functions of these gene products include regulation of the composition of the extracellular matrix, cell volume, and outflow resistance.

Cultured human trabecular meshwork cells exposed to dexamethasone also exhibit an unusual stacked arrangement of smooth and rough endoplasmic reticulum, proliferation of the Golgi apparatus, pleomorphic nuclei,683 and perhaps most intriguing in terms of outflow resistance, increased amounts of extracellular matrix material and unusual geodesic domelike cross-linked actin networks.682,683 Human trabecular meshwork cell monolayers grown on filters exhibit enhanced tight junction formation and decreased hydraulic conductivity in the presence of dexamethasone.703 Glucocorticoid glaucoma and POAG eyes both exhibit increased amounts of extracellular matrix material in the meshwork.674,704,705 However, the extracellular material that accumulates in eyes with corticosteroid-induced glaucoma differs from that seen in eyes with POAG.706

Prostaglandins are produced by human trabecular endothelial cells in culture, and dexamethasone inhibits trabecular cell prostaglandin synthesis by up to 90%.686,707 However, it has been argued that the dose-response relationship for the dexamethasone effect on prostaglandin synthesis is quite different from those for steroid effects on IOP, facility, and the MYOC induction.687,708–710

Prostaglandins and other eicosanoids in the trabecular meshwork may play important physiologic and pharmacologic roles in the aqueous outflow pathway. Weinreb,711 using radiolabeled arachidonic acid, demonstrated that prostaglandin E2 and prostaglandin F are produced as major cyclooxygenase products in cultured human trabecular meshwork cells, as well as 6-keto-prostaglandin F as a relatively minor product. Prostaglandins can increase outflow facility in monkeys and humans.269,712,713 However, the tonography and constant pressure perfusion techniques typically used to make these determinations are not specific for trabecular outflow, but rather are a combination of trabecular, uveoscleral and pseudofacility (IOP dependence of aqueous humor formation).

The protein profile of the normal human trabecular meshwork changes with age. Specifically, the α1 and α2 components of type IV collagen both increase with age, while a protein of molecular weight 31 kd starts to gradually disappear by age 31 years.714 Concomitant with the loss of α-sm (smooth muscle) actin filaments in the trabecular meshwork with age, an increase of synthesizing organelles such as rough endoplasmic reticulum was observed that could contribute to the age-related increase of extracellular material in this region. Both the increase in synthetic activity and loss of contractile protein might contribute to a decrease in outflow facility with age that is more pronounced with glaucoma.715 Fibronectin, an extracellular glycoprotein, plays a role in the cellular attachment to basement membrane and cell-matrix interaction.716,717 Quantitative morphometric evaluation of the trabecular drainage zone of normotensive eyes demonstrates a slow but significant rise in fibronectin content and concentration with aging,543,718 Additionally, a plaque material, derived from the sheaths of the elastic-like fibers in the cribriform layer of the trabecular meshwork, develops with age in the inner and outer wall of Schlemm's canal in both normal and glaucomatous eyes;598 the material is less evenly distributed around the circumference in glaucomatous eyes (with the exception of pseudoexfoliation glaucoma). Cross-linking of macromolecules limits their function and fate in cells, which lead to the hypothesis that cross-linking plays a mechanistic role in the aging process.719 Oxidative stress contributes to the morphologic and physiologic alterations in the aqueous outflow pathway in aging and glaucoma.720,721

Other factors in the aqueous humor may play a role in regulating trabecular meshwork extracellular matrix composition. Transforming growth factor (TGF)-β 2 is a component of normal aqueous humor detected in many mammalian eyes219–222 and may play a dominant role in glaucoma pathogenesis. TGFβ2 in the aqueous of diabetes plus POAG eyes is significantly higher than control aqueous; active TGFβ2 in the aqueous of POAG plus diabetes eyes is significantly higher than in controls and in patients with diabetes alone.722 Similarly increased levels of total and active TGFβ2 are found in the aqueous humor of POAG patients compared to age-matched controls.222,723 TGFβ2 enhancement of plasminogen activator inhibitor–1 expression inhibits the plasminogen/plasmin system necessary for activation of matrix metalloproteinases (MMP), thus decreasing MMP activity and possibly contributing to increased extracellular material in the trabecular meshwork of glaucomatous eyes.724 TGFβs have an inhibitory effect on the rate of cell proliferation and motility of trabecular meshwork cells in vitro that could contribute to the decreased cellularity of the trabecular meshwork.725 Perfusion of human anterior segments in vitro with TGFβ2 results in decreased outflow facility and increased focal accumulation of extracellular material under the inner wall of Schlemm's canal.726

As alluded to earlier (see the Aqueous Humor Flow section), long-term suppression of aqueous humor outflow can have a detrimental effect on the trabecular meshwork. A reduction in outflow facility occurs after long-term acetazolamide treatment of rabbits and humans.455 In cynomolgus monkeys treated with topical timolol for up to 7.4 months, underperfusion of the trabecular meshwork results in meshwork densification, activation of meshwork endothelial cells, and increased extracellular material within the cribriform region.456,727 More recently, a study was conducted in cynomolgus monkeys in which unilateral aqueous flow suppression was accomplished with timolol (β2 antagonist) plus dorzolamide (carbonic anhydrase inhibitor) and aqueous outflow in the same eye was redirected with topical prostaglandin F-ie (enhanced uveoscleral outflow, see next section). After 4 weeks, outflow facility was significantly lower in treated versus control eyes.728 The reduction in outflow facility was not correlated with any change in fibronectin levels in the aqueous humor.728

Back to Top
UVEOSCLERAL PATHWAY
The anterior chamber and the spaces within the trabecular meshwork are continuous with those between the ciliary muscle bundles. No epithelial or endothelial barrier separates them. Water and larger molecules from the anterior chamber can pass into and through the ciliary muscle via its anterior face, and from there, into the suprachoroidal space to be carried away, some perhaps by the choroidal vessels, but most actually through the sclera into the orbit.268,729 This pathway is called the uveoscleral pathway. It may be estimated as a relatively constant 0.5 μL/min,268 based on normal values for the other parameters in the modified Goldmann equation. In the young adult monkey, normally 40% to 70% of the aqueous is drained via the uveoscleral route.21,143–145,460,730,731 This decreases by about half in very old monkeys.731 The uveoscleral pathway was once thought to be less important in the human, based on direct measurements from very few eyes, most of which were in elderly persons and all of which were being enucleated for posterior segment tumors.729 Indirect measurements in young healthy conscious humans, although incorporating some assumptions, indicate that uveoscleral outflow may routinely account for nearly 50% of total aqueous drainage.278,459,732 This decreases somewhat with age459 although not as dramatically as found in monkeys.731 Aqueous draining via the uveoscleral route takes 2 hours or more145,268,460 to pass from the anterior chamber to the general circulation, because it has to negotiate first the supraciliary and suprachoroidal spaces, then the scleral emissarial channels, or the sclera itself, and finally the lymphatic vessels of the conjunctiva, orbit, or beyond to eventually drain into the systemic circulation. A small proportion of aqueous draining via the uveoscleral route may pass through the orbital fissure to be reabsorbed into the orbital intracranial blood vessels, once outside the eye. Conversely, aqueous draining via the trabecular route appears in the general circulation almost immediately.730

In the absence of drugs, outflow of aqueous via the uveoscleral route is virtually independent of IOP (i.e., uveoscleral flow is constant, within the physiologic range of IOP from approximately 10 to 40 mm Hg),238 unlike outflow via the trabecular pathway, and thus is not measurable by traditional methods for estimating facility of outflow (see elsewhere in this chapter). Presently it is believed that the driving force for uveoscleral flow is provided by the difference in pressure between the anterior chamber and the suprachoroidal space, which in the monkey is approximately 4 mm Hg. Within the range of IOP of approximately 10 to 40 mm Hg, a change in IOP is reflected by an equal change in pressure in the suprachoroidal space, thereby maintaining the pressure difference at a constant value.733 Uveoscleral outflow is decreased by muscarinic agonists such as pilocarpine and increased by muscarinic antagonists such as atropine. This is the opposite situation to that observed for the trabecular meshwork—Schlemm's canal route. Pilocarpine causes contraction of the ciliary muscle, squeezing the spaces between the ciliary muscle bundles, reducing the access of aqueous to this pathway. Atropine, on the other hand, opens access to this pathway.268 Fortunately, in the treatment of glaucoma, the effect of pilocarpine on the trabecular meshwork in facilitating outflow overshadows the decrease in uveoscleral outflow.

This system likely evolved to protect the eye in several ways during inflammation. The trabecular meshwork may become compromised by inflammation or obstructed by inflammatory debris, and the choroid may be overloaded with debris and extravasated proteins that must be removed from the eye.133 In this situation, prostaglandins would be released and, as autacoids or hormones that are synthesized, released, and locally acting, would induce the changes described. Because the eye has no lymphatics, uveoscleral outflow may serve as an analogue to an intraocular lymphatic drainage system.14 The normal low flow rate that is sufficient to remove normal levels of extravascular protein may be inadequate when protein levels are increased as in uveitis. Redirection of aqueous outflow from the trabecular to the uveoscleral pathway would both rid the eye of excess proteins and maintain physiologic IOP. This could also explain the low IOP that often accompanies uveitis; during experimental iridocyclitis in monkeys, uveoscleral outflow is increased approximately fourfold.340

Nilsson and colleagues145 reported a 60% increase in aqueous outflow via the uveoscleral pathway in monkeys after a single submaximal dose of prostaglandin F-1-isopropyl ester (PGF-ie). After multiple submaximal dosing, there was a greater than 100% increase in uveoscleral outflow.143 In both instances, aqueous outflow was substantially redirected from the trabecular to the uveoscleral pathway. The increase in uveoscleral outflow results from an enhancement of flow from the anterior chamber through the ciliary muscle and into the suprachoroid apparently results in part from relaxation of the muscle,574,575,734 but primarily from a remodeling of the extracellular matrix within the muscle. There is narrowing of the ciliary muscle fiber bundles, widening of the intermuscular spaces, and, perhaps most importantly, dissolution of collagen types I735 and III736 within the connective tissue-filled spaces between the outer longitudinally oriented muscle bundles.737,738 The latter probably results from PG-stimulated induction of MMP enzymes.739,740 PGF-ie and latanoprost have been shown to increase scleral MMP-1, MMP-2, and MMP-3 as well as scleral permeability and transscleral absorption of fibroblast growth factor-2.741,742

Prostaglandin analogues are the most potent and efficacious topical ocular hypotensive agents currently known for the treatment of human glaucoma.743 The most effective PGs for lowering IOP in humans are derivatives of PGF-ie, modified structurally to enhance ocular penetration and specifically activate the FP-prostanoid receptor. Side effects of early analogues included ocular irritation, conjunctival hyperemia and headache.142,744 These have been largely eliminated with latanaprost, a 17-phenyl-substituted isopropylester prodrug derivative of PGF-ie, which maintains a approximate 30% IOP reduction with once daily topical application of a 30-μL drop of 0.005% solution in ocular hypertensive patients with starting IOP of approximately 26 mm Hg.149,745–748 Latanoprost has also been shown to be effective in lowering IOP and with fewer side effects after once weekly dosing.749 Other analogues with similar IOP lowering efficacy but slightly higher prevalence of side effects include 0.03% bimatoprost and 0.004% travoprost.748,750

There is some controversy over whether or not bimatoprost is more efficacious than latanoprost and travoprost, perhaps as a result of enhancing outflow via the trabecular as well as the uveoscleral routes.748,751,752 Some studies suggest such a difference, others do not. Possible explanations for the discrepancies in these studies may be that the sample sizes are too small, so that differences in the populations of the different studies could either mask a small difference or make it more apparent. If there were a larger difference in efficacy between the drugs, it would be apparent in nearly all the studies, even if the magnitude of the difference varied because of these other factors. It is unlikely that such small and variable differences really matter clinically in terms of general approaches to treatment.753

Bimatoprost mildly stimulates aqueous humor flow and decreases tonographic resistance to outflow in normal human volunteers.754 In patients with ocular hypertension or glaucoma treated with bimatoprost, aqueous humor flow is unchanged while pressure-sensitive and pressure insensitive flow are increased,755 similar to earlier findings with latanoprost.149

Latanoprost, travoprost and bimatoprost are all FP agonists in HTM cells in vitro.756 In glaucomatous monkeys, bimatoprost and travoprost are additive with latanoprost in lowering IOP.757 Long-term (1 year) unilateral topical treatment of normotensive cynomolgus monkeys with bimatoprost, latanoprost, sulprostone (EP3/EP1 agonist) or AH13205(EP2 agonist) results in similar morphological changes in all groups as well as in the contralateral untreated eyes. Uveoscleral outflow pathways are enlarged and appear organized. More myelinated nerve fiber bundles are found. Changes in the trabecular meshwork are also noted.758

The increase in uveoscleral outflow in response to these compounds is so great that a larger pharmacologic reduction in IOP is possible than with any other known substance. It has yet to be established, however, whether or not endogenous prostaglandins have a physiologic role in regulating uveoscleral outflow, or whether they play only a pathophysiologic role.

In ocular hypertensive patients, chronic treatment with brimonidine lowers IOP initially by decreasing aqueous flow and, after chronic treatment, by increasing uveoscleral outflow.469,759 No other studies have examined the effect of long-term administration of brimonidine on these parameters.

Epinephrine, in addition to increasing trabecular outflow, also increases uveoscleral outflow in monkeys276 and humans.278,353 The mechanism of this phenomenon is unknown. It may in part be due to the mildly relaxant effect of epinephrine on the ciliary muscle, presumably acting via its β-adrenergic receptors.573–576 However, adrenergic agonists also stimulate prostaglandin biosynthesis in several tissues, including rabbit760 and bovine761 iris. Pretreatment with the cyclooxygenase inhibitor indomethacin inhibits the ocular hypotensive effect of topically applied epinephrine in humans762 suggesting that the IOP-lowering action of epinephrine may be mediated at least in part by prostaglandins or other cyclooxygenase products.762,763

Back to Top
ACKNOWLEDGMENTS
Supported by National Eye Institute grant EY02698, Research to Prevent Blindness, and the Ocular Physiology Research and Education Fund.
Back to Top
REFERENCES

1. Davson H: The aqueous humor and the intraocular pressure. In Davson H, ed.: , ed.: Physiology of the Eye. New York: Pergamon Press, 1990:3–95

2. Freddo TF: Shifting the paradigm of the blood-aqueous barrier. Exp Eye Res 73:581, 2001

3. Brubaker RF: Flow of aqueous humor in humans [The Friedenwald Lecture]. Invest Ophthalmol Vis Sci 32:3145, 1991

4. Hogan MJ, Alvarado JA, Weddell JE: Histology of the Human Eye. An Atlas and Textbook. Philadelphia: WB Saunders Co., 1971:136–153, 260–319

5. Cole DF: Aqueous humor formation. Doc Ophthalmol 21:116, 1966

6. Alm A, Bill A: Ocular and optic nerve blood flow at normal and increased intraocular pressure in monkeys (Macaca Irus): A study with radioactively labelled microspheres, including flow determinatins in brain and some other tissues. Exp Eye Res 15:15, 1973

7. Morrison JC, Van Buskirk EM: Anterior collateral circulation in the primate eye. Ophthalmology 90:707, 1983

8. Morrison JC, Van Buskirk EM: Ciliary process microvasculature. Am J Ophthalmol 97:372, 1984

9. Morrison JC, van Buskirk EM: Sequential microdissection and scanning electron microscopy of ciliary microvascular castings. Scan Electron Microsc 2:857, 1984

10. Caprioli J, Sears ML, Mead A: Ocular blood flow in phakic and aphakic monkey eyes. Exp Eye Res 39:1, 1984

11. Morrison JC, de Frank MP, van Buskirk EM: Comparative microvascular anatomy of mammalian ciliary processes. Invest Ophthalmol Vis Sci 28:1325, 1987

12. Morrison JC, de Frank MP, van Buskirk EM: Regional microvascular anatomy of the rabbit ciliary body. Invest Ophthalmol Vis Sci 28:1314, 1987

13. Funk R, Rohen JW: Scanning electron microscopic study on the vasculature of the human anterior eye segment, especially with respect to the ciliary processes. Exp Eye Res 51:651, 1990

14. Bill A: Blood circulation and fluid dynamics in the eye. Physiol Rev 55:383, 1975

15. Seidël E: Weitre experimentelle Untersuchungen über die Quelle und den Verlauf der introkulären Saftsrömung. XII. Metteilung. Uber den manometrischen Nachweis des physiologischen Druckgefalles zwishen Vorderkammer und Schlemmshen Kanal. Graefe's Arch Clin Exp Ophthalmol 107:101, 1921

16. Seidël E: Weitre experimentelle Untersuchungen über die Quelle und den Verlauf der introkulären Saftströmung. IX. Uber der Abfluss des Kammerwassers aus der vorderen Augenkammer. Graefe's Arch Clin Exp Ophthalmol 104:357, 1921

17. Knutson SL, Sears ML: Herman Boerhaave and the history of vessels carrying aqueous humor from the eye. Am J Ophthalmol 76:648, 1973

18. Ascher KW: Aqueous veins: Preliminary note. Am J Ophthalmol 25:31, 1942

19. Goldmann H: Enhalten die Kammervasservener Kammervasser? Ophthalmologica 117:240, 1949

20. Ashton N: Anatomical study of Schlemm's canal and aqueous veins by means of neoprene casts: Part I. Aqueous veins. Br J Ophthalmol 35:291, 1951

21. Bill A: The aqueous humor drainage mechanism in the cynomolgus monkey (Macaca irius) with evidence for unconventional routes. Invest Ophthalmol 4:911, 1965

22. Bill A: The role of ciliary blood flow and ultrafiltration in aqueous humor formation. Exp Eye Res 16:287, 1973

23. Brodwell J, Fischbarg J: The hydraulic conductivity of rabbit ciliary epithelium in vitro. Exp Eye Res 34:121, 1982

24. Caprioli J: The ciliary epithelia and aqueous humor. In Hart WM, ed.: Adler's Physiology of the Eye: Clinical Application, 9th ed. St. Louis: CV Mosby, 1992:228–247

25. Cole DF: Effects of some metabolic inhibitors upon the formation of the aqueous humor in rabbits. Br J Ophthalmol 44:739, 1960

26. Cole DF: Secretion of the aqueous humor. Exp Eye Res 25(suppl):161, 1977

27. Green K, Pederson JE: Contribution of secretion and filtration to aqueous humor formation. Am J Physiol 222:1218, 1972

28. Maren TH: HCO3-formation in aqueous humor: Mechanism and relation to the treatment of glaucoma. Invest Ophthalmol 13:479, 1974

29. Pederson JE: Fluid permeability of monkey ciliary epithelium in vivo. Invest Ophthalmol Vis Sci 23:176, 1982

30. Duke-Elder S: The aqueous humor. In Duke-Elder S, ed.: The Physiology of the Eye and of Vision System of Ophthalmology, Vol. 4. St. Louis: CV Mosby, 1968:104–200

31. Saier MH: Neurotransmission. In Zubay G, ed.: , ed.: Biochemistry. Menlo Park, CA: Addison-Wesley, 1983:1147–1167

32. Chattoraj SC, Watts NB: Endocrinology. In Tietz NW, ed.: Textbook of Clinical Chemistry. Philidelphia: WB Saunders, 1986:1116–1117

33. Weinbaum S, Langham ME, Goldgraben JR, et al: The role of secretion and pressure-dependent slow in aqueous humor formation. Exp Eye Res 13:266, 1972

34. Green K, Pederson JE: Aqueous humor formation. Exp Eye Res 16:273, 1973

35. Bonting SL, Becker B: Studies on sodium-potassium activated adenosinetriphosphatase: XIV. Inhibition of enzyme activity and aqueous humor flow in the rabbit eye after intravitreal injection of ouabain. Invest Ophthalmol 3:523, 1964

36. Becker B: Ouabain and aqueous humor dynamics in the rabbit eye. Invest Ophthalmol 2:325, 1963

37. Becker B: Vanadate and aqueous humor dynamics. Proctor Lecture. Invest Ophthalmol Vis Sci 19:1156, 1980

38. Krupin T, Becker B, Podos SM: Topical vanadate lowers intraocular pressure in rabbits. Invest Ophthalmol Vis Sci 19:1360, 1980

39. Podos SM, Lee PY, Severin C, et al: The effect of vanadate on aqueous humor dynamics in cynomolgus monkeys. Invest Ophthalmol Vis Sci 25:359, 1984

40. Shahidullah M, Wilson WS, Yap M, et al: Effects of ion transport and channel-blocking drugs on aqueous humor formation in isolated bovine eye. Invest Ophthalmol Vis Sci 44:1185, 2003

41. Hochgesand DH, Dunn JJ, Crook RB: Catecholaminergic regulation of Na-K-Cl contransport in pigmented ciliary epithelium: Difference between PE and NPE. Exp Eye Res 72:1, 2001

42. Friedland BR, Maren TH: Carbonic anhydrase: Pharmacology of inhibitors and treatment of glaucoma. In Sears ML, ed.: , ed.: Pharmacology of the Eye. Berlin: Springer-Verlag, 1984:279–309

43. Brechue WF, Maren TH: A comparison between the effect of topical and systemic carbonic anhydrase inhibotors on aqueous humor secretion. Exp Eye Res 57:67, 1993

44. Pierce WMJ, Sharir M, Waite KJ: Topically active ocular carbonic anhydrase inhibitors-novel biscarbonylamidothiadiazole sulfonamides as ocular hypotensive agents. Proc Soc Exp Biol Med 203:360, 1993

45. Becker B: Hypothermia and aqueous humor dynamics of the rabbit eye. Trans Am Ophthalmol Soc 58:337, 1960

46. Bonting SL, Simon KA, Hawkins NM: Studies on sodium-potassium-activated adenosine triphosphatase. I. Quantitative distribution in several tissues of the cat. Arch Biochem 95:416, 1961

47. Riley MV: The sodium-potassium-stimulated adensosine triphosphatase of rabbit ciliary epithelium. Exp Eye Res 3:76, 1964

48. Riley MV, Kishida K: ATPases of ciliary epithelium: Cellular and subcellular distribution and probable role in secretion of aqueous humor. Exp Eye Res 42:559, 1986

49. Coca-Prados M, Lopez-Briones LG: Evidence that the a and the a(+) isoforms of the catalytic subunit of (Na+, K+)-ATPase reside in distinct ciliary epithelial cells of the mammalian eye. Biochem Biophys Res Commun 145:460, 1987

50. Flügel C, Lütjen-Drecoll E: Presence and distribution of Na+/K+-ATPase in the ciliary epithelium of the rabbit. Histochemistry 88:613, 1988

51. Usukura J, Fain GL, Bok D: [3H]Ouabain localization of Na-K ATPase in the epithelium of the rabbit ciliary body pars plicata. Invest Ophthalmol Vis Sci 29:606, 1988

52. Holland MG, Gipson CC: Chloride ion transport in the isolated ciliary body. Invest Ophthalmol Vis Sci 9:20, 1970

53. Cole DF: Electrochemical changes associated with the formation of the aqueous humor. Br J Ophthalmol 45:202, 1961

54. Cole DF: Evidence for active transport of chloride in ciliary epithelium of the rabbit. Exp Eye Res 8:5, 1969

55. Kinsey VE, Reddy DVN: Chemistry and dynamics of aqueous humor. In Prince JH, ed.: , ed.: The Rabbit in Eye Research. Springfield, IL: Charles C. Thomas, 1964:218–319

56. Krupin T, Reinach PS, Candia OA, et al: Transepithelial electrical measurements on the isolated rabbit irisciliary body. Exp Eye Res 38:115, 1984

57. Do CW, To CH: Chloride secretion by bovine ciliary epithelium: A model of aqueous humor formation. Invest Ophthalmol Vis Sci 41:1853, 2000

58. Verkman AS: Role of aquaporin water channels in eye function. Exp Eye Res 76:137, 2003

59. Patil RV, Han Z, Yiming M, et al: Fluid transport by human nonpigmented ciliary epithelial layers in culture: A homeostatic role for aquaporin-1. Am J Physiol 281:1139, 2001

60. Frigeri A, Gropper M, Turck CW, et al: Immunolocalization of the mercurial-insensitive water channel and glycerol intrinsic protein in epithelial cell plasma membranes. Proc Nat Acad Sci 92:4328, 1995

61. Wistrand PJ: Carbonic anhydrase in the anterior uvea of the rabbit. Acta Physiol Scand 24:144, 1951

62. Ballintine EJ, Maren TH: Carbonic anhydrase activity and the distribution of Diamox in the rabbit eye. Am J Ophthalmol 40:148, 1955

63. Bhattacherjee P: Distribution of carbonic anhydrase in the rabbit eye as demonstrated histochemically. Exp Eye Res 12:356, 1971

64. Tsukahara S, Maezara N: Cytochemical localization of adenyl cyclase in the rabbit ciliary body. Exp Eye Res 26:99, 1978

65. Mishima H, Sears M, Bausher L, et al: Ultracytochemistry of cholera toxin binding sites in ciliary processes. Cell Tissue Res 223:241, 1982

66. White A, Handler P, Smith EL: Introduction to metabolism. In White A, Handler P, Smith EL, eds.: Principles of Biochemistry, 5th ed. New York: McGraw-Hill, 1973:279–310

67. Mudge GH, Weiner IM: Agents affecting volume and composition of body fluids. In Goodman A, Rall TW, Nies AS, Taylor P, eds.: The Pharmacological Basis of Therapeutics, 8th ed. New York: McGraw-Hill, 1990:682–731

68. Muther TF, Friedland BR: Autoradiographic localization of carbonic anhydrase in the rabbit ciliary body. J Histochem Cytochem 28:1119, 1980

69. Lütjen-Drecoll E, Lonnerholm G: Carbonic anhydrase distribution in the rabbit eye by light and electron microscopy. Invest Ophthalmol Vis Sci 21:782, 1981

70. Lütjen-Drecoll E, Lonnerholm G, Eichhorn M: Carbonic anhydrase distribution in the human and monkey eye by light and electron microscopy. Graefe's Arch Clin Exp Ophthalmol 220:285, 1983

71. Maren TH, Mayer E, Wadsworth BC: Carbonic anhydrase inhibition. I. The pharmacology of Diamox (1-acetylamino 1,3,4-thiadiazole-5-sulfonamide). Bull Johns Hopkins Hosp 95:199, 1954

72. Maren TH: Carbonic anhydrase: Chemistry, physiology, and inhibition. Physiol Rev 47:595, 1967

73. Maren TH: The development of ideas concerning the role of carbonic anhydrase in the secretion of aqueous humor: Relations to the treatment of glaucoma. In Drance SM, Neufeld AH, eds.: Glaucoma Applied Pharmacology in Medical Treatment. Orlando: Grune & Stratton; 1984:325–555

74. Eller MG, Schoenwald RD, Dixson JA: Topical carbonic anhydrase inhibitors, III. Optimization model for corneal penetration of ethoxzolamide analogues. J Pharm Sci 74:155, 1985

75. Maren TH, Jankowska L, Sanyal G, et al: The transcorneal permeability of sulfonamide carbonic anhydrase inhibitors and their effect on aqueous humor secretion. Exp Eye Res 36:457, 1983

76. Schoenwald RD, Eller MG, Dixson JA, et al: Topical carbonic anhydrase inhibitors. J Med Chem 27:810, 1984

77. Lippa EA, Carlson LE, Ehinger B, et al: Dose response and duration of action of dorzolamide, a topical carbonic anhydrase inhibitor. Arch Ophthalmol 110:495, 1992

78. Wilkerson M, Cyrlin M, Lippa EA, et al: Four-week safety and efficacy study of dorzolamide, a novel, active topical carbonic anhydrase inhibitor. Arch Ophthalmol 111:1343, 1993

79. Gunning FP, Greve EL, Bron AM, et al: Two topical carbonic anhydrase inhibitors sezolamide and dorzolamide in Gelrite vehicle: A multiple-dose efficacy study. Graefes Arch Clin Exp Ophthalmol 231:384, 1993

80. Herkel U, Pfeiffer N: Update on topical carbonic anhydrase inhibitors. Curr Opin Ophthalmol 12:88, 2001

81. Cvetkovic RS, Perry CM: Brinzolamide: A review of its use in the management of primary open-angle glaucoma and ocular hypertension. Drugs Aging 20:919, 2003

82. DeSantis L: Preclinical overview of brinzolamide. Surv Ophthalmol 44:S119, 2000

83. Wistrand P: Local action of the carbonic anhydrase inhibitor, acetazolamide, on the intraocular pressure in cats. Acta Pharm Toxicol 14:27, 1957

84. Macri FJ: The constrictive action of acetazolamide on the arteries of the cat. Arch Ophthalmol 66:570, 1961

85. Macri FJ, Cevario SJ: A possible vascular mechanism for the inhibition of aqueous humor formation by ouabain and acetazolamide. Exp Eye Res 20:563, 1975

86. Bill A: Effects of acetazolamide and carotid occlusion on the ocular blood flow in unanesthetized rabbits. Invest Ophthalmol 13:954, 1974

87. Maren TH: Biochemistry of aqueous humor inflow. In Kaufman PL, Mittag TW, eds.: Glaucoma. London: Mosby-Year Book Europe Ltd., 1994:1.35–1.46

88. Maren TH: The rates of movement of Na+, Cl, and HCO3 from plasma to posterior chamber: Effect of acetazolamide and relation to the treatment of glaucoma. Invest Ophthalmol 15:356, 1976

89. Becker B: The mechanism of the fall in intraocular pressure induced by the carbonic anhydrase inhibitor Diamox. Am J Ophthalmol 39:177, 1955

90. Kaufman PL, Mittag TW: Medical therapy of glaucoma. In Kaufman PL, Mittag TW, eds.: Glaucoma. London: Mosby-Year Book Europe Ltd., 1994:9.7–9.30

91. Sugrue MF, Mallorga P, Schwamm H: A comparison of L-671,152 and MK927, two topically effective ocular hypotensive carbonic anhydrase inhibitors, in experimental animals. Curr Eye Res 9:607, 1990

92. Wang RF, Serle JB, Podos SM, et al: MK-507 (L 671, 152), a topically active carbonic anhydrase inhibitor, reduces aqueous humor production in monkeys. Arch Ophthalmol 109:1297, 1991

93. Maren TH, Bar-Ilan A, Conroy CW, et al: Chemical and pharmacological properties of MK-927, a sulfonamide carbonic anhydrase inhibitor tha lowers intraocular pressure by the topical route. Exp Eye Res 50:27, 1990

94. Maus TL, Larsson LI, McLaren JW, et al: Comparison of dorzolamide and acetazolamide as suppressors of aqueous humor flow in humans. Arch Ophthalmol 115:45, 1997

95. Michaud JE, Friren B: Comparison of topical brinzolamide 1% and dorzolamide 2% eye drops given twice daily in addition to timolol 0.5% in patients with primary open-angle glaucoma or ocular hypertension. Am J Ophthalmol 132:235, 2001

96. Seong GJ, Lee SC, Lee JH, et al: Comparisons of intraocular-pressure-lowering efficacy and side effects of 2% dorzolamide and 1% brinzolamide. Ophthalmologica 215:188, 2001

97. Krause U, Raunio V: Proteins of the normal human aqueous humor. Ophthalmologica 159:178, 1969

98. Krause U, Raunio V: The proteins of the normal aqueous humor. Ophthalmologica 160:280, 1970

99. Rodriguez-Peralta L: The blood-aqueous barrier in five species. Am J Ophthalmol 80:713, 1975

100. Bill A: The drainage of albumin from the uvea. Exp Eye Res 3:179, 1964

101. Bill A: Capillary permeability to and extravascular dynamics of myoglobin, albumin, and gammaglobulin in the uvea. Acta Physiol Scand. 73:204, 1968

102. Vegge T: An epithelial blood-aqueous barrier to horseradish peroxidase in the ciliary processes of the vervet monkey (Cercopithecus aethiops). Zeitschr Zellforsch Mikrosk Anat 114:309, 1971

103. Shiose Y: Electron microscopic studies on blood-retinal and blood-aqueous barriers. Jpn J Ophthalmol 14:73, 1970

104. Smith RS: Ultrastructural studies of the blood-aqueous barrier. 1. Transport of an electron dense tracer in the iris and ciliary body of the mouse. Am J Ophthalmol 71:1066, 1971

105. Uusitalo R, Palkama A, Stjernschantz J: An electron microscopic study of the blood-aqueous barrier in the ciliary body and iris of the rabbit. Exp Eye Res 17:49, 1973

106. Uusitalo R, Stjernschantz J, Palkama A: Studies on the ultrastructure of the blood-aqueous barrier in the rabbit. Acta Ophthalmol 123:61, 1974

107. Freddo T, Raviola G: The homogeneous structure of blood vessels in the vascular tree in Macaca mulatta iris. Invest Ophthalmol Vis Sci 22:279, 1982

108. Hirsch M, Montcourrier P, Arguillere P, et al: The structure of tight junctions in the ciliary epithelium. Curr Eye Res 4:493, 1995

109. Shabo AL, Maxwell DS: The blood-aqueous barrier to tracer protein. A light and electron microscopic study of the primate ciliary process. Microvasc Res 4:142, 1972

110. Smith RS, Rudt LA: Ultrastructural studies of the blood-aqueous barrier 2. The barrier to horseradish peroxidase in primates. Am J Ophthalmol 76:937, 1973

111. Vegge T: An electron microscopic study of the permeability of iris capillaries to horseradish peroxidase in the vervet monkey. Z Zellforsch. 121:74, 1971

112. Davson H, Duke-Elder WS, Maurice DM: Changes in ionic distribution following dialysis of aqueous humor against plasma. J Physiol (Lond) 109:32, 1949

113. Ross EJ: The transfer of non-electrolytes across the blood-aqueous barrier. J Physiol (Lond) 112:229, 1951

114. Davson H, Matchett PA: The kinetics of penetration of the blood-aqueous barrier. J Physiol (Lond) 122:11, 1953

115. Hart WM: Intraocular pressure. In Hart WM, ed.: , ed.: Adler's Physiology of the Eye, 9th ed. St. Louis: CV Mosby, 1992:248–267

116. Bairati AJ, Orzalesi N: The ultrastructure of the epithelium of the ciliary body: A study of the junctional complexes and of the changes associated with the production of plasmoid aqueous humor. Z Zellforsch Milrosk Anat 69:635, 1966

117. Okisaka S: Effects of paracentesis on the blood-aqueous barrier: A light and electron microscopic study on cynomolgus monkey. Invest Ophthalmol 15:824, 1976

118. Al-Ghadyan A, Mead A, Sears ML: Increased pressure after paracentesis of the rabbit eye is completely accounted for by prostaglandin synthesis and release plus pupillary block. Invest Ophthalmol Vis Sci 18:361, 1979

119. Bartels SP, Pederson JE, Gaasterland DE, et al: Sites of breakdown of the blood-aqueous barrier after paracentesis of the rhesus monkey eye. Invest Ophthalmol Vis Sci 18:1050, 1979

120. Ohnishi Y, Tanaka M: Effects of pilocarpine and paracentesis on occluding junctions between the nonpigmented ciliary epithelial cells. Exp Eye Res 32:635, 1981

121. Gaasterland DF, Barranger JA, Rapoport SI: Long-term ocular effects of osmotic modification of the blood-brain barrier in monkeys. 1. Clinical examination, aqueous ascorbate and protein. Invest Ophthalmol Vis Sci 24:153, 1983

122. Miyake Y, Asakura M, Maekubo K: Consensual reactions of human blood-aqueous barrier to implant operations. Arch Ophthalmol 102:558, 1984

123. Kondo K, Coca-Prados M, Sears ML: Human ciliary epithelia in monolayer culture. Exp Eye Res 102:423, 1984

124. Jampol L, Neufeld A, Sears M: Pathways for the response of the eye to injury. Invest Ophthalmol Vis Sci 14:184, 1975

125. Shabo AL, Maxwell DS, Kreiger AE: Structural alterations in the ciliary process and the blood-aqueous barrier of the monkey after systemic urea injections. Am J Ophthalmol 81:162, 1976

126. Raviola G: Blood-aqueous barrier can be circumvented by lowering intraocular pressure. Proc Natl Acad Sci USA 73:638, 1976

127. Raviola G: Effects of paracentesis on the blood-aqueous barrier: An electron microscopic study on Macaca mullata using horseradish peroxidase as a tracer. Invest Ophthalmol 13:828, 1974

128. Whitelocke RA, Eakins KE: Vascular changes in the anterior uvea of the rabbit produced by prostaglandins. Arch Ophthalmol 89:495, 1973

129. Whitelocke RAF, Eakins KE, Bennett A: Acute anterior uveitis and prostaglandins. Proc R Soc Med 66:429, 1973

130. Green K, Kim K: Pattern of ocular response to topical and systemic prostaglandin. Invest Ophthalmol 14:36, 1975

131. Laties AM, Neufeld AH, Vegge T, et al: Differential reactivity of rabbit iris and ciliary process to topically applied prostaglandin-E2 (dinoprostone). Arch Ophthalmol 94:1966, 1976

132. Crawford K, Kaufman PL, True-Gabelt B: Prostaglandins and aqueous humor dynamics. In Shields MB, Pollack IP, Kolker AE, eds.: Perspectives in Glaucoma Transactions of the First Scientific Meeting of the American Glaucoma Society. Thorofare: Slack, Inc., 1988:259–267

133. Kaufman PL, Crawford K, Gabelt BT: The effects of prostaglandins on aqueous humor dynamics. In Kooner KS, Zimmerman TJ, eds.: New Ophthalmic Drugs. Philadelphia: WB Saunders, 1989:141–150

134. Kaufman PL, Rohen JW, Gabelt BT, et al: Parasympathetic denervation of the ciliary muscle following panretinal photocoagulation. Curr Eye Res 10:437, 1991

135. Green K: Permeability properties of the ciliary epithelium in response to prostaglandins. Invest Ophthalmol 12:752, 1973

136. Neufeld AH, Sears ML: Prostaglandin and eye. Prostaglandins 4:157, 1973

137. Neufeld AH, Jampol LM, Sears ML: Aspirin prevents the disruption of the blood-aqueous barrier in the rabbit eye. Nature 238:158, 1972

138. Bhattacherjee P: Autoradiographic localization of intravitreally or intracamerally injected [3H]-prostaglandins. Exp Eye Res 18:181, 1974

139. Bhattacherjee P, Hammond BR: Inhibition of increased permeability of the blood-aqueous barrier by nonsteroidal anti-inflammatory compounds as demonstrated by fluorescein angiography. Exp Eye Res 21:499, 1975

140. Bengtsson E: The effect of imidazole on the disruption of the blood-aqueous barrier in the rabbit eye. Invest Ophthalmol 15:315, 1976

141. Chavis RM, Vygantis CM, Vygantis A: Experimental inhibition of prostaglandin-like inflammatory response after cryotherapy. Am J Ophthalmol 82:310, 1976

142. Giuffré G: The effects of prostaglandin F2a in the human eye. Graefe's Arch Clin Exp Ophthalmol 222:139, 1985

143. Gabelt BT, Kaufman PL: Prostaglandin F2a increases uveoscleral outflow in the cynomolgus monkey. Exp Eye Res 49:389, 1989

144. Kaufman PL, Crawford K: Aqueous humor dynamics: How PGF lowers intraocular pressure. Prog Clin Biol Res 312:387, 1989

145. Nilsson SFE, Samuelsson M, Bill A, et al: Increased uveoscleral outflow as a possible mechanism of ocular hypotension caused by prostaglandin F2a-1–isopropylester in the cynomolgus monkey. Exp Eye Res 48:707, 1989

146. Villumsen J, Alm A: Prostaglandin F2a-isopropylester eye drops: Effects in normal human eyes. Br J Ophthalmol 73:419, 1989

147. Camras CB, Siebold EC, Lustgarten JS, et al: Maintained reduction of intraocular pressure by prostaglandin F2a-1–isopropylester applied in multiple doses in ocular hypertensive and glaucoma patients. Ophthalmology 96:1329, 1989

148. Kobayashi H, Kobayashi K, Okinami S: A comparison of intraocular pressure-lowering effect of prostaglandin F2-alpha analogues, latanoprost and unoprotone isopropyl. J Glaucoma 10:487, 2001

149. Ziai N, Dolan JW, Kacere RD, et al: The effects on aqueous dynamics of PhXA41, a new prostaglandin F2a Analogue, after topical application in normal and ocular hypertensive human eyes. Arch Ophthalmol 111:1351, 1993

150. Alm A, Villumsen J, Tornquist P, et al: Intraocular-pressure-reducing effect of PhXA41 in patients with increased eye pressure. A one-month study. Ophthalmology 100:1312, 1993

151. Alm A, Widengard I, Kjellgren D, et al: Latanoprost administered once daily caused a maintained reduction of intraocular pressure in glaucoma patients treated concomitantly with timolol. Br J Ophthalmol 79:12, 1995

152. Toris CB, Camras CB, Yablonski ME: Effects of PhXA41, a new prostaglandin F2 alpha, on aqueous humor dynamics in human eyes. Ophthalmology 100:1297, 1993

153. Racz P, Ruzsonyi MR, Nagy ZT, et al: Maintained intraocular pressure reduction with once-a-day application of a new prostaglandin-F2-alpha analogue (PhXA41)—An in-hospital, placebo-controlled study. Arch Ophthalmol 111:657, 1993

154. Cantor LB: Bimatoprost: A member of a new class of agents, the prostamides, for glaucoma management. Expert Opin Invest Drugs. 10:721, 2001

155. Brubaker RF: Mechanism of action of bimatoprost (Lumigan). Surv Ophthalmol 45(suppl)S347, 2001

156. Jaffe NS, Jaffe MS, Jaffe GF: Cataract surgery and its complications, 5th ed. St. Louis: CV Mosby; 1990:148

157. Benham GH, Duke-Elder WS, Hodgson TH: The osmotic pressure of the aqueous humor in the normal and glaucomatous eye. J Physiol (Lond) 92:355, 1938

158. Roepke RR, Hetherington WA: Osmotic relation between aqueous humor and blood plasma. Am J Physiol 130:340, 1940

159. Kinsey VE: The chemical composition and the osmotic pressure of the aqueous humor and plasma of the rabbit. J Gen Physiol 34:389, 1950

160. Cole DF: Electrolyte composition of anterior and posterior aqueous humor in the sheep. Ophthalmic Res 4:1, 1973

161. Stjernschantz J, Uusitalo R, Palkama A: The aqueous proteins of the rat in the normal eye and after aqueous withdrawal. Exp Eye Res 16:215, 1973

162. Ghose T, Quigley JH, Landrigan PL, et al: Immunoglobulins in aqueous humor and iris from patients with endogenous uveitis and patients with cataract. Br J Ophthalmol 57:897, 1973

163. Sen DK, Saren GS, Saha K: Immunoglobulins in human aqueous humor. Br J Ophthalmol 61:216, 1977

164. Fielder AR, Rahi AHS: Immunoglobulins of normal aqueous humor. Trans Ophthalmol Soc UK 99:120, 1979

165. Mondino BJ, Rao H: Complement levels in normal and inflamed aqueous humor. Invest Ophthalmol Vis Sci 24:380, 1983

166. Pandolfi M, Nilsson IM, Martinsson G: Coagulation and fibrinolytic components in primary and plasmoid aqueous humor. Acta Ophthalmol 42:820, 1964

167. Sandberg HO, Class O: The alpha and gamma crystallin content in aqueous humor of eyes with clear lens and with cataracts. Exp Eye Res 28:601, 1979

168. Fujiwara H: Lens crystallin reactive protein in the aqueous humor of cataract patients. Jpn J Ophthalmol 33:418, 1989

169. Kolodny NH, Freddo TF, Lawrence B, et al: Contrast-enhanced MRI confirmation of an anterior protein pathway in normal rabbit eyes. Invest Ophthalmol Vis Sci 37:1602, 1996

170. Bert R, Freddo T, Caruthers SD, et al: Confirmation of anterior large-molecule diffusion pathway in the normal human eye. Invest Ophthalmol Vis Sci 40(suppl):S198, 1999

171. Freddo TF, Bartels SP, Barsotti MF, et al: The source of proteins in the aqueous humor of the normal rabbit. Invest Ophthalmol Vis Sci 31:125, 1990

172. Barsotti M, Bartels SP, Barsotti MF, et al: The source of proteins in the aqueous humor of the normal monkey eye. Invest Ophthalmol Vis Sci 33:581, 1992

173. Kee C, Gabelt BT, Gange SJ, et al: Serum effects on aqueous outflow during anterior chamber perfusion in monkeys. Invest Ophthalmol Vis Sci 37:1840, 1996

174. Johnson M, Gong H, Freddo TF, et al: Serum proteins and aqueous outflow resistance in bovine eyes. Invest Ophthalmol Vis Sci 34:3549, 1993

175. Sit AJ, Gong H, Ritter N, et al: The role of soluble proteins in generating aqueous outflow resistance in the bovine and human eye. Exp Eye Res 64:813, 1997

176. Reddy DVN, Kinsey VE: Chemistry and dynamics of aqueous humor. In Prince JH, ed.: The Rabbit in Eye Research: Springfield: CC Thomas, 1966:218

177. Bito LZ, Davson H, Levin E: The relationship between the concentrations of amino acids in the ocular fluids and blood plasma of dogs. Exp Eye Res 4:374, 1965

178. Ehlers N, Kristensen K, Schonheyder F: Amino acid transport in human ciliary epithelium. Acta Ophthalmol 56:777, 1978

179. Kinsey VE: Further study of the distribution of chloride between lasma and intraocular fluids in the rabbit eye. Invest Ophthalmol 6:395, 1967

180. de Berardinis E, Tieri O, Polzella A, et al: The chemical composition of the human aqueous humor in normal and pathological conditions. Exp Eye Res 4:179, 1965

181. DiMatteo J: Active transport of ascorbic acid into lens epithelium of the rat. Exp Eye Res 49:873, 1989

182. Reddy VN, Giblin FJ, Lin LR, et al: The effect of aqueous humor ascorbate on ultraviolet-B–induced DNA damage in lens epithelium. Invest Ophthalmol Vis Sci 39:344, 1998

183. Ringvold A: Aqueous humor and ultraviolet radiation. Acta Ophthalmol 58:69, 1980

184. Ringvold A: The significance of ascorbate in the aqueous humour protection against UV-A and UV-B. Exp Eye Res 62:261, 1996

185. Koskela TK, Reiss GR, Brubaker RF, et al: Is the high concentration of ascorbic acid in the eye an adaptation to intense solar radiation? Invest Ophthalmol Vis Sci 31:2265, 1989

186. Rose RC, Richer SP, Bode AM: Ocular oxidants and antioxidant protection. Proc Soc Exp Biol Med. 217:397, 1998

187. Spector A, Garner WH: Hydrogen peroxide and human cataract. Exp Eye Res 33:673, 1981

188. Kahn MG, Giblin FJ, Epstein DL: Glutathione in calf trabecular meshwork and its relation to aqueous humor outflow facility. Invest Ophthalmol Vis Sci 24:1283, 1983

189. Zhou L, Li Y, Yue BYJT: Oxidative stress affects cytoskeletal structure and cell-matrix interactions in cells from a ocular tissue: The trabecular meshwork. J Cell Physiol 180:182, 1999

190. Alvarado J, Murphy C, Polansky J, et al: Age-related changes in trabecular meshwork cellularity. Invest Ophthalmol Vis Sci 21:714, 1981

191. Grierson I, Wang Q, McMenanin PG, et al: The effects of age and antiglaucoma drugs on the meshwork cell population. Res Clin Forums. 4:69, 1982

192. Grierson I, Howes RC: Age-related depletion of the cell population in thhe human trabecular meshwork. Eye 1:204, 1987

193. Alvarado J, Murphy C, Juster R: Trabecular meshwork cellularity in primary open-angle glaucoma and nonglaucomatous normals. Ophthalmology 91:564, 1984

194. Riley MV: Intraocular dynamics of lactic acid in the rabbit. Invest Ophthalmol Vis Sci 11:600, 1972

195. Sato T, Roy S: Effect of high glucose on fibronectin expression and cell proliferation in trabecular meshwrok cells. Invest Ophthalmol Vis Sci 43:170, 2002

196. Hogg P, Calthorpe M, Batterbury M, et al: Aqueous humor stimulates the migration of human trabecular meshwork cells in vitro. Invest Ophthalmol Vis Sci 41:1091, 2000

197. Stefansson E, Wolbarsht ML, Landers MB: The corneal contact lens and aqueous humor hypoxia in cats. Invest Ophthalmol Vis Sci 24:1052, 1983

198. Stefansson E, Robinson D, Wolbarsht ML: Effects of epinephrine on PO2 in anterior chamber. Arch Ophthalmol 101:636, 1983

199. Heald K, Langham ME: Permeability of the cornea and the blood-aqueous barrier to oxygen. Br J Ophthalmol 40:705, 1956

200. Kleinstein RN, Kwan M, Fatt I, et al: In vivo aqueous humor oxygen tension-as estimated from measurements on bare stroma. Invest Ophthalmol Vis Sci 21:415, 1981

201. Khodadoust AA, Stark WJ, Bill WR: Coagulation properties of intraocular humors and cerebrospinal fluid. Invest Ophthalmol Vis Sci 24:1616, 1983

202. Berzelius JJCJ: The ciliary epithelia and aqueous humor. In Hart WMJ, ed.: Adler's Physiology of the Eye, 9th ed. Philadelphia: CV Mosby, 1992:241

203. Varma SD, Reddy DVN: Phospholipid composition of aqueous humor plasma and lens in normal and alloxan diabetic rabbits. Exp Eye Res 13:120, 1972

204. Kim JO, Cotlier E: Phospholipid distributions and fatty acid composition of lysophosphatidylcholine in rabbit aqueous humor, lens and vitreous. Exp Eye Res 22:569, 1976

205. Obenberger J, Starka L, Hampl R: Quantitative determination of endogenous corticosteroids in the rabbit plasma and aqueous humor. Graefe's Arch Clin Exp Ophthalmol 183:203, 1971

206. Andersson H: Monoamine metabolites in aqueous humor. J Pharmacol 24:998, 1972

207. Farkas TG, Plusec J: The occurrence of trivalent chromium in the aqueous and lens of rats. Invest Ophthalmol 5:398, 1966

208. Ainley RG, Phillips CI, Gibbs A: Aqueous humor vitamin B12 and intramuscular cobalamins. Br J Ophthalmol 53:854, 1969

209. Falbe-Hansen I, Degn JK: Sialic acid in the aqueous humor and the vitreous of normal human eyees and of eyes with malignant melanoma of the choroid. Acta Ophthalmol 47:972, 1969

210. Laurent UBG: Hyaluronate in aqueous humor. Exp Eye Res 33:147, 1981

211. Lütjen-Drecoll E: Functional morphology of the trabecular meshwork in primate eyes. Prog Retin Eye Res 18:91, 1998

212. Knepper PA, Mayanil CSK, Goossens W, et al: Aqueous humor in primary open-angle glaucoma contains an increased level of CD44S. Invest Ophthalmol Vis Sci 43:133, 2002

213. Rauz S, Walker EA, Shackleton CHL, et al: Expression and putative role of 11β-hydroxysteroid dehydrogenase isozymes within the human eye. Invest Ophthalmol Vis Sci 42:2037, 2001

214. Rauz S, Cheung CMG, Wood PJ, et al: Inhibition of 11β-hydroxysteroid dehydrogenase type 1 lowers intraocular pressure in patients with ocular hypertension. Q J Med 96:481, 2003

215. Stokes J, Noble J, Brett L, et al: Distribution of glucocorticoid and mineralocorticoid receptors and 11β-hydroxysteroid dehydrogenases in human and rat ocular tissues. Invest Ophthalmol Vis Sci 41:1629, 2000

216. Whorwood CB, Ricketts ML, Stewart PM: Regulation of sodium-potassium adenosine triphosphate subunit gene expression by corticosteroids and 11β-hydroxysteroid dehydrogenase. Endocrinology 135:901, 1994

217. Ewart HS, Klip A: Hormonal regulation of Na+K+ATPase: Mechanisms underlying rapid and sustained changes in pump activity. Am J Physiol 38:C295, 1995

218. Walker EA, Stewart PM: 11β-Hydroxysteroid dehydrogenase: Unexpected connections. Trends Endocrinol Metab 14:334, 2003

219. Cousins SW, McCabe MM, Danielpour D, et al: Identification of transforming growth factor-beta as an immunosuppressive factor in aqueous humor. Invest Ophthalmol Vis Sci 32:2201, 1991

220. Granstein RD, Staszewski R, Knisely TL, et al: Aqueous humor contains transforming growth factor b and a small (<3500 dalton) inhibitor of thymocyte proliferation. J Immunol 144:3021, 1990

221. Jampel HD, Roche N, Stark WJ, et al: Transforming growth factor-beta in human aqueous humor. Curr Eye Res 9:963, 1990

222. Tripathi RC, Li J, Chan WF, et al: Aqueous humor in glaucomatous eyes contains an increased level of TGF-beta 2. Exp Eye Res 59:723, 1994

223. Wilbanks GA, Mammolenti M, Streilein JW: Studies on the induction of anterior chamber-associated immune deviation (ACAID). III. Induction of ACAID depends upon intraocular transforming growth factor. Eur J Immunol 22:165, 1992

224. Green H, Mann MJ, Paul SD: Elaboration of bicarbonate in intraocular fluids: V. Contribution of the lens to chemistry of aqueous humor of the rabbit eye. Arch Ophthalmol 59:590, 1958

225. Bryson JM, Wolter JR, O'Keefe NT: Ganglion cells in the human ciliary body. Arch Ophthalmol 75:57, 1966

226. Becker B: Iodide transport by the rabbit eye. Am J Physiol 200:804, 1961

227. Bito LZ: Species differences in the response of the eye to irritation and trauma: A hypothesis of divergence in ocular defense mechanisms, and the choice of experimental animals for eye research. Exp Eye Res 39:807, 1984

228. Bito LZ: Prostaglandins: Old concepts and new perspectives. Arch Ophthalmol 105:1036, 1987

229. Bárány EH: Inhibition by hippurate and probenecid of in vitro uptake of iodipamide and o-iodohippurate-composite uptake system for iodipamide in choroid plexus, kidney cortex, and anterior uvea of several species. Acta Physiol Scand 86:12, 1972

230. Bárány EH: In vitro uptake of bile acids by choroid plexus, kidney cortex, and anterior uvea: I. The iodipamide sensitive transprot systems in the rabbit. Acta Physiol Scand 93:250, 1975

231. Bárány EH: The liver-like anion transport system in rabbit kidney, uvea, and choroid plexus: I. Selectivity of some inhibitors, dirction of transport, possible physiological substrates. Acta Physiol Scand 88:412, 1973

232. Bárány EH: Organic cation uptake in vitro by the rabbit iris-ciliary body, renal cortex, and choroid plexus. Invest Ophthalmol 15:341, 1976

233. Bárány EH: Elimination of organic cations from the eye. Klinische Monatsblatter fur Augenheilkunde 184:290, 1984

234. Goldmann H: Abflussdruck, minutemvolumen und Wider-stand der Kammerwasser-Stromung des Menschen. Doc Ophthalmol 5-6:278, 1951

235. Langham ME, Taylor CB: The influence of superior cervical ganglionectomy on intraocular dynamics. J Physiol (Lond) 152:447, 1960

236. Langham ME, Rosenthal AR: The role of cervical sympathetic nerve in regulation of the intraocular pressure and circulation. Am J Physiol 210:786, 1966

237. Bill A, Bárány EH: Gross facility, facility of conventional routes, and pseudofacility of aqueous humor outflow in the cynomolgus monkey. Arch Ophthalmol 75:665, 1966

238. Bill A: Effects of longstanding stepwise increments in eye pressure on the rate of aqueous humor formation in a primate (Cercopithecus ethiops). Exp Eye Res 12:184, 1971

239. Brubaker RF, Riley FCJ: The filtration coefficient of the blood-aqueous barrier. Invest Ophthalmol 11:752, 1972

240. Brubaker RF, Worthen DM: The filtration coefficient of the intraocular vasculature as measured by low-pressure perfusion in a primate eye. Invest Ophthalmol 12:321, 1973

241. Kaufman PL: Aqueous humor dynamics following total iridectomy in the cynomolgus monkey. Invest Ophthalmol Vis Sci 18:870, 1979

242. Bito LZ: Accumulation and apparent active transport of prostaglandins by some rabbit tissues in vitro. J Physiol 221:371, 1972

243. Grant WM: Tonographic method for measuring the facility and rate of aqueous flow in human eyes. Arch Ophthalmol 44:204, 1950

244. Holm O, Krakau CET: Measurements of the flow of aqueous humor according to a new principle. Experientia 22:773, 1966

245. Holm O, Krakau CET: A method of measuring pupillary aqueous flow. Acta Ophthalmol 46:558, 1968

246. O'Rourke J, Macri FJ: Studies in uveal physiology: II. Clinical studies of the anterior chamber clearance of isotopic tracers. Arch Ophthalmol 84:415, 1970

247. Nagataki S: Aqueous humor dynamics of human eyes as studied using fluorescein. Jpn J Ophthalmol 19:235, 1975

248. Jones RF, Maurice DM: New methods of measuring the rate of aqueous flow in man with fluorescein. Exp Eye Res 5:208, 1966

249. Araie M, Sawa M, Nagataki S, et al: Aqueous humor dynamics in man as studied by oral fluorescein. Jpn J Ophthalmol 24:346, 1980

250. Bárány EH, Kinsey VE: The rate of flow of aqueous humor: I. The rate of disappearance of para-aminohippuric acid, radioactive rayopake, and radioactive diodrast from the aqueous humor of rabbits. Am J Ophthalmol 32:177, 1949

251. Linnér E, Friedenwald JS: The appearance time of fluorescein as an index of aqueous flow. Am J Ophthalmol 44:225, 1957

252. Starr PA: Changes in aqueous flow determined by fluorophtometry. Trans Ophthalmol Soc UK 86:639, 1966

253. Bloom JN, Levene RZ, Thomas G, et al: Fluorophotometry and the rate of aqueous flow in man. I. Instrumentation and normal values. Arch Ophthalmol 94:435, 1976

254. Langley D, MacDonald RK: Clinical method for observing changes in the rate of flow of aqueous humor in the human eye: I. Normal eyes. Br J Ophthalmol 36:432, 1952

255. McLaren JW, Brubaker RF: A scanning ocular spectrofluorophotometer. Invest Ophthalmol Vis Sci 29:1285, 1988

256. Langham ME, Taylor CB: Fluorophotometric apparatus for the objective determination of fluorescence in teh anterior chamber of the living eye. Br J Ophthalmol 38:52, 1954

257. Klein R, Ernest JT, Engerman RL: Fluorophotometry. I. Technique. Arch Ophthalmol 98:2231, 1980

258. Kinsey VE, Bárány EH: The rate of flow of aqueous humor: II. Derivation of rate of flow and its physiological significance. Am J Ophthalmol 32:189, 1949

259. Goldmann H: Uber Fluorescein in der menschlichen Vord-erkammer. Das Kammerwasser-Minutenvolumen des Menschen. Ophthalmologica 119:65, 1950

260. Hayashi M, Yablonski ME, Mindel JS: Methods for assisting the effects of pharmacologic agents on aqueous humor dynamics. In Tasman W, Jaeger EA, eds.: Biomedical Foundations of Ophthalmology: Philadelphia: JB Lippincott, 1990:1–9

261. McLaren JW, Trocme SD, Relf S, et al: Rate of flow aqueous humor determined from measurements of aqueous flare. Invest Ophthalmol Vis Sci 31:339, 1990

262. Maurice DM: Theory and methodology of vitreous fluorophotometry. Jpn J Ophthalmol 29:119, 1985

263. Gaul GR, Brubaker RF: Measurement of aqueous flow in rabits with corneal and vitreous depots of fluorescent dye. Invest Ophthalmol Vis Sci 27:1331, 1986

264. Becker B: The measurement of rate of aqueous flow with iodide. Invest Ophthalmol 1:52, 1962

265. Nagataki S, Brubaker RF: The effect of pilocarpine on aqueous humor formation in human beings. Arch Ophthalmol 100:818, 1982

266. Brubaker RF: Measurement with fluorophotometry: I. Plasma binding. II. Anterior segment, and III. Aqueous humor flow. Graefes Arch Clin Exp Ophthalmol 222:190, 1985

267. Bárány EH: Simultaneous measurements of changing intraocular pressure and outflow facility in the vervet monkey by constant pressure infusion. Invest Ophthalmol 3:135, 1964

268. Bill A: Conventional and uveoscleral drainage of aqueous humor in the cynomolgus monkey (Macaca irus) at normal and high intraocular pressu Res Exp Eye Res 5:45, 1966

269. Gabelt BT, Kaufman PL: The effect of prostaglandin F2a on trabecular outflow facility in cynomolgus monkeys. Exp Eye Res 51:87, 1990

270. Bill A: Circulation in the eye. In Renkin EM, Michel CC, eds.: The Cardiovascular System IV. Washington, DC: American Physiological Society, 1984:1001

271. Galin MA, Baras I, Mandell GKL: Measurement of aqueous flow utilizing the perilimbal suction cup. Arch Ophthalmol 66:65, 1961

272. Maus TL, McLaren JW, Shepard JWJ, et al: The effects of sleep on circulating catecholamines and aqueous flow in human subjects. Exp Eye Res 62:351, 1996

273. Wetzel RK, Eldred WD: Specialized neuropeptide Y- and glucagon-like immunoractive amacrine cells in the peripheral retina of the turtle. Vis NeuroSci 14:867, 1997

274. Wilson WS, Shahidullah M, Millar C: The bovine arterially-perfused eye: An in vitro method for the study of drug mechanisms on IOP, aqueous humour formation and uveal vasculature. Curr Eye Res 12:609, 1993

275. Shahidullah M, Wilson WS, Millar C: Effects of timolol, terbutaline and forskolin on IOP, aqueous humour formation and ciliary cyclic AMP levels in the bovine eye. Curr Eye Res 14:519, 1995

276. Bill A: Early effects of epinephrine on aqueous humor dynamics in vervet monkeys (Cercopithecus ethiops). Exp Eye Res 8:35, 1969

277. Bill A: Effects of norepinephrine, isoproterenol and sympathetic stimulation on aqueous humour dynamics in vervet monkeys. Exp Eye Res 10:31, 1970

278. Townsend DJ, Brubaker RF: Immediate effect of epinephrine on aqueous formation in the normal human eye as measured by fluorophotometry. Invest Ophthalmol Vis Sci 19:256, 1980

279. Topper JE, Brubaker RF: Effects of timolol, epinephrine, and acetazolamide on aqueous flow during sleep. Invest Ophthalmol Vis Sci 26:1315, 1985

280. Vareilles P, Silverstone D, Plazonnet B, et al: Comparison of the effects of timolol and other adrenergic agents on intraocular pressure in the rabbit. Invest Ophthalmol 16:987, 1977

281. Gregory D, Sears M, Bausher L: Intraocular pressure and aqueous flow are decreased by cholera toxin. Invest Ophthalmol Vis Sci 20:371, 1981

282. Schmitt CJ, Lotti VJ, Varailles P, et al: β-Adrenergic blockers: Lack of relationship between antagonism of isoproterenol and lowering of intraocular pressure in rabbits. In Sears M, ed.: New Directions in Ophthalmic Research. New Haven: Yale University Press, 1981:147–162

283. Caprioli J, Sears ML: Forskolin lowers intraocular pressure in rabbits, monkeys, and man. Lancet 1:958, 1983

284. Smith BR, Gaster RN, Leopold IH, et al: Forskolin, a potent adenylate cyclase activator, lowers rabbit intraocular pressure. Arch Ophthalmol 102:146, 1984

285. Bartels SP, Lee SR, Neufeld AH: The effects of forskolin on cyclic AMP, intraocular pressure and aqueous humor formation in rabbits. Curr Eye Res 6:307, 1987

286. Shibata T, Mishima H, Kurokawa T: Ocular pigmentation and intraocular pressure response to forskolin. Curr Eye Res 7:667, 1988

287. Elena PP, Fredj-Reygrobellet D, Moulin G, et al: Pharmacological characteristics of β-adrenergic adenylate cyclase in nonpigmented and in pigmented cells of bovine ciliary processes. Curr Eye Res 3:1383, 1984

288. Nathanson JA: Adrenergic regulation of intraocular pressure: Identification of beta 2-adrenergic-stimulated adenylate cyclase in the ciliary process epithelium. Proc Natl Acad Sci USA 77:7420, 1980

289. Nathanson JA: Effects of a potent and specific β2-adrenoceptor antagonist on intraocular pressure. Br J Ophthalmol 73:97, 1981

290. Nathanson JA: Differential inhibition of β-adrenergic receptors in human and rabbit ciliary processes and heart. J Pharmacol Exp Ther 232:119, 1985

291. Nathanson JA: Biochemical and physiological effects of S-32-468, a new β-adrenoceptor antagonist with possible oculoselectivity. Curr Eye Res 4:191, 1985

292. Nathanson JA: Atriopeptin-activated guanylate cyclase in the anterior segment. Identification, localization and effects of atriopeptins on IOP. Invest Ophthalmol Vis Sci 28:1357, 1987

293. Nathanson JA: Stereospecificity of beta adrenergic antagonists: R-enantiomers show increased selectivity for beta-2 receptors in ciliary process. J Pharmacol Exp Ther 245:94, 1988

294. Nathanson JA: Direct application of a guanylate cyclase activator lowers intraocular pressure. Eur J Pharmacol 147:155, 1988

295. Wax MB, Molinoff PB: Distribution and properties of β-adrenergic receptors in human iris-ciliary body. Invest Ophthalmol Vis Sci 28:420, 1987

296. Millar C, Wilson WS: Comparison of the effects of vasodilator drugs on intraocular pressure and vascular relaxation. Br J Pharmacol 104 (suppl):55, 1991

297. Burke JA, Potter DE: Ocular effects of a relatively selective α2-agonist (UK14304-18) in cats, rabbits, and monkeys. Curr Eye Res 5:665, 1986

298. Serle JB, Steidl S, Wang R-F, et al: Selective α2-adrenergic agonists B-HT 920 and UK14304-18. Effects on aqueous humor dynamics in monkeys. Arch Ophthalmol 109:1158, 1991

299. Gabelt BT, Robinson JC, Hubbard WC, et al: Apraclonidine and brimonidine effects of anterior ocular and cardiovascular physiology in normal and sympathectomized monkeys. Exp Eye Res 59:633, 1994

300. Gharagozloo NZ, Relf SJ, Brubaker RF: Aqueous flow is reduced by the alpha-adrenergic agonist, apraclonidine hydrochloride (ALO 2145). Ophthalmology 95:1217, 1988

301. Burke J, Kharlamb A, Shan T, et al: Adrenergic and imidazoline receptor-mediated responses to UK-14,304-18 (brimonidine) in rabbits and monkeys. A species difference. Ann NY Acad Sci 763:78, 1995

302. Allen RC, Langham ME: The intraocular pressure response of conscious rabbits to clonidine. Invest Ophthalmol 15:815, 1976

303. Harvey SC: Hypnotics and sedatives. In Gilman AG, Goodman LS, Rall TW, Murad R, eds.: , eds.: The Pharmacological Basis of Therapeutics, 7th ed. New York: Macmillan, 1985:339–371

304. Marshall BE, Longnecker DE: General anesthetics. In Gilman AG, Rall TW, Nies AS, Taylor P, eds.: The Pharmacological Basis of Therapeutics, 8th ed. New York: McGraw-Hill, Inc., 1993:285–310

305. Robinson JC, Kaufman PL: Dose-dependent suppression of aqueous humor formation by timolol in the cynomolgus monkey. J Glaucoma. 2:251, 1993

306. Nilsson SFE, Maepea O, Samuelsson M, et al: Effects of timolol on terbutaline- and VIP-stimulated aqueous humor flow in the cynomolgus monkey. Curr Eye Res 9:863, 1990

307. Gartner S: Blood vessels of the conjunctiva. Arch Ophthalmol 32:464, 1944

308. Wilke K: Early effects of epinephrine and pilocarpine on the intraocular pressure and the episcleral venous pressure in the normal human eye. Acta Ophthalmol 52:231, 1974

309. Alm A, Bill A, Young FA: The effects of pilocarpine and neostigmine on the blood flow through the anterior uvea in monkeys. A study with radioactively labelled microspheres. Res Exp Eye Res 15:31, 1973

310. James RG, Calkins JP: Effect of certain drugs on iris vessels. Arch Ophthalmol 57:414, 1957

311. Kolker AE, Hetherington JJ: Becker-Shaffer's Diagnosis and Therapy of the Glaucomas, 4th ed. St. Louis: CV Mosby, 1976:325–334

312. Swan K, Hart W: A comparative study of the effects of mecholyl, doryl, pilocarpine, atropine, and epinephrine on the blood-aqueous barrier. Am J Ophthalmol 23:1311, 1940

313. Bito LZ, Davson H, Snider N: The effects of autonomic drugs on mitosis and DNA synthesis in the lens epithelium and on the composition of the aqueous humor. Exp Eye Res 4:54, 1965

314. Wålinder P-E: Influence of pilocarpine on iodopyracet and iodide accumulation by rabbit ciliary body-iris preparations. Invest Ophthalmol 5:378, 1966

315. Wålinder P-E, Bill A: Influence of intraocular pressure and some drugs on aqueous flow and entry of cycloleucine into the aqueous humor of vervet monkeys (Cercopithecus ethiops). Invest Ophthalmol 8:446, 1969

316. Bárány EH: A mathematical formulation of intraocular pressure as dependent on secretion, ultrafiltration, bulk outflow, and osmotic reabsorption of fluid. Invest Ophthalmol 2:584, 1963

317. Gaasterland D, Kupfer C, Ross K: Studies of aqueous humor dynamics in man: IV. Effects of pilocarpine upon measurements in young normal volunteers. Invest Ophthalmol 14:848, 1975

318. Kupfer C: Clinical significance of pseudofacility. Sanford R. Gifford Memorial Lecture. Am J Ophthalmol 75:193, 1973

319. Wålinder P-E, Bill A: Aqueous flow and entry of cycloleucine into the aqueous humor of vervet monkeys (Cercopithecus ethiops). Invest Ophthalmol 8:434, 1969

320. Macri FJ, Cevario SJ: The induction of aqueous humor formation by the use of Ach + eserine. Invest Ophthalmol 12:910, 1973

321. Macri FJ, Cevario SJ: The dual nature of pilocarpine to stimulate or inhibit the formation of aqueous humor. Invest Ophthalmol 13:617, 1974

322. Bill A: Effects of atropine and pilocarpine on aqueous humour dynamics in cynomolgus monkeys (macaca irus). Exp Eye Res 6:120, 1967

323. Bill A, Walinder P-E: The effects of pilocarpine on the dynamics of aqueous humor in a primate (macaca irus). Invest Ophthalmol 5:170, 1966

324. Miichi H, Nagataki S: Effects of pilocarpine, salbutamol, and timolol on aqueous humor formation in cynomolgus monkeys. Invest Ophthalmol Vis Sci 24:1269, 1983

325. Polansky JR, Wood IS, Maglio MT, et al: Trabecular meshwork cell culture in glaucoma research: Evaluation of biological activity and structural properties of human trabecular cells in vitro. Ophthalmology 91:580, 1984

326. Polansky JR, Bloom E, Konami D, et al: Cultured human trabecular cells: Evaluation of hormonal and pharmacological responses in vitro. In Ticho U, David R, eds.: Recent Advances in Glaucoma. Amsterdam: Excerpta Medica, 1984:201–206

327. Carlson KH, Bourne WM, McLaren JW, et al: Variations in human corneal endothelial cell morphology and permeability to fluorescein with age. Exp Eye Res 47:27, 1988

328. Lütjen-Drecoll E: Functional morphology of the ciliary epithelium. In Lütjen-Drecoll E, ed.: Basic Aspects of Glaucoma Research. Stuttgart: Schattauer, 1982:69–87

329. Bárány EH: Pseudofacility and uveoscleral outflow routes: Some nontechnical difficulties in the determination of outflow facility rate and rate of formation of aqueous humor. In Leydhecker W, ed.: Glaucoma Symposium. Tutzing Castle. Basel: Karger, 1966:27–51

330. Bill A: Further studies on the influence of the intraocular presssure on aqueous humor dynamics in cynomolgus monkeys. Invest Ophthalmol 6:364, 1967

331. Goldmann H: On pseudofacility. Bibl Ophthalmol 76:1, 1968

332. Moses RA: Intraocular pressure. In Moses RA, ed.: Adler's Physiology of the Eye: Clinical Application. St Louis: CV Mosby, 1981:227

333. Kaufman PL, Wiedman T, Robinson JR: Cholinergics. In Sears ML, ed.: Handbook of Experimental Pharmacology. Berlin: Springer-Verlag, 1984:149–191

334. Bill A: The effect of ocular hypertension caused by red cells on the rate of formation of aqueous humor. Invest Ophthalmol 7:162, 1968

335. Mäepea O, Nilsson SF: Suppression of VIP- and terbutaline stimulated aqueous humor flow by increased intraocular pressure in the cynomolgus monkey. Curr Eye Res 10:703, 1991

336. Householder JR, Clausen DF, Harris JE: The IOP sensitivity of the anterior chamber appearance time of intraarterially injected fluorescein. Exp Eye Res 4:83, 1965

337. Hoskins HD, Jr, Kass MA: Medical treatment. In Hoskins HD, Jr, Kass MA, eds.: Becker-Shaffer's Diagnosis and Therapy of the Glaucomas, 6th ed. ed. St. Louis: CV Mosby, 1989:405–493

338. Bill A: Aspects on suppressability of aqueous humor formation. Doc Ophthalmol 26:73, 1969

339. Kaufman PL, Bill A, Bárány EH: Formation and drainage of aqueous humor following total iris removal and ciliary muscle disinsertion in the cynomolgus monkey. Invest Ophthalmol Vis Sci 16:226, 1977

340. Toris CB, Pederson JE: Aqueous humor dynamics in experimental iridocyclitis. Invest Ophthalmol Vis Sci 28:477, 1987

341. Toris CB, Gregerson DS, Pederson JE: Uveoscleral outflow using different-sized fluorescent tracers in normal and inflamed eyes. Exp Eye Res 45:525, 1987

342. Cole DF: Reduction in aqueous humor formation as caused by iodate, spironolactone and polyphloretin phosphate. Br J Ophthalmol 46:291, 1962

343. Maus TL, Young WFJ, Brubaker RF: Aqueous flow in humans after adrenalectomy. Invest Ophthalmol Vis Sci 35:3325, 1994

344. Gharagozloo NZ, Brubaker RF: The correlation between serum progesterone and aqeuous dynamics during the menstrual cycle. Acta Ophthalmol 69:791, 1991

345. Horven I, Lilleaasen P, Aasen A: Intraocular pressure before, during, and after extracorporeal circulation in pigs. Scand J Thorac Cardiovasc Surg 15:269, 1981

346. van Loenen AC, Van Bijsterveld OP, Nijkamp F: Some aspects of waterloading in rabbits. Doc Ophthalmol 56:345, 1984

347. Kolker AE: Hyperosmotic agents in glaucoma. Invest Ophthalmol Vis Sci 9:418, 1970

348. Lorimer DW, Hakanson NE, Pion PD, et al: The effect of intravenous mannitol or oral glycerol on intraocular pressure in dogs. Cornell Vet 79:249, 1989

349. Gafter U, Pinkas M, Hirsch J: Intraocular pressure in uremic patients on chronic hemodialysis. Nephron 40:74, 1985

350. Zimmerman TJ: Topical ophthalmic beta blockers. A comparative review. J Ocular Pharmacol 9:373, 1993

351. Coakes RL, Brubaker RF: The mechanism of timolol in lowering intraocular pressure. Arch Ophthalmol 96:2045, 1978

352. Yablonski ME, Zimmerman TJ, Waltman SR: A fluorophotometric study of the effect of topical timolol on aqueous humor dynamics. Exp Eye Res 27:135, 1978

353. Schenker JI, Yablonski ME, Podos SM, et al: Fluorophotometric study of epinephine and timolol in human subjects. Arch Ophthalmol 99:1212, 1981

354. Stewart RH, Kimbrough RL, Ward RL: Betaxolol vs. timolol. A six-month double-blind comparison. Arch Ophthalmol 104:46, 1986

355. Araie M, Takase M: Effects of S-596 and carteolol, new β-adrenergic blockers, and flurbiprofen on the human eye: A fluorometric study. Graefes Arch Clin Exp Ophthalmol 222:259, 1985

356. Yablonski ME, Novack GD, Burke PJ: The effect of levobunolol on aqueous humor dynamics. Exp Eye Res 44:49, 1987

357. Mills KB, Wright G: A blind randomized cross-over trial comparing metipranolol 0.3% with timolol 0.25% in open-angle glaucoma: A pilot study. Br J Ophthalmol 70:39, 1986

358. Main BG, Tucker H: Recent advances in beta-adrenergic blocking agents. In Ellis GP, West GB, eds.: Progress in Medicinal Chemistry. Amsterdam: Elsevier, 1985:121–164

359. Novack GD: Ophthalmic β-blockers since timolol. Surv Ophthalmol 31:307, 1987

360. Bonomi L, Steindler P: Effect of pindolol on intraocular pressure. Br J Ophthalmol 59:301, 1975

361. Murray DL, Podos SM, Wei C, et al: Ocular effects in normal rabbits of topically applied labetalol, a combined α- and β-adrenergic antagonist. Arch Ophthalmol 97:723, 1979

362. Bouzoubaa M, Leclerc G, Decker N, et al: Synthesis and β-adrenergic blocking activity of new aliphatic and alicyclic oxime ethers. J Med Chem 27:1291, 1984

363. Sugrue MF, Gautheron P, Grove J: MK-927: A topically active ocular hypotensive carbonic anhydrase inhibitor. J Ocul Pharmacol 6:9, 1990

364. Becker B: Decrease in intraocular pressure in man by a carbonic anhydrase inhibitor, Diamox. Am J Ophthalmol 37:13, 1954

365. Hoskins HD Jr, Kass MA: Aqueous humor formation. In Klein EA, ed.: Becker-Shaffers's Diagnosis and Therapy of the Glaucomas, 6th ed. St. Louis: CV Mosby, 1989:18–40

366. Woltersdorf OWJ, Schwam H, Bicking JB: Topically active carbonic anhydrase inhibitors. 1. O-acyl derivatives of 6-hydroxybenzothiazole-2-sulfonamide. J Med Chem 32:2486, 1989

367. Sugrue MF, Viader MP: Synthetic atrial natriuretic factor lowers rabbit intraocular pressure. Eur J Pharmacol 130:349, 1986

368. Millar JC, Carr RD, Humphries RG, et al: Drug effects on intraocular pressure and vascular flow in the bovine perfused eye using radiolabelled microsphere. Res J Ocular Pharmacol Ther 11:11, 1995

369. Mittag TW, Tormay A, Ortega M, et al: Atrial natriuretic peptide (ANP), guanylate cyclase, and intraocular pressure in the rabbit eye. Curr Eye Res 6:1189, 1987

370. Korenfeld MS, Becker B: Atrial natriuretic peptides: Effects on intraocular pressure, cGMP, and aqueous flow. Invest Ophthalmol Vis Sci 30:2385, 1989

371. Karnezis TA, Murphy MB: Dopamine receptors and intraocular pressure. Trends Pharmacol Sci 9:389, 1988

372. Krupin T, Weiss A, Becker B, et al: Increased intraocular pressure following topical azide or nitroprusside. Invest Ophthalmol Vis Sci 16:1002, 1977

373. Wiederholt M, Sturm A, Lepple-Wienhues A: Relaxation of trabecular meshwork and ciliary muscle by release of nitric oxide. Invest Ophthalmol Vis Sci 35:2515, 1994

374. Nathanson JA, McKee M: Identification of an extensive system of nitric oxide-producing cells in the ciliary muscle and outflow pathway of the human eye. Invest Ophthalmol Vis Sci 36:1765, 1995

375. Rasmussen CA, Gabelt BT, Kaufman PL: Effect of nitric oxide compounds on ciliary muscle in vitro [Abstract 5026]. Invest Ophthalmol Vis Sci 45:S5026, 2004

376. Kee C, Kaufman PL, Gabelt BT: Effect of 8-Br cGMP on aqueous humor dynamics in monkeys. Invest Ophthalmol Vis Sci 35:2769, 1994

377. Becker B: Topical 8-bromo-cyclic BMP lowers intraocular pressure in rabbits. Invest Ophthalmol Vis Sci 31:1647, 1990

378. Payne LJ, Slagle TM, Cheeks LT, et al: Effect of calcium channel blockers on intraocular pressure. Ophthalmic Res 22:337, 1990

379. Beatty JF, Krupin T, Nichols PF, et al: Elevation of intraocular pressure by calcium channel blockers. Arch Ophthalmol 102:1072, 1984

380. Chang FW, Burke JA, Potter DE: Mechanism of the ocular hypotensive action of ketanserin. J Ocular Pharmacol 1:137, 1985

381. Krootila K, Palkama A, Uusitalo H: Effect of serotonin and its antagonist (ketanserin) on intraocular pressure in the rabbit. J Ocular Pharmacol 3:279, 1987

382. Barnett NL, Osborne NN: The presence of serotonin (5-HT1) receptors negatively coupled to adenylate cyclase in rabbit and human ciliary processes. Exp Eye Res 57:209, 1993

383. Meyer-Bothling U, Bron AJ, Osborne NN: Topical application of serotonin or the 5-HT1a-agonist 5-CT increases intraocular pressure in rabbits. Invest Ophthalmol Vis Sci 34:3035, 1993

384. Chidlow G, Cupido A, Melena J, et al: Flesinoxan, a 5-HT1A receptor agonist/alpha 1-adrenoceptor antagonist, lowers intraocular pressure in NZW rabbits. Curr Eye Res 23:144, 2001

385. Gabelt BT, Millar CJ, Kiland JA, et al: Effects of serotonergic compounds on aqueous humor dynamics in monkeys. Curr Eye Res 23:120, 2001

386. May JC, McLaughlin MA, Sharif NA, et al: Evaluation of the ocular hypotensive response of serotonin 5-HT1A and 5-HT-2 receptor ligands in conscious ocular hypertensive cynomolgus monkeys. J Pharmacol Exp Ther 306:301, 2003

387. May JA, Chen HH, Rusinko A, et al: A novel and selective 5-HT2 receptor agonist with ocular hypotensive activity: (S)-(+)-1-(2-aminopropyl)-8,9-dihydropyrano[3,2-e]indole. J Med Chem 46:4188, 2003

388. Mastropasqua L, Costagliola C, Ciancaglinin M, et al: Ocular hypotensive effect of ketanserin in patients with primary open angle glaucoma. Acta Ophthalmol Scand (Suppl) 224:24, 1997

389. Takenaka H, Mano T, Maeno T, et al: The effect of anplag (sarpogrelate HCl), novel selective 5-HT2 antagonist on intraocular pressure in glaucoma patients [Abstract 3390]. Invest Ophthalmol Vis Sci 36:S734, 1995

390. Okka M, Gabelt BT, Dean TR, et al: Aqueous humor dynamics in monkeys after topical R(+)-DOI. Invest Ophthalmol Vis Sci 45:5034, 2004

391. Crawford KS, Kaufman PL: Dose-related effects of prostaglandin F2a isopropylester on intraocular pressure, refraction and pupil diameter in monkeys. Invest Ophthalmol Vis Sci 32:510, 1991

392. Sharif NA, Kelly CR, Crider JY, et al: Human ciliary muscle and trabecular meshwork cells express functional serotonin-2 (5HT2) receptors coupled to phosphoinositide turnover and intracellular Ca2+ mobilization [Abstract 2084]. Invest Ophthalmol Vis Sci 44:S2084, 2003

393. Mallorga P, Sugrue MF: Characterization of serotonin receptors in the iris+ciliary body of the albino rabbit. Curr Eye Res 6:527, 1987

394. De Vries GW, Mobasser A, Wheeler LA: Stimulation of endogenous cyclic AMP levels in ciliary body by SK&F28526, a novel dopamine receptor agonist. Curr Eye Res 5:449, 1986

395. Chu E, Chu TC, Potter DE: Potential sites of action of TNPA: A dopamine-2 agonist. Exp Eye Res 69:611, 1999

396. Chu E, Chu TC, Potter DE: Mechanisms and sites of ocular action of 7-hydroxy-2-dipropylaminotetralin: A dopamine(3) receptor agonist. J Pharmacol Exp Ther 293:710, 2000

397. Leopold IH, Duzman E: Observations on the pharmacology of glaucoma. Ann Rev Pharmacol Toxicol 26:401, 1986

398. Chu E, Socci R, Chu TC: PD128,907 induces ocular hypotension in rabbits: Involvement of D2/D3 dopamine receptors and brain natriuretic peptide. J Ocul Pharmacol Ther 20:15, 2004

399. Potter DE, Burke JA: Effects of ergoline derivatives on intraocular pressure and iris function in rabbits and monkeys. Curr Eye Res 2:281, 1982

400. Potter DE, Burke JA, Chang FW: Ocular hypotensive action of ergoline derivatives in rabbits: Effects of sympathectomy and domperidone pretreatment. Curr Eye Res 3:307, 1984

401. Langer SZ: Presynaptic regulation of the release of catecholamines. Phamacol Rev 32:337, 1980

402. Siegel MJ, Lee PY, Podos SM, et al: Effect of topical pergolide on aqueous dynamics in noral and glaucomatous monkeys. Exp Eye Res 44:227, 1987

403. Mekki QA, Hassan SM, Turner P: Bromocriptine lowers intraocular pressure without affecting blood pressure. Lancet 1:1250, 1983

404. Mekki QA, Warrington SJ, Turner P: Bromocriptine eye-drops lower intraocular pressure without affecting prolactin levels. Lancet 1:287, 1984

405. Geyer O, Robinson D, Lazar M: Hypotensive effect of bromocriptine in normal eyes. J Ocul Pharmacol 3:291, 1987

406. Chiou GCY: Treatment of ocular hypertension and glaucoma with dopamine antagonists. Ophthalmic Res 16:129, 1984

407. Chiou GCY: Ocular hypotensive actions of haloperidol, a dopaminergic antagonist. Arch Ophthalmol 102:143, 1984

408. Krupin T, Feitl M, Becker B: Effect of prazosin on aqueous humor dynamics in rabbits. Arch Ophthalmol 98:1639, 1980

409. Mittag TW: Ocular effects of selective alpha-adrenergic agents: A new drug paradox? Ann Ophthalmol 15:201, 1983

410. Mittag TW, Tormay A, Severin C, et al: Alpha-adrenergic antagonists: Correlation of the effect on intraocular pressure and on α2-adrenergic receptor binding specificity in the rabbit eye. Exp Eye Res 40:591, 1985

411. Hedler L, Stamm G, Weitzell R, et al: Functional characterization of central α-adrenoceptors by yohimbine diastereoisomers. Eur J Pharmacol 70:43, 1981

412. Serle JB, Stein AJ, Podos SM, et al: Coryanthine and aqueous humor dynamics in rabbits and monkeys. Arch Ophthalmol 102:1385, 1984

413. Jampel HD, Robin AL, Quigley HA, et al: Apraclonidine. A one-week dose-response study. Arch Ophthalmol 106:1069, 1988

414. Robin AL: Short-term effects of unilateral 1% apraclonidine therapy. Arch Ophthalmol 106:912, 1988

415. Abrams DA, Robin AL, Pollack IP, et al: The safety and efficacy of topical 1% ALO 2145 (p-aminoclonidine hydrochloride) in normal volunteers. Arch Ophthalmol 105:1205, 1987

416. Coleman AL, Robin AL, Pollack IP, et al: Cardiovascular and intraocular pressure effects and plasma concentrations of apraclonidine. Arch Ophthalmol 108:1264, 1990

417. Robin AL, Coleman AL: Apraclonidine hydrochloride: An evaluation of plasma concentrations, and a comparison of its intraocular pressure lowering and cardiovascular effects to timolol maleate. Trans Am Ophthalmol Soc 88:159, 1990

418. McCannel C, Koskela T, Brubaker RF: Topical flurbiprofen pretreatment does not block apraclonidine's effect on aqueous flow in humans. Arch Ophthalmol 109:810, 1991

419. Schadlu R, Maus TL, Nau CB, et al: Comparison of the eficacy of apraclonidine and brimonidine as aqueous suppressants in humans. Arch Ophthalmol 116:1441, 1998

420. Barnebey HS, Robin AL, Zimmerman TJ, et al: The efficacy of brimonidine in decreaseing elevations in intaocular pressure after laser trabeculoplasty. Ophthalmology 100:1083, 1993

421. Igic R, Kojovic V: Angiotensin I converting enzyme (kininase II) in ocular tissues. Exp Eye Res 30:299, 1980

422. Vita JB, Anderson JA, Hulem CD, et al: Angiotensin-converting enzyme activity in ocular fluids. Invest Ophthalmol Vis Sci 20:255, 1981

423. Weinreb RN, Sandman R, Ryder MI, et al: Angiotensin-converting enzyme activity in human aqueous humor. Arch Ophthalmol 103:34, 1985

424. Ferrari-Dileo G, Ryan JW, Rockwook EJ, et al: Angiotensin-converting enzyme in bovine, feline, and human ocular tissues. Invest Ophthalmol Vis Sci 29:876, 1988

425. Sramek SJ, Wallow IH, Tewksbury DA, et al: An ocular renin-angiotensin system. Immunohistochemistry of angiotensinogen. Invest Ophthalmol Vis Sci 33:1627, 1992

426. Wallow IH, Sramek SJ, Bindley CD, et al: Ocular renin angiotensin: EM immunocytochemical localization of prorenin. Curr Eye Res 12:945, 1993

427. Sramek SJ, Wallow IHL, Day RP, et al: Ocular renin-angiotensin: Immunohistochemical evidence for the presence of prorenin in eye tissue. Invest Ophthalmol Vis Sci 29:1749, 1988

428. Watkins RW, Baum T, Cedeno K: Topical ocular hypotensive effects of the novel angiotensin converting enzyme inhibitor SCH 33861 in conscious rabbits. J Ocul Pharmacol 3:295, 1987

429. Constad WH, Fiore P, Samson C, et al: Use of an angiotensin converting enzyme inhibitor in ocular hypertension and primary open-angle glaucoma. Am J Ophthalmol 105:674, 1988

430. Krupin T, Siverstein B, Feitl M: The effect of H1-blocking antihistamines on intraocular pressure in rabbits. Ophthalmology. 87:1167, 1980

431. Purnell WD, Gregg JM: Δ9-tetrahydrocannabinol, euphoria and intraocular pressure in man. Ann Ophthalmol 7:921, 1975

432. Green K, Pederson JE: Effect of Δ1 -tetrahydrocannabinol on aqueous dynamics and ciliary body permeability in the rabbit. Exp Eye Res 15:499, 1973

433. Green K, Roth M: Ocular effects of topical administration of Δ9-tetrahydrocannabinol in man. Arch Ophthalmol 100:265, 1982

434. Jay WM, Green K: Multiple-drop study of topically applied 1% Δ9-tetrahydrocannabinol in human eyes. Arch Ophthalmol 101:591, 1983

435. Seeman JL, Gabelt BT, Kiland JA, et al: HU-210 and WIN 55,212,2 effects on IOP, pupil and refraction in cynomolgus monkeys [Abstract 4501]. Invest Ophthalmol Vis Sci 42:S839, 2001

436. Chien FY, Wang RF, Mittag TW, et al: Effect of WIN 55212-2, a cannabinoid receptor agonist, on aqueous humor dynamics in monkeys. Arch Ophthalmol 121:87, 2003

437. Oltmanns MH, Lattanzio FA, Allen RC, et al: Topical WIN 55 212-2 reduces intraocular pressure in ocular hypertensive rats [Abstract 2093]. Invest Ophthalmol Vis Sci 45(suppl):S2093, 2004

438. Laine K, Järvinen K, Mechoulam R, et al: Comparison of the enzymatic stability and intraocular pressure effects of 2-arachidonylglycerol and noladin ether, a novel putative endocannabinoid. Invest Ophthalmol Vis Sci 43:3216, 2002

439. Song A-H, Slowey C-A: Involvement of cannabinoid receptors in the intraocular pressure-lowering effects of WIN55212-2. J Pharmacol Exp Ther 292:136, 2000

440. Porcella A, Chiara M, Gessa GL, et al: The synthetic cannabinoid WIN55212-2 decreases the intraocular pressure in human glaucoma resistant to conventional therapies. Eur J Neurosci 13:409, 2001

441. Stamer WD, Golightly SF, Hosohata Y, et al: Cannabinoid CB(1) receptor expression, activation and detection of endogenous ligand in trabecular meshwork and ciliary process tissues. Eur J Pharmacol 431:277, 2001

442. Straiker AJ, Maguire G, Mackei K, et al: Localization of cannabinoid CB1 receptors in the human anterior eye and retina. Invest Ophthalmol Vis Sci 40:2442, 1999

443. Russell KR, Wang DR, Potter DE: Modulation of ocular hydrodynamics and iris function by bremeazocine, a kappa opioid receptor agonist. Exp Eye Res 70:675, 2000

444. Ruskell KR, Potter DE: Dynorphin modulates ocular hydrodynamics and releases atrial natriuretic peptide via activation of kappa-Opioid receptors. Exp Eye Res 75:259, 2002

445. Russell KR, Moore TT, Potter DE: Elevation of atrial natriuretic peptide levels in aqueous humor of the rabbit by kappa opioid receptor agonists. Neuropeptides 35:232, 2001

446. Potter DE, Russell KR, Manhiani M: Bremazocine increases C-type natriuretic peptide levels in aqueous humor and enhances outflow facility. J Pharmacol Exp Ther 309:548, 2004

447. Tisha T, Potter DE: Kappa opioid agonist-induced changes in IOP: Correlation with 3H-NE release and cAMP accumulation. Exp Eye Res 73:167, 2001

448. Rasmussen CA, Gabelt BT, Russell KR, et al: Blood pressure effects on bremazocine IOP resp outflow facility in cynomolgus monkeys [Abstract 3436]. Invest Ophthalmol Vis Sci 44:S3436, 2003

449. Anderson L, Wilson WS: Inhibition by indomethacin of the increased facility of outflow induced by adrenaline. Exp Eye Res 50:119, 1990

450. Coakes RL, Siah PB: Effects of adrenergic drugs on aqueous humor dynamics in the normal human eye. I. Salbutamol. Br J Ophthalmol 68:393, 1984

451. Larson RS, Brubaker RF: Isoproterenol stimulates aqueous flow in humans with Horner's syndrome. Invest Ophthalmol Vis Sci 29:621, 1988

452. Gharagozloo NZ, Larson RS, Kullerstrand W: Terbutaline stimulates aqueous humor flow in humans during sleep. Arch Ophthalmol 106:1218, 1988

453. Lee DA, Topper JE, Brubaker RF: Effect of clonidine on aqueous humor flow in normal human eyes. Exp Eye Res 38:239, 1984

454. Fechtner RD: Beta blockers. In: Glaucoma Medical Therapy Principles and Management. San Francisco: The Foundation of the American Academy of Ophthalm 1999:25–39

455. Becker B: Does hyposecretion of aqueous humor damage the trabecular meshwork? J Glaucoma. 4:303, 1995

456. Lütjen-Drecoll E, Kaufman PL: Long-term timolol and epinephrine in monkeys. II. Morphological alterations in trabecular meshwork and ciliary muscle. Trans Ophthalmol Soc UK 105:196, 1986

457. Hoskins HD, Kass MA: Aqueous humor outflow. In Hoskins HD, Kass MA, eds.: Becker-Shaffer's Diagnosis and Therapy of the Glaucomas. St. Louis: CV Mosby, 1989:41–66

458. Hoskins HD, Kass MA: Secondary open-angle glaucoma. In Hoskins HD, Kass MA, eds.: Becker-Shaffer's Diagnosis and Therapy of the glaucomas. St. Louis: CV Mosby, 1989:308–350

459. Toris CB, Yablonski ME, Wang YL, et al: Aqueous humor hynamics in the aging human eye. Am J Ophthalmol 127:407, 1999

460. Sperber GO, Bill A: A method for near-continuous determination of aqueous humor flow: Effects of anaesthetics, temperature and indomethacin. Exp Eye Res 39:435, 1984

461. Suguro K, Toris CB, Pederson JE: Uveoscleral outflow following cyclodialysis in the monkey eye using a fluorescent tracer. Invest Ophthalmol Vis Sci 26:810, 1985

462. Friedenwald JS: Contribution to the theory and practice of tonometry. Am J Ophthalmol 20:985, 1937

463. Friedenwald JS: Tonometer calibration: An attempt to remove discrepancies found in the 1954 calibration scale for Schiotz tonometers. Trans Am Acad Ophthalmol Otolaryngol 61:108, 1957

464. Prijot E, Weekers R: Mesure de la resistance a l'ecoulement de l'humeur aqueuse au moyen du tonometre electonique. Ophthalmologica 123:1, 1952

465. Topper JE, McLaren J, Brubaker RF: Measurement of aqueous humor flow with scanning ocular fluorophotometers. Curr Eye Res 3:1391–1395,1984

466. Rosengren B: A method for producing intraocular rise of tension. Acta Ophthalmol 12:403, 1934

467. Itoi M: Pv tonography: Tonometers for Pv tonography. Nippon Ganka Gakkai Zasshi 76:97, 1972

468. Yablonski ME, Cook DJ, Gray J: A fluorophotometric study of the effect of argon laser trabeculoplasty on aqueous humor dynamics. Am J Ophthalmol 99:579, 1985

469. Toris CB, Gleason ML, Camras CB, et al: Effects of brimonidine on aqueous humor dynamics in human eyes. Arch Ophthalmol 113:1514, 1995

470. Toris CB, Tafoya ME, Camras CB, et al: Effects of apraclonidine on aqueous humor dynamics in human eyes. Ophthalmology. 102:456–461,1995

471. Toris CB, Zhan GL, Camras CB, McLaughlin MA: Effects of travoprost on aqueous humor dynamics in monkeys. J Glaucoma 14:70, 2005

472. McLaughlin MA, Toris CB, Brooks DE, et al: Effects of Rho kinase inhibitors on intraocular pressure (IOP) and aqueous humor dynamics in rabbits and monkeys [Abstract 3534]. Invest Ophthalmol Vis Sci 45(suppl):S3534, 2004

473. Tafoya ME, Toris CB, Camras CB, et al: Effects of apraclonidine on aqueous humor dynamics in human eyes. Invest Ophthalmol Vis Sci 34:929, 1993

474. Toris CB, Yablonski ME, Wang YL, Camras CB: Aqueous humor dynamics in the aging human eye. Am J Ophthalmol 127:407, 1999

475. Toris CB, Zhan GL, Zhao J, et al: Potential mechanism for the additivity of pilocarpine and latanoprost. Am J Ophthalmol 131:722, 2001

476. Toris CB, Camras CB, Yablonski ME: Aqueous humor dynamics in ocular hypertensive patients. J Glaucoma 11:253, 2002

477. Ashton N, Brini A, Smith R: Anatomical studies of the trabecular meshwork of the normal human eye. Br J Ophthalmol 40:257, 1956

478. Rohen JW, Lütjen E, Bárány E: The relation between the ciliary muscle and the trabecular meshwork and its importance for the effect of miotics on aqueous outflow resistance. Albrecht von Graefes Arch Klin Exp Ophthalmol 172:23, 1967

479. Rohen JW, Futa R, Lütjen-Drecoll E: The fine structure of the cribriform meshwork in normal and glaucomatous eyes as seen in tangential sections. Invest Ophthalmol Vis Sci 21:574, 1981

480. Vegge T: Ultrastructure of normal human trabecular endothelium. Acta Ophthalmol 41:193, 1963

481. Holmberg AS: Schlemm's canal and the trabecular meshwork. An electron microscopic study of the normal structure in man and monkey (Cercopithecus ethiops). Doc Ophthalmol 19:339, 1965

482. Spencer WH, Alvarado J, Hayes TL: Scanning electron microscopy of human ocular tissues: The trabecular meshwork. Invest Ophthalmol 7:651, 1968

483. Peterson WS, Jocson VL: Hyaluronidase effects on aqueous outflow resistance: Quantitative and localizing studies in the rhesus monkey eye. Am J Ophthalmol 77:573, 1974

484. Grant WM: Experimental aqueous perfusion in enucleated human eyes. Arch Ophthalmol 69:783, 1963

485. Ellingsen BA, Grant WM: Influence of intraocular pressure and trabeculotomy on aqueous outflow in enucleated monkey eyes. Invest Ophthalmol 10:705, 1971

486. Ellingsen BA, Grant WM: Trabeculotomy and sinusotomy in enucleated human eyes. Invest Ophthalmol 11:21, 1972

487. Erickson-Lamy K, Rohen JW, Grant WM: Outflow facility studies in the perfused human ocular anterior segment. Exp Eye Res 52:723, 1991

488. Bahler CK, Fautsch MP, Hann CR, et al: Factors influencing intraocular pressure in cultured human anterior segments. Invest Ophthalmol Vis Sci 45:3137, 2004

489. Bahler CK, Hann CR, Fautsch MP, et al: Pharmacologic disruption of Schlemm's canal cells and outflow facility in anterior segments of human eyes. Invest Ophthalmol Vis Sci 45:2246, 2004

490. Rosenquist R, Epstein DL, Melamed S, et al: Outflow resistance of enucleated human eyes perfused at two different perfusion pressures and different extents of trabeculotomy. Curr Eye Res 8:1233, 1989

491. Inomata H, Bill A, Smelser GK: Aqueous humor pathways through the trabecular meshwork and into Schlemm's canal in the cynomolgus monkey (Macaca irus): An electron microscopic study. Am J Ophthalmol 73:760, 1972

492. Raviola G, Raviola E: Paracellular route of aqeuous outflow in the trabecular meshwork and canal of Schlemm. A freeze-fracture study of the endothelial junctions in the sclerocorneal angle of the macaque monkey eye. Invest Ophthalmol 21:52, 1981

493. Alvarado JA, Yun AJ, Murphy CG: Juxtacanalicular tissue in primary open-angle glaucoma and in nonglaucomatous normals. Arch Ophthalmol 104:1517, 1986

494. Murphy CG, Johnson M, Alvarado JA: Juxtacanalicular tissue in pigmentary and primary open-angle glaucoma. The hydrodynamic role of pigment and other constituents. Arch Ophthalmol 110:1779, 1992

495. Hamard P, Valtot F, Sourdille P, et al: Confocal microscopic examination of trabecular meshwork removed during ab externo trabeculectomy. Br J Ophthalmol 86:1046, 2002

496. Gong H, Ruberti J, Overby D, et al: A new view of the human trabecular meshwork using quick-freeze, deep-etch electron microscopy. Exp Eye Res 75:347, 2002

497. Sabanay I, Gabelt BT, Tian B, et al: H-7 effects on structure and fluid conductance of monkey trabecular meshwork. Arch Ophthalmol 118:955, 2000

498. Sabanay I, Tian B, Gabelt BT, et al: Functional and structural reversibility of H-7 effects on the conventional aqueous outflow pathway in monkeys. Exp Eye Res 78:137, 2004

499. Overby D, Gong H, Qiu G, et al: The mechanism of increasing outflow facility during washout in the bovine eye. Invest Ophthalmol Vis Sci 42:3455, 2002

500. Homberg AS: Our present knowledge of the structure of the trabecular meshwork. In Leydhecker W, ed.: Glaucoma Twentieth International Congress of Ophthalmology, Tutzing Symposium. Basel: S. Karger, 1966

501. Johnstone MA: The morphology of the aqueous outflow system. In Drance SM, Neufeld AH, eds.: Glaucoma: Applied Pharmacology. New York: Grune & Stratton, 1984:87–109

502. Feeney L: Outflow studies using an electron dense tracer. Trans Am Acad Ophthalmol Otolaryngol 70:792, 1966

503. Anderson DR: Scanning electron microscopy of primate trabecular meshwork. Am J Ophthalmol 71:90, 1971

504. Kayes J: Pressure gradient changes on the trabecular meshwork of monkeys. Am J Ophthalmol 79:549, 1975

505. Shabo AL, Reese TS, Gaasterland D: Postmortem formation of giant endothelial vacuoles in Schlemm's canal of monkey. Am J Ophthalmol 76:896, 1973

506. van Buskirk EM, Grant WM: Lens depression and aqueous outflow in enucleated primate eyes. Am J Ophthalmol 76:632, 1973

507. Tripathi RC: The functional morphology of the outflow systems of ocular and cerebrospinal fluids. Exp Eye Res 25(suppl):65, 1977

508. Tripathi RC: Mechanism of the aqueous outflow across the trabecular wall of Schlemm's canal. Exp Eye Res 11:116, 1971

509. Tripathi RC: Aqueous outflow pathway in normal and glaucomatous eyes. Br J Ophthalmol 56:157, 1972

510. Cole DF, Tripathi RC: Theoretical considerations on the mechanism of the aqueous outflow. Exp Eye Res 12:25, 1971

511. Epstein DL, Rohen JW: Morphology of the trabecular meshwork and inner wall endothelium after cationized ferritin perfusion in the monkey eye. Invest Ophthalmol Vis Sci 32:160, 1991

512. Grierson I, Lee WR: Pressure-induced changes in the ultrastructure of the endothelium lining of Schlemm's canal. Am J Ophthalmol 80:863, 1975

513. Brilakis HS, Johnson DH: Giant vacuole survival time and implications for aqeuous humor outflow. J Glaucoma 10:277, 2001

514. Bill A, Svedbergh B: Scanning electron microscopic studies of the trabecular meshwork and the canal of Schlemm: An attempt to localize the main resistance to outflow of aqueous humor in man. Acta Ophthalmol 50:295, 1972

515. Johnson M, Shapiro A, Ethier CR, et al: Modulation of outflow resistance by the pores of the inner wall endothelium. Invest Ophthalmol Vis Sci 33:1670, 1992

516. Johnson M, Chan D, Read AT, et al: The pore density in the inner wall endothelium of Schlemm's canal of glaucomatous eyes. Invest Ophthalmol Vis Sci 43:2950, 2002

517. Pollack IP, Becker B, Constant MA: The effect of hypothermia on aqueous humor dynamics: I. Intraocular pressure and outflow facility of the rabbit eye. Am J Ophthalmol 49:1126, 1960

518. Stamer WD, Snyder RW, Smith BL, et al: Localization of aquaporin CHIP in the human eye: Implications in the pathogenesis of glaucoma and other cisorders of ocular fluid balance. Invest Ophthalmol Vis Sci 35:3867, 1994

519. Hamann S, Zeuthern T, La Cour M, et al: Aquaporins in complex tissues: Distribution of aquaporins 1-5 in human and rat eye. Am J Physiol 274:C1332, 1998

520. Stamer WD, Seftor RE, Snyder RW, et al: Cultured human trabecular meshwork cells express aquaporin-1 water channels. Curr Eye Res 14:1095, 1995

521. Stamer WD, Peppel K, O'Donnell ME, et al: Expression of aquaporin-1 in human trabecular meshwork cells: Role in resting cell volume. Invest Ophthalmol Vis Sci 42:1803, 2001

522. Brandt JD, O'Donnell ME: How does the trabecular meshwork regulate outflow? Clues from the vascular endothelium [Review]. J Glaucoma 8:328, 1999

523. O'Donnell ME, Brandt JD, Curry F-RE: Na-K-Cl cotransport regulates intracellular volume and monolayer permeablility of trabecular meshwork cells. Am J Physiol 268:C1067, 1995

524. Al-Aswad LA, Gong H, Lee D, et al: Effects of Na-K-2Cl cotransport regulators on outflow facility in calf and human eyes in vitro. Invest Ophthalmol Vis Sci 40:1695, 1999

525. Gabelt BT, Wiederholt M, Clark AF, et al: Anterior segment physiology following bumetanide inhibition of Na-K-Cl cotransport. Invest Ophthalmol Vis Sci 38:1700, 1997

526. Putney LK, Brandt JD, O'Donnell ME: Na-K-Cl cotransport in normal and glaucomatous human trabecular meshwork cells. Invest Ophthalmol Vis Sci 40:425, 1999

527. Mitchell CH, Fleischhauer JC, Stamer WD, et al: Human trabecular meshwork cell volume regulation. Am J Physiol Cell Physiol. 283:C315, 2002

528. Johnstone MA: The aqueous outflow system as a mechanical pump. Evidence from examination of tissue and aqueous movement in human and non-human primates. J Glaucoma 13:421, 2004

529. Tamm ER, Flügel C, Stefani FH, et al: Nerve endings with structural characteristics of mechanoreceptors in the human scleral spur. Invest Ophthalmol Vis Sci 35:1157, 1994

530. Ruskell GL: Trigeminal innervation of the scleral spur in cynomolgus monkeys. J Anat 184:511, 1994

531. Tamm ER, Koch TA, Mayer B, et al: Innervation of myofibroblast-like scleral spur cells in human and monkey eyes. Invest Ophthalmol Vis Sci 36:1633, 1995

532. Holland M, Sallman LV, Collins E: A study of the innervation of the chamber angle. Am J Ophthalmol 42:148, 1956

533. Stone RA, McGlinn AM: Calcitonin gene-related peptide immunoreactive nerves in human and rhesus monkey eyes. Invest Ophthalmol Vis Sci 29:305, 1988

534. Laties AM, Stone RA, Brecha NC: Substance P-like immunoreactive nerve fibers in the trabecular meshwork. Invest Ophthalmol Vis Sci 21:484, 1981

535. Stone RA, Laties AM, Brecha NC: Substance P-like immunoreactive nerves in the anterior segment of the rabbit, cat, and monkey eye. Neuroscience 7:2459, 1982

536. Stone RA, Laties AM, Emson PC: Neuropeptide Y and the ocular innervation of rat, guinea pig, cat, and monkey. Neuroscience. 17:1207, 1986

537. Selbach JM, Gottanka J, Wittmann M, et al: Efferent and afferent innervation of primate trabecular meshwork and scleral spur. Invest Ophthalmol Vis Sci 41:2184, 2000

538. Gabelt BT, Robinson JC, Gange SJ, et al: Superior cervical ganglionectomy in monkeys: Aqueous humor dynamics and their responses to drugs. Exp Eye Res 60:575, 1995

539. Wentworth WO, Brubaker RF: Aqueous humor dynamics in a series of patients with third neuron Horner's syndrome. Am J Ophthalmol 92:407, 1981

540. Hubbard WC, Robinson JC, Schmidt K, et al: Superior cervical ganglionectomy in monkeys: Effects on refraction and intraocular pressure. Exp Eye Res 68:637, 1999

541. Marshall GE, Konstas AG, Lee WR: Immunogold localization of type IV collagen and laminin in the aging human outflow system. Exp Eye Res 51:691, 1990

542. Marshall GE, Konstas AG, Lee WR: Immunogold ultrastructural localization of collagens in the aged human outflow system. Ophthalmology. 98:692, 1991

543. Hann CR, Springett MJ, Wang X, et al: Ultrastructural localization of collagen IV, fibronectin, and laminin in the trabecular meshwork of normal and glaucomatous eyes. Ophthalmic Res 33:314, 2001

544. Ueda J, Wentz-Hunter K, Yue BYJT: Distribution of myocilin and extracellular matrix components in the juxtacanalicular tissue of human eyes. Invest Ophthalmol Vis Sci 43:1068, 2002

545. Bhatt K, Gong H, Freddo TF: Freeze-fracture studies of interendothelial junctions in the angle of the human eye. Invest Ophthalmol Vis Sci 36:1379, 1995

546. Ye W, Gong H, Sit A, et al: Interendothelial junctions in normal human schlemm's canal respond to changes in pressure. Invest Ophthalmol Vis Sci 38:2460, 1997

547. Alvarado JA, Murphy CG: Outflow obstruction in pigmentary and primary open angle glaucoma. Arch Ophthalmol 110:1769, 1992

548. Shirato S, Murphy CG, Bloom E, et al: Kinetics of phagocytosis in trabecular meshwork cells: Flow cytometry and morphometry. Invest Ophthalmol Vis Sci 30:2499, 1989

549. Zhou L, Fukuchi T, Kawa JE, et al: Loss of cell-matrix cohesiveness after phagocytosis by trabecular meshwork cells. Invest Ophthalmol Vis Sci 36:787, 1995

550. Kaufman PL, Gabelt BT: Presbyopia, prostaglandins and primary open angle glaucoma. In Krieglstein GK, ed.: Glaucoma Update V Proceedings of the Symposium of the Glaucoma Society of the International Congress of Ophthalmology in Quebec City, June 1994. New York-Berlin: Springer-Verlag, 1995:224–241

551. Funk RH, Rohen JW: Scanning electron microscopic study of episcleral arterioveous anastomoses in the owl and cynomolgus monkey. Curr Eye Res 15:321, 1996

552. Ascher KW: Aqueous veins: Their status eleven years after their detection. Arch Ophthalmol 49:438, 1953

553. Sears ML: The mechanism of action of adrenergic drugs in glaucoma. Invest Ophthalmol 5:115, 1966

554. Krasnov MM: Microsurgery of glaucoma. Indications and choice of techniques. Am J Ophthalmol 67:857, 1969

555. Ellingsen BA, Grant WM: The relationship of pressure and aqueous outflow in enucleated human eyes. Invest Ophthalmol 10:430, 1971

556. Lütjen-Drecoll E, Futa R, Rohen JW: Ultrahistochemical studies on tangential sections of the trabecular meshwork in normal and glaucomatous eyes. Invest Ophthalmol Vis Sci 21:563, 1981

557. Rohen JW: The evolution of the primate eye in relation to the problem of glaucoma. In Lutjen-Drecoll E, ed.: Basic Aspects of Glaucoma Research. Stuttgart: Schattauer, 1982:3–33

558. Rohen JW: Why is intraocular pressure elevated in chronic simple glaucoma? Anatomical considerations. Ophthalmology 90:758, 1983

559. Lütjen-Drecoll E, Rohen JW: Morphologic basis for drug action on aqueous outflow. In Drance SM, ed.: Applied Pharmacology in the Medical Treatment of Glaucomas. New York: Grune & Stratton, 1984:459–475

560. Tamm E, Flügel C, Stefani FH, et al: Contractile cells in the human scleral spur. Exp Eye Res 54:531, 1992

561. Kaufman PL, Bárány EH: Residual pilocarpine effects on outflow facility after ciliary muscle disinsertion in the cynomolgus monkey. Invest Ophthalmol 15:558, 1976

562. Kaufman PL, Bárány EH: Loss of acute pilocarpine effect on outflow facility following surgical disinsertion and retrodisplacement of the ciliary muscle from the scleral spur in the cynomolgus monkey. Invest Ophthalmol 15:793, 1976

563. Kaufman PL, Bárány EH: Recent observations concerning the effects of cholinergic drugs on outflow facility in monkeys. In Bito LZ, Davson H, Fenstermacher JD, eds.: The Ocular and Cerebrospinal Fluids. Fogarty International Center Symposium, 1977:415–418

564. James RJ, Croft MA, Rasmussen CA, et al: Effect of centrally stimulated ciliary muscle contraction on outflow facility in young rhesus monkeys [Abstract 4669]. Invest Ophthalmol Vis Sci 45(suppl):S4669, 2004

565. Armaly MF, Burian HM: Changes in the tonogram during accommodation. Arch Ophthalmol 60:60, 1958

566. Flügel C, Bárány EH, Lütjen-Drecoll E: Histochemical differences within the ciliary muscle and its function in accommodation. Exp Eye Res 50:219, 1990

567. Erickson-Lamy KA, Kaufman PL: Effect of cholinergic drugs on outflow facility after ciliary ganglionectomy. Invest Ophthalmol Vis Sci 29:491, 1988

568. Erickson-Lamy K, Schroeder A: Dissociation between the effect of aceclidine on outflow facility and accommodation. Exp Eye Res 50:143, 1990

569. Gabelt BT, Kaufman PL: Inhibition of outflow facility, accommodative, and miotic responses to pilocarpine in rhesus monkeys by muscarinic receptor subtype antagonists. J Pharmacol Exp Ther 263:1133, 1992

570. Gabelt BT, Kaufman PL: Inhibition of aceclidine-stimulated outflow facility, accommodation and miosis by muscarinic receptor subtype antagonists in rhesus monkeys. Exp Eye Res 58:623, 1994

571. Poyer JF, Kaufman PL, Flügel C: Age does not affect contractile responses of the isolated rhesus monkey ciliary muscle to muscarinic agonists. Curr Eye Res 12:413, 1993

572. Poyer JF, Gabelt BT, Kaufman PL: The effect of muscarinic agonists and selective receptor subtype antagonists on the contractile response of the isolated rhesus monkeys ciliary muscle. Exp Eye Res 59:729, 1994

573. Ehinger B: Adrenergic nerves to the eye and to related structures in man and in the cynomolgus monkey (Macaca irus). Invest Ophthalmol 5:42, 1966

574. van Alphen GWHM, Robinette SL, Macri FJ: Drug effects on ciliary muscle and choroid preparations in vitro. Arch Ophthalmol 68:111, 1962

575. van Alphen GWHM, Kern R, Robinette SL: Adrenergic receptors of the intraocular muscles. Comparison to cat rabbit, and monkey. Arch Ophthalmol 74:253, 1965

576. Törnqvist G: Effect of cervical sympathetic stimulation on accommodation in monkeys. An example of a beta-adrenergic inhibitory effect. Acta Physiol Scand. 67:363, 1966

577. Kaufman PL: The effects of drugs on the outflow of aqueous humor. In Drance SM, Neufeld AH, eds.: Glaucoma: Applied Pharmacology in Medical Treatment. Orlando: Grune & Stratton, 1984:429–458

578. Kaufman PL, Bárány EH: Adrenergic drug effects on aqueous outflow facility following ciliary muscle retrodisplacement in the cynomolgus monkey. Invest Ophthalmol Vis Sci 20:644, 1981

579. Nilsson SFE, Bill A: Physiology and neurophysiology of aqueous humor inflow and outflow. In Kaufman PL, Mittag TW, eds.: Glaucoma. London: Mosby-Yearbook Europe Ltd., 1994:1.17–1.34

580. Wax MB, Coca-Prados M: Receptor-mediated phosphoinositide hydrolysis in human ocular ciliary epithelial cells. Invest Ophthalmol Vis Sci 30:1675, 1989

581. Alvarado JA, Franse-Carman L, McHolm G, et al: Epinephrine effects on major cell types of the aqueous outflow pathway: In vitro studies/clinical implications. Trans Am Ophthalmol Soc 88:267, 1990

582. Sears ML, Neufeld AH: Adrenergic modulation of the outflow of aqueous humor. Invest Ophthalmol 14:83, 1975

583. Thomas JV, Epstein DL: Timolol and epinephrine in primary open angle glaucoma. Transient additive effect. Arch Ophthalmol 99:91, 1981

584. Cyrlin MN, Thomas JV, Epstein DL: Additive effect of epinephrine to timolol therapy in primary open-angle glaucoma. Arch Ophthalmol 100:414, 1982

585. Allen RC, Epstein DL: Additive effects of betaxolol and epinephrine in primary open angle glaucoma. Arch Ophthalmol 104:1178, 1986

586. Robinson JC, Kaufman PL: Effects and interactions of epinephrine, norepinephrine, timolol and betaxolol on outflow facility in the cynomolgus monkey. Am J Ophthalmol 109:189, 1990

587. Busch MJWM, van Oosterhout JGM, Hoyng PFJ: Effects of cyclic nucleotide analogs on intraocular pressure and trauma-induced inflammation in the rabbit eye. Curr Eye Res 11:5, 1992

588. Neufeld AH, Dueker DK, Vegge T, et al: Adenosine 3',5'-monophosphate increases the outflow of aqueous humor from the rabit eye. Invest Ophthalmol Vis Sci 14:40, 1975

589. Kaufman PL: Adenosine 3',5' cyclic monophosphate and outflow facility in monkey eyes with intact and retrodisplaced ciliary muscle. Exp Eye Res 44:415, 1987

590. Erickson-Lamy KA, Nathanson JA: Epinephrine increases facility of outflow and cyclic AMP content in the human eye in vitro. Invest Ophthalmol Vis Sci 33:2672, 1992

591. Crawford KS, Robinson JC, Heideman W, et al: Indomethacin and epinephrine effects on outflow facility and cAMP formation in monkeys. Invest Ophthalmol Vis Sci 37:1348, 1996

592. Erickson K, Liang L, Shum P, et al: Adrenergic regulation of aqueous outflow. J Ocular Pharmacol 10:241, 1994

593. Alvarado JA, Murphy CG, Franse-Carman L, et al: Effect of β-adrenergic agonists on paracellular width and fluid flow across outflow pathway cells. Invest Ophthalmol Vis Sci 39:1813, 1998

594. Wiederholt M: Direct involvement of trabecular meshwork in the regulation of aqueous humor outflow. Curr Opin Ophthalmol. 9:46, 1998

595. Robinson JC, Kaufman PL: Cytochalasin B potentiates epinephrine's outflow facility increasing effect. Invest Ophthalmol Vis Sci 32:1614, 1991

596. Robinson JC, Kaufman PL: Phalloidin inhibits epinephrine's and cytochalsin B's facilitation of aqueous outflow. Arch Ophthalmol 112:1610, 1994

597. Tripathi BJ, Tripathi RC: Effect of epinephrine in vitro on the morphology, phagocytosis, and mitotic activity of human trabecular endothelium. Exp Eye Res 39:731, 1984

598. Lütjen-Drecoll E, Shimuzu R, Rohrbach M, et al: Quantitative analysis of "plaque material" in the inner and outer wall of Schlemm's canal in normal and glaucomatous eyes. Exp Eye Res 42:443, 1986

599. Rodriques M, Spaeth G, Weinreb S, et al: Spectrum of trabecular pigmentation in open-angle glaucoma: A clinicopathologi study. Trans Am Acad Ophthalmol Otolaryngol 81:258, 1976

600. Alvarado J, Murphy C, Juster R, et al: Studies on the pathogenesis of primary open-angle glaucoma: Regional analysis of trabecular meshwork cellularity and dense collagen. In Ticho U, David R, eds.: Recent Advances in Glaucoma. New York: Elsevier Science Publishing, 1984:3–8

601. Grierson I: What is open angle glaucoma? Eye 1:15, 1987

602. Rohen JW, Lutjen-Drecoll E: Age changes of the trabecular meshwork in human and monkey eyes. A light and electron microscopic study. Altern Entwickl Aging Dev 1:1, 1971

603. Gabelt BT, Kaufman PL: Changes in aqueous humor dynamics with age and glaucoma. Prog Retin Eye Res May 23:[Epub ahead of print], 2005

604. Tian B, Geiger B, Epstein DL, et al: Cytoskeletal involvement in the regulation of aqueous humor outflow. Invest Ophthalmol Vis Sci 41:619, 2000

605. Volberg T, Geiger B, Citi S, et al: Effect of protein kinase inhibitor H-7 on the contractility, integrity and membrane anchorage of the microfilament system. Cell Motility Cytoskel 29:321, 1994

606. Birrell GB, Hedberg KK, Habliston DL, et al: Protein kinase C inhibitor H-7 alters the actin cytoskeleton of cultured cells. J Cell Physiol 141:74, 1989

607. Liu X, S: C, Glasser A, et al. Effect of H-7 on cultured human trabecular meshwork cells. Mol Vis Jun 27;7:145-53, 2001

608. Yu JC, Gottlieb AI: Disruption of endothelial actin microfilaments by protein kinase C inhibitors. Microvasc Res 43:100, 1992

609. Bershadsky A, Chausovsky A, Becker E, et al: Involvement of microtubules in the control of adhesion-dependent signal transduction. Curr Biol 6:1279, 1996

610. Gills JP, Roberts BC, Epstein DL: Microtubule disruption leads to cellular contraction in human trabecular meshwork cells. Invest Ophthalmol Vis Sci 39:653, 1998

611. Spector I, Shochet NR, Kashman Y, et al: Latrunculins: Novel marine toxins that disrupt microfilament organization in cultured cells. Science 219:493, 1983

612. Coué M, Brenner SL, Spector I, et al: Inhibition of actin polymerization by latrunculin A. FEBS Lett 213:316, 1987

613. Cai S, Liu X, Glasser A, et al: Effect of latrunculin-A on morphology and actin-associated adhesions of cultured human trabecular meshwork cells. Mol Vis 6:132, 2000

614. Perkins TW, Alvarado JA, Polansky JR, et al: Trabecular meshwork cells grown on filters: Conductivity and cytochalasin effects. Invest Ophthalmol Vis Sci 29:1836, 1988

615. Tian B, Kaufman PL, Volber T, et al: H-7 disrupts the actin cytoskeleton and increases outflow facility. Arch Ophthalmol 116:633, 1998

616. Tian B, Gabelt BT, Peterson JA, et al: H-7 increases trabecular facility and facility after ciliary muscle disinsertion in monkeys. Invest Ophthalmol Vis Sci 40:239, 1999

617. Peterson JA, Tian B, Bershadsky AD, et al: Latrunculin-A increases outflow facility in the monkey. Invest Ophthalmol Vis Sci 40:931, 1999

618. Svedbergh B, Lütjen-Drecoll E, Ober M, et al: Cytochalasin B-induced structural changes in the anterior ocular segment of the cynomolgus monkey. Invest Ophthalmol Vis Sci 17:718, 1978

619. Johnstone M, Tanner D, Chau B, et al: Concentration-dependent morphologic effects of cytochalasin B in the aqueous outflow system. Invest Ophthalmol Vis Sci 19:835, 1980

620. Kaufman PL, Bárány EH: Cytochalasin B reversibly increases outflow facility in the eye of the cynomolgus monkey. Invest Ophthalmol Vis Sci 16:47, 1977

621. Kaufman PL, Erickson KA: Cytochalasin B and D dose-outflow facility response relationships in the cynomolgus monkey. Invest Ophthalmol Vis Sci 23:646, 1982

622. Kaufman PL: Non-additivity of maximal pilocarpine and cytochalasin effects on outflow facility. Exp Eye Res 44:283, 1987

623. Bill A: Effects of Na2EDTA and alpha-chymotrypsin on aqueous humor outflow conductance in monkey eyes. Uppsala J Med Sci 85:311, 1980

624. Bill A, Lutjen-Drecoll E, Svedbergh B: Effects of intracameral Na2EDTA and EGTA on aqueous outflow routes in the monkey eye. Invest Ophthalmol Vis Sci 19:492, 1980

625. Hamanaka T, Bill A: Morphological and functional effects of Na2EDTA on the outflow routes for aqueous humor in monkeys. Exp Eye Res 44:171, 1987

626. Tian B, Gabelt BT, Kaufman PL: Effect of staurosporine on outflow facility in monkeys. Invest Ophthalmol Vis Sci 40:1009, 1999

627. Tian B, Brumback LC, Kaufman PL: ML-7, chelerythrine and phorbol ester increase outflow facility in the monkey eye. Exp Eye Res 71:551, 2000

628. Tian B, Kiland JA, Kaufman PL: Effects of the marine macrolides swinholide A and jasplakinolide on outflow facility in monkeys. Invest Ophthalmol Vis Sci 42:3187, 2001

629. Peterson JA, Tian B, McLaren JW, et al: Latrunculin's effects on intraocular pressure, aqueous humor flow and corneal endothelium. Invest Ophthalmol Vis Sci 41:1749, 2000

630. Peterson JA, Tian B, Geiger B, et al: Effect of latrunculin-B on outflow facility in monkeys. Exp Eye Res 70:307, 2000

631. Tian B, Kaufman PL: Effects of the Rho kinase inhibitor Y-27632 and the phosphatase inhibitor calyculin on outflow facility in monkeys. Invest Exp Eye Res 80:215, 2005

632. Epstein DL, Rowlette L-L, Roberts BC: Acto-myosin drug effects and aqueous outflow function. Invest Ophthalmol Vis Sci 40:74, 1999

633. Honjo M, Tankhara H, Inatani M, et al: Effects of rho-associated protein kinase inhibitor Y-27632 on intraocular pressure and outflow facility. Invest Ophthalmol Vis Sci 42:137, 2001

634. Honjo M, Inatani M, Kido N, et al: A myosin light chain kinase inhibitor, ML-9, lowers the intraocular pressure in rabbit eyes. Exp Eye Res 75:135, 2002

635. Waki M, Yoshida Y, Oka T, et al: Reduction of intraocular pressure by topical administration of an inhibitor of the rho-associated protein kinase. Curr Eye Res 22:470, 2001

636. Rao PV, Deng P-F, Kumar J, et al: Modulation of aqueous humor outflow facility by the rho kinase-specific inhibitor Y-27632. Invest Ophthalmol Vis Sci 42:1029, 2001

637. Sward K, Dreja K, Susnjar M, et al: Inhibition of Rho-associated kinase blocks agonist-induced Ca2+ sensitization of myosin phosphorylation and force in guinea-pig ileum. J Physiol 522:33, 2000

638. Burridge K, Chrzanowska-Wodnicka M: Focal adhesions, contractility, and signaling. Ann Rev Cell Devel Biol 12:463, 1996

639. Tian B, Wang R-F, Podos SM, et al: Effects of topical H-7 on outflow facility, intraocular pressure and corneal thickness in monkeys. Arch Ophthalmol 122:1171, 2004

640. Tian B, Sabanay I, Gabelt BT, et al: Latrunculin B effects on aqueous outflow and trabecular meshwork and corneal endothelium structure in the monkey eye [Abstract 2092]. Invest Ophthalmol Vis Sci 45:S2092, 2004

641. Kaufman PL: Aqueous humor outflow. In Zadunaisky JA, Davson H, eds.: Current Topics in Eye Research. New York: Academic Press, 1984:97–138

642. Sawaguchi S, Yue BY, Yeh P, et al: Effects of intracameral injection of chondroitinase ABC in vivo. Arch Ophthalmol 110:110, 1992

643. Bárány EH, Scotchbrook S: Influence of testicular hyaluronidase on the resistance to flow through the angle of the anterior chamber. Acta Physiol Scand. 30:240, 1954

644. Van Buskirk EM, Brett J: The canine eye: In vitro dissolution of the barriers to aqueous outflow. Invest Ophthalmol Vis Sci 17:258, 1978

645. Knepper PA, Farbman AI, Tesler AG: Exogenous hyaluronidases and degradation of hyaluronic acid in the rabbit eye. Invest Ophthalmol Vis Sci 25:286, 1984

646. Kaufman PL, Svedbergh B, Lütjen-Drecoll E: Medical trabeculocanalotomy in monkeys with cytochalasin B or EDTA [Editorial]. Ann Ophthalmol 11:795, 1979

647. Gabelt BT, Crawford K, Kaufman PL: Outflow facility and its response to pilocarpine decline in aging rhesus monkeys. Arch Ophthalmol 109:879, 1991

648. Liu X, Filla MS, Brandt CR, et al: Exoenzyme C3 transferase gene transfer to cultured human trabecular meshwork cells and ciliary muscle cells—Towards glaucoma gene therapy [Abstract 4563]. Invest Ophthalmol Vis Sci 45(suppl):S4563, 2004

649. Vittitow JL, Garg R, Rowlette LL, et al: Gene transfer of dominant-negative RhoA increases outflow facility in perfused human anterior segment cultuRes Mol Vis. 8:32, 2002

650. Kee C, Sohn S, Hwang J-M: Stromelysin gene transfer into cultured human trabecular cells and rat trabecular meshwork in vivo. Invest Ophthalmol Vis Sci 42:2856, 2001

651. Santas AJ, Bahler CK, Peterson JA, et al: Effect of heparin II domain of fibronectin on aqueous outflow in cultured anterior segments of human eyes. Invest Ophthalmol Vis Sci 44:4796, 2003

652. Becker B, Hahn K: Topical corticosteroids and heredity in primary open-angle glaucoma. Am J Ophthalmol 57:543, 1964

653. Becker B: Intraocular pressure response to topical corticosteroids. Invest Ophthalmol Vis Sci 4:198, 1965

654. Kolker AE, Becker B: Topical corticosteroids and glaucoma: Current status. In Bellows JG, ed.: Contemporary Ophthalmology. Baltimore: Williams & Wilkins, 1972:161–167

655. Woods DC, Contaxis I, Sweet D, et al: Response of rabbits to corticosteroids. I. Influence on growth, introacular pressure and lens transparency. Am J Ophthalmol 63:841, 1967

656. Bonomi L, Perfetti S, Noya E, et al: Experimental corticosteroid ocular hypertension in the rabbit. Albrcht Von Graefes Arch Klin Exp Ophthalmol 209:73, 1978

657. Francois J: Corticosteroid glaucoma. Ophthalmologica 188:76, 1984

658. Green K, Phillips CI, Gore SM, et al: Ocular fluid dynamics in response to topical RU486, a steroid blocker. Curr Eye Res 4:605, 1985

659. Tchernitchin A, Wenk EJ, Southren AL, et al: Radioautography of dexamethasone in rabbit gingiva and buccal mucosa. J Dental Res 59:2100, 1980

660. Weinreb RN, Bloom E, Baxter JD: Detection of glucocorticoid receptors in cultured human trabecular cells. Invest Ophthalmol Vis Sci 21:403, 1981

661. Hernandez MR, Wenk EJ, Weinstein BI, et al: Glucocorticoid target cells in human outflow pathway: Autopsy and surgical specimens. Invest Ophthalmol Vis Sci 24:1612, 1983

662. Armaly MF: Effect of corticosteroids on intraocular pressure and fluid dynamics. II. The effect of dexamethasone in the glaucomatous eye. Arch Ophthalmol 70:492, 1963

663. Fingert JH, Clark AF, Craig JE, et al: Evaluation of the myocilin (MYOC) glaucoma gene in monkey and human steroid-induced ocular hypertension. Invest Ophthalmol Vis Sci 42:145, 2001

664. Armaly MF: Effect of corticosteroids on intraocular pressure and fluid dynamics. I. The effect of dexamethasone in the nromal eye. Arch Ophthalmol 70:482, 1963

665. Levi L, Schwartz B: Decrease of ocular pressure with oral metyrapone. A double masked crossover trial. Arch Ophthalmol 105:777, 1987

666. Polansky JR, Palmberg P, Matulich D: Cellular sensitivity to glucocorticoids in patients with POAG. Steroid receptors and responses in cultured skin fibroblasts. Invest Ophthalmol Vis Sci 26:805, 1985

667. Schwartz B, Levene RZ: Plasma cortisol differences between normal and glaucomatous patients: Before and after dexamethasone suppression. Arch Ophthalmol 87:369, 1972

668. Schwartz B, McCarty G, Rosner B: Increased plasma free cortisol in ocular hypertension and open-angle glaucoma. Arch Ophthalmol 105:1060, 1987

669. Weinreb RN, Polansky JR, Kramer SG, et al: Acute effects of dexamethasone on intraocular pressure in glaucoma. Invest Ophthalmol Vis Sci 26:170, 1985

670. Southren AL, Gordon GG, Munnangi PR, et al: Altered cortisol metabolism in cells cultured from trabecular meshwork specimen obtained from patients with primary open-angle glaucoma. Invest Ophthalmol Vis Sci 26:1413, 1983

671. Weinstein BI, Munnangi P, Gordon GG, et al: Defects in cortisol-metabolizing enzymes in primary open-angle glaucoma. Invest Ophthalmol Vis Sci 26:890, 1985

672. Southren AL, Wandel T, Gordon GG, et al: Treatment of glaucoma with 3a,5b-tetrahydrocortisol: A new therapeutic modality. J Ocular Pharmacol 10:385, 1994

673. Clark AF, Lane D, Wilson K, et al: Inhibition of dexamethasone-induced cytoskeletal changes in cultured human trabecular meshwork cells by tetrahydrocortisol. Invest Ophthalmol Vis Sci 37:805, 1996

674. Rohen JW, Linner E, Witmer R: Electron microscopic studies on the trabecular meshwork in two cases of corticosteroid glaucoma. Exp Eye Res 17:19, 1973

675. Yun AJ, Murphy CG, Polansky JR, et al: Proteins secreted by human trabecular cells. Glucocorticoid and other effects. Invest Ophthalmol Vis Sci 30:2012, 1989

676. Steely HT, Browder S, Julian MB: The effects of dexamethasone on fibronectin expression in cultured trabecular meshwork cells. Invest Ophthalmol Vis Sci 33:2242, 1992

677. Johnson DH, Bradley JM, Acott TS: The effect of dexamethasone on glycosaminoglycans of human trabecular meshwork in perfusion organ culture. Invest Ophthalmol Vis Sci 31:2568, 1990

678. Zhou L, Li Y, Yue BYJT: Glucocorticoid effects on extracellular matrix proteins and integrins in bovine trabecular meshwork cells in relation to glaucoma. Int J Mol Med 1:339, 1998

679. Dickerson JEJ, Steely HTJ, English-Wright SL, et al: The effect of dexamethasone on integrin and laminin expression in cultured human trabecular meshwork cells. Exp Eye Res 66:731, 1998

680. Snyder RW, Stamer WD, Kramer TR, et al: Corticosteroid treatment and trabecular meshwork proteases in cell and organ culture supernatants. Exp Eye Res 57:461, 1993

681. Samples JR, Alexander JP, Acott TS: Regulation of the levels of human trabecular matrix metalloproteinases and inhibitor by interleukin-1 and dexamethasone. Invest Ophthalmol Vis Sci 34:3386, 1993

682. Clark AF, Wilson K, McCartney MD, et al: Glucocorticoid-induced formation of cross-linked actin networks in cultured human trabecular meshwork cells. Invest Ophthalmol Vis Sci 35:281, 1994

683. Wilson K, McCartney MD, Miggans ST, et al: Dexamethasone induced ultrastructural changes in cultured human trabecular meshwork cells. Exp Eye Res 12:783, 1993

684. Tripathi BJ, Tripathi RC, Swift HH: Hydrocortisone-induced DNA endoreplication in human trabecular cells in vitro. Exp Eye Res 49:259, 1989

685. Matsumoto Y, Johnson DH: Dexamethasone decreases phagocytosis by human trabecular meshwork cells in situ. Invest Ophthalmol Vis Sci 38:1902, 1997

686. Partridge CA, Weinstein BI, Southren AL: Dexamethasone induces specific proteins in human trabecular meshwork cells. Invest Ophthalmol Vis Sci 30:1843, 1989

687. Polansky JR, Kurtz RM, Alvarado JA, et al: Eicosanoid production and glucocorticoid regulatory mechanisms in cultured human trabecular meshwork cells. Prog Clin Biol Res 312:113, 1989

688. Nguyen TD, Huang WD, Bloom E, et al: Glucocorticoid (GC) effects on HTM cells: Molecular biology approaches. In Lütjen-Drecoll E, ed.: Basic Aspects of Glaucoma Research III. Stuttgart: Schattauer, 1993:331–343

689. Ueda J, Wentz-Hunter K, Cheng EL, et al: Ultrastructural localization of myocilin in human trabecular meshwork cells and tissues. J Histochem Cytochem 48:1321, 2000

690. Filla MS, Liu X, Nguyen TD, et al: In vitro localization of TIGR/MYOC in trabecular meshwork extracellular matrix and binding to fibronectin. Invest Ophthalmol Vis Sci 43:151, 2002

691. Mertts M, Garfield S, Tanemoto K, et al: Identification of the region in the N-terminal domain responsible for the cytoplasmic localization of MYOC/TIGR and its association with microtubules. Lab Invest 79:1237, 1999

692. Wentz-Hunter K, Ueda J, Shimizu N, et al: Myocilin is associated with mitochondria in human trabecular meshwork cells. J Cell Physiol 190:46, 2002

693. Stone EM, Fingert JH, Alward WLM, et al: Identification of a gene that causes primary open angle glaucoma. Science 275:668, 1997

694. Alward WL, Fingert JH, Coote MA, et al: Clinical features associated with mutations in the chromosome 1 open-angle glaucoma gene. N Engl J Med 338:1022, 1998

695. Tamm ER: Myocilin and glaucoma: Facts and ideas. Prog Retin Eye Res 21:395, 2002

696. Polansky JR, Fauss DJ, Chen P, et al: Cellular pharmacology and molecular biology of the trabecular meshwork inducible glucocorticoid response gene product. Ophthalmologica 211:126, 1997

697. Nguyen TD, Chen P, Huang WD, et al: Gene structure and properties of TIGR, an olfactomedin-related glycoprotein cloned from glucocorticoid-induced trabecular meshwork cells. J Biol Chem 273:6341, 1988

698. Wentz-Hunter K, Kubota R, Shen X, et al: Extracellular myocilin affects activity of human trabecular meshwork cells. J Cell Physiol. 200:45, 2004

699. Caballero M, Rowlette LL, Borrás T: Altered secretion of a TIGR/MYOC mutant lacking the olfactomedin domain. Bioch Biophys Acta 1502:447, 2000

700. Jacobson N, Andrews M, Shepard AR, et al: Non-secretion of mutant proteins of the glaucoma gene myocilin in cultured trabeuclar meshwork cells and in aqueous humor. Hum Mol Genet 10:117, 2001

701. Lo WR, Rowlette LL, Caballero M, et al: tissue differential microarray analysis of dexamethasone induction reveals potential mechanisms of steroid glaucoma. Invest Ophthalmol Vis Sci 44:473, 2003

702. Ishihashi T, Takagi Y, Mori K, et al: cDNA microarray analysis of gene expression changes induced by dexamethasone in cultured human trabecular meshwork cells. Invest Ophthalmol Vis Sci 43:3691, 2002

703. Underwood JL, Murphy CG, Chen J, et al: Glucocorticoids regulate transendothelial fluid flow resistance and formation of intercellular junctions. Am J Physiol 277:C330, 1999

704. Rohen JW, Lütjen-Drecoll E: Morphology of aqueous outflow pathways in normal and glaucomatous eyes. In Ritch R, Shields MB, Krupin T, eds.: The Glaucomas. St. Louis: CV Mosby, 1989:41

705. Lütjen-Drecoll E, Rohen JW: Pathology of the trabecular meshwork in primary open angle glaucoma. In Kaufman PL, Mittag TW, eds.: Glaucoma. Textbook of Ophthalmology Series. New York: CV Mosby, 1994

706. Johnson D, Gottanka J, Flügel C, et al: Ultrastructural changes in the trabecular meshwork of human eyes treated with corticosteroids. Arch Ophthalmol 115:375, 1997

707. Weinreb RN, Mitchell MD, Alvarado JA, et al: Glucocorticoid regulation of eicosandoid biosynthesis in cultured human trabecular cells. In Ticho U, David R, eds.: Recent Advances in Glaucoma. New York: Elsevier Sciences Publishers B.V., 1984:213–218

708. Polansky JR, Kurtz RM, Fauss DJ: In vitro correlates of glucocorticoid effects on intraocular pressure. In Krieglestein GK, ed.: Glaucoma Update IV. Berlin: Springer-Verlag, 1991:20–29

709. Polansky JR, Alvarado JA: Cellular mechanisms influencing the aqeuous humor outflow pathway. In Albert DM, Jakobiec FA, eds.: Principles and Practice of Ophthalmology: Basic Sciences. Philadelphia: WB Saunders, 1994:26

710. Polansky JR: HTM cell culture model for steroid effects on intraocular pressure: Overview. In Lütjen-Drecoll E, ed.: Basic Aspects of Glaucoma Research III. Stuttgart: Schattauer, 1993:307–318

711. Weinreb RN, Polansky JR, Alvarado JA, et al: Arachidonic acid metabolism in human trabecular meshwork cells. Invest Ophthal Visual Sci 29:1708, 1988

712. Lee PY, Podos SM, Severin C: Effect of prostaglandin F on aqueous humor dynamics of rabbit, cat, and monkey. Invest Ophthalmol Vis Sci 25:1087, 1984

713. Alm A, Villumsen J: PhXA34, a new potent ocular hypotensive drug. A study on dose-response relationship and on aqueous humor dynamics in healthy volunteers. Arch Ophthalmol 109:1564, 1991

714. Millard CB, Tripathi BJ, Tripathi RC: Age-related changes in protein profiles of the normal human trabecular meshwork. Exp Eye Res 45:623, 1987

715. Flügel C, Tamm E, Lütjen-Drecoll E, et al: Age-related loss of a-smooth muscle actin in normal and glaucomatous human trabecular meshwork of different age groups. J Glaucoma 1:165, 1992

716. Spiro RG: Nature of the glycoprotein components of basement membranes. Ann NY Acad Sci 312:106, 1978

717. Bornstein P, Duksin D, Balian G, et al: Organization of extracellular proteins on the connective tissue cell surface: Relevance to cell-matrix interactions in vitro and in vivo. Ann NY Acad Sci 312:93, 1978

718. Babizhayev MA, Brodskaya MW: Fibronectin detection in drainage outflow system of human eyes in ageing and progression of open angle glaucoma. Mech Ageing Dev 47:145, 1989

719. Bjorksten J: The crosslinkage theory of aging. J Am Geriatr Soc 16:408, 1968

720. Babizhayev MA, Bunin AY: Lipid peroxidation in open-angle glaucoma. Acta Ophthalmol 67:371, 1989

721. Green K: Free radicals and aging of anterior segment tissues of the eye: A hypothesis. Ophthalmic Res 27(suppl):143, 1995

722. Ochiai Y, Ochiai H: Higher concentration of transforming growth factor-beta in aqueous humor of glaucomatous eyes and diabetic eyes. Jpn J Ophthalmol 46:249, 2002

723. Picht G, Welge-Luessen U, Grehn F, et al: Transforming growth factor beta 2 levels in the aqeuous humor in different types of glaucoma and the relation to filtering bleb development. Graefes Arch Clin Exp Ophthalmol 239:199, 2001

724. Fuchshofer R, Welge-Lüssen U, Lütjen-Drecoll E: The effect of TGF-β2 on human trabecular meshwork extracellular proteolytic system. Exp Eye Res 77:757, 2003

725. Borisuth NS, Tripathi BJ, Tripathi RC: Identification and partial characterization of TGF-beta 1 receptors on trabecular cells. Invest Ophthalmol Vis Sci 33:596, 1992

726. Gottanka J, Chan D, Eichhorn M, et al: Effects of TGF-beta2 in perfused human eyes. Invest Ophthalmol Vis Sci 45:153, 2004

727. Lütjen-Drecoll E, Kaufman PL: Morphological changes in primate aqueous humor formation and drainage tissues after long-term treatment with antiglaucomatous drugs. J Glaucoma 2:316, 1993

728. Kiland JA, Gabelt BT, Kaufman PL: Studies on the mechanism of action of timolol and on the effects of suppression and redirection of aqueous flow on outflow facility. Exp Eye Res 78(Special Issue):639, 2004

729. Bill A, Phillips I: Uveoscleral drainage of aqueous humor in human eyes. Exp Eye Res 21:275, 1971

730. Bill A: Aqueous humor dynamics in monkeys (Macaca irus and Cercopithecus ethiops). Exp Eye Res 11:195, 1971

731. Gabelt BT, Gottanka J, Lütjen-Drecoll E, et al: Aqueous humor dynamics and trabecular meshwork and anterior ciliary muscle morphologic changes with age in rhesus monkeys. Invest Ophthalmol Vis Sci 44:2118, 2003

732. Kaufman PL: Physiology and phamacology of aqueous humor outflow: Implications for treatment. In Caprioli J, ed.: Contemporary Issues in Glaucoma. Philadelphia: WB Saunders, 1991:781–802

733. Emi K, Pederson JE, Toris CB: Hydrostatic pressure of the suprachoroidal space. Invest Ophthalmol Vis Sci 30:233, 1989

734. Poyer JF, Millar C, Kaufman PL: Prostaglandin F2a effects on isolated rhesus monkey ciliary muscle. Invest Ophthalmol Vis Sci 36:2461, 1995

735. Sagara T, Gaton DD, Lindsey JD, et al: Reduction of collagen type I in the ciliary muscle of inflamed monkey eyes. Invest Ophthalmol Vis Sci 40:2568, 1999

736. Sagara T, Gaton DD, Lindsey JD, et al: Topical prostaglandin F2a treatment reduces collagen types I, III, and IV in the monkey uveoscleral outflow pathway. Arch Ophthalmol 117:794, 1999

737. Lütjen-Drecoll E, Tamm E: Morphological study of the anterior segments of cynomolgus monkey eyes following treatment with prostaglandin F2a. Exp Eye Res 47:761, 1988

738. Tamm E, Rittig M, Lütjen-Drecoll E: Elektronenmikroskopische und immunhistochemische Untersuchungen zur augendrucksenkenden Wirkung von Prostaglandin F2a. Fortschr Ophthalmol 87:623, 1990

739. Gaton DD, Sagara T, Lindsey JD, et al: Increased matrix metalloproteinases 1, 2, and 3 in the monkey uveoscleral outflow pathway following topical prostaglandin F(2 alpha)-isopropyl ester treatment. Arch Ophthalmol 119:1165, 2001

740. Lindsey JD, Kashiwagi K, Kashiwagi F, et al: Prostaglandin action on ciliary smooth muscle extracellular matrix metabolism—Implications for uveoscleral outflow. Surv Ophthalmol 41(Suppl 2):S53, 1997

741. Kim J-W, Lindsey JD, Wang N, et al: Increased human scleral permeability with prostaglandin exposure. Invest Ophthalmol Vis Sci 42:1514, 2001

742. Weinreb RN: Enhancement of scleral macromolecular permeability iwht prostaglandins. Trans Am Ophthalmol Soc 99:319, 2001

743. Stjernschantz J, Selén G, Ocklind A, et al: Effects of latanoprost and related prostaglandin analogues. In Alm A, Weinreb RN, eds.: Uveoscleral Outflow Biology and Clinical Aspects. London: Mosby-Wolfe Medical Communications, 1998:57–72

744. Lee P-Y, Shoo H, Xu L, et al: The effect of prostaglandin F2a on intraocular pressure in normotensive human subjects. Invest Ophthalmol Vis Sci 29:1474, 1988

745. Bito LZ, Stjernschantz J, Resul B, et al: The ocular effects of prostaglandins and the therapeutic potential of a new PGF2α analog, PhXA41 (latanoprost), for glaucoma management. J Lipid Mediators 6:535, 1993

746. Camras CB: Comparison of latanoprost and timolol in patients with ocular hypertension and glaucoma. A six-month, masked, multicenter trial in the United States. Ophthalmology 103:138, 1996

747. Camras CB, Alm A, Watson P, et al: Latanoprost, a prostaglandin analog, for glaucoma therapy. Ophthalmol 103:1916, 1996

748. Parrish RK, Palmberg P, Sheu W-P, et al: A comparison of latanoprost, bimatoprost, and travoprost in patients with elevated intraocular pressure: A 12-week, randomized, masked-evaluator multicenter study. Am J Ophthalmol 135:688, 2003

749. Kurtz S, Shemesh G: The efficacy and safety of once-daily verss once-weekly latanoprost treatment for increased intraocular pressure. J Ocul Pharmacol Ther 20:321, 2004

750. Eisenberg DL, Toris CB, Camras CB: Bimatoprost and travoprost: A review of recent studies of two new glaucoma drugs. Surv Ophthalmol 47:S105, 2002

751. Noecker RS, Dirks MS, Choplin NT, et al: A six-month randomized clinical trial comparing the intraocular prssure-lowering efficacy of bimatoprost and latanoprost in patients with ocular hypertension or glaucoma. Am J Ophthalmol 135:55, 2003

752. Gandolfi SA, Cimino L: Effect of bimatoprost on patients with primary open-angle glaucoma or ocular hypertension who are nonresponders to loatanoprost. Ophthalmology 110:609, 2003

753. Kaufman PL: The prostaglandin wars [Editorial]. Am J Ophthalmol 136:727, 2003

754. Brubaker RF, Schoff EO, Nau CB, et al: Effects of AGN 192024, a new ocular hypotensive agent, on aqueous dynamics. Am J Ophthalmol 131:19, 2001

755. Christensen GA, Nau CB, McLaren JW, et al: Mechanism of ocular hypotensive action of bimatoprost (Lumigan) in patients with ocular hypertension or glaucoma. Ophthalmology 111:1658, 2004

756. Sharif NA, Kelly CR, Crider JY: Human trabecular meshwork cell responses induced by bimatoprost, travoprost, unoprostone, and other FP prostaglandin receptor agonist analogues. Invest Ophthalmol Vis Sci 44:715, 2003

757. Gagliuso DJ, Wang R-F, Mittag TW, et al: Additivity of bimatoprost or travoprost to latanoprost in glaucomatous monkey eyes. Arch Ophthalmol 122:1342, 2004

758. Richter M, Krauss AH-P, Woodward DF, et al: Morphological changes in the anterior eye segment after long-term treatment with different receptor selective prostaglandin agonists and a prostamide. Invest Ophthalmol Vis Sci 44:4419, 2003

759. Toris CB, Camras CB, Yablonski ME: Acute versus chronic effects of brimonidine on aqueous humor dynamics in ocular hypertensive patients. Am J Ophthalmol 128:8, 1999

760. Engstrom P, Dunham EW: Alpha-adrenergic stimulation of prostaglandin release from rabbit iris-ciliary body in vitro. Invest Ophthalmol Vis Sci 22:757, 1982

761. Yousufzai SY, Abdel-Latif AA: Effects of norepinephrine and other pharmacological agents on prostaglandin E2 release by rabbit and bovine irides. Exp Eye Res 37:279, 1983

762. Camras CB, Feldman SG, Podos SM, et al: Inhibition of the epinephrine-induced reduction of intraocular pressure by systemic indomethacin in humans. Am J Ophthalmol 100:169, 1985

763. Bhattacherjee P, Hammond BR: Effect of indomethacin on the ocular hypotensive action of adrenaline in the rabbit. Exp Eye Res 24:307, 1977

764. Lutjen-Drecoll E, Rohen JW: Pathology of the trabecular meshwork in primary open angle glaucoma. In Kaufman PL, Mittag TW, eds.: Glaucoma Textbook of Ophthalmology Series. New York: CV Mosby, 1994:1.1–1.16

765. Wiederholt M, Helbig H, Korbmacher C: Ion transport across the ciliary epithelium: Lessons from cultured cells and proposed role of the carbonic anhydrase. In Botrè F, Gross G, eds.: Carbonic Anhydrase. Cambridge: Verlag-Chemie, 1991:232–244

766. Maren TH: The kinetics of HCO3 synthesis related to fluid secretion, pH control, and CO2 elimination. Ann Rev Physiol 50:695, 1988

767. Kanski JJ: Glaucoma. In Kanski JJ, ed.: Clinical Ophthalmology, 2nd ed. London: Butterworth-Heinemann, 1989:182–231

768. Shields MB: Aqueous humor dynamics. I. Anatomy and physiology. In Shield MB, ed.: Textbook of Glaucoma, 4th ed. Baltimore: Williams & Wilkins, 1998:5–34

769. Wiederholt M, Stumpff F: The trabecular meshwork and aqueous humor reabsorption. In Civan MM, ed.: Current Topics in Membranes. San Diego: Academic Press, 1998:163–202

770. Kolker AE, Hetherington JJ: Gonioscopic and microscopic anatomy of the angle of the anterior chamber of the eye. In Kolker AE, Hetherington JJ, eds.: Becker-Schaffer's Diagnosis and Therapy of the Glaucomas, 4th ed. St Louis: Mosby-Year Book, 1976:21–41

771. Kaufman PL, Tian B, Gabelt BT, et al: Outflow enhancing drugs and gene therapy in glaucoma. In Weinreb R, Krieglstein G, Kitazawa Y, eds.: Glaucoma in the 21st Century. London: Harcourt-Mosby, 2000:117–128

Back to Top