Comprehensive Physiology Wiley Online Library

Urinary Concentration and Dilution: Models

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Whole‐Body Water and Solute Balance
1.1 Osmolal Clearance
1.2 Effect on Plasma Osmolality of Urine Flow of a Given Osmolality
1.3 Hypertonic Urine Implies a Hypotonic Tubular Reabsorbate and Vice Versa
1.4 Free Water Clearance
1.5 Antidiuretic Hormone Controls Water Excretion
2 Functional Implications of Renal Architecture
2.1 Production of Hyperosmotic Urine Is Related to the Loop of Henle
2.2 Significance of the S‐Shaped Configuration of the Mammalian Nephron for the Concentration of Urine
2.3 New Functional Features Introduced by the S‐Shaped Configuration
3 A Short History of Models of the Renal Counterflow System
3.1 Active Water Transport
3.2 Concentration by Differential Permeabilities and Counterflow
3.3 Concentration by Multiplication of a “Single Effect” along the Loop of Henle
3.4 Active Salt Transport out of the AHL Is the Probable Single Effect
3.5 Experimental Predictions of the Kuhn Models
3.6 Problems with the Kuhn Models
3.7 Vasa Recta Models
3.8 Nephrovascular Models
4 Equations for the Renal Medulla
4.1 Differential Equations
4.2 Integrated Mass Balance Equations
4.3 Generalized Multiplication Rule
4.4 Nephrovascular Coupling
4.5 Transmural Fluxes
5 Analytical Solution of the Differential Equations
5.1 Solutions for Single Flow Tubes
5.2 Analytical Solutions of Multitube Counterflow Problems
5.3 Solutions for Central Core Model
5.4 Externally Driven Central Core Multiplier
5.5 Multiplication in Inner and Outer Medullae
5.6 Non‐Zero Diffusion Coefficient
5.7 Solution of Time‐Dependent Equations
6 Numerical Solution of the Differential Equations
6.1 Method of Lines
6.2 Shooting Method
6.3 Quasilinearization
6.4 Finite Difference Methods
7 Results of Computer Simulation
7.1 Origin of Inner Medullary Concentration Gradient
7.2 Role of Collecting Duct
7.3 Role of Vasa Recta
7.4 Transition from Diuresis to Antidiuresis
8 Free‐Energy Balance in Renal Counterflow Systems
9 Summary and Future Directions
Figure 1. Figure 1.

Response of whole kidney to varying rates of vasopressin infusion in rats. Note the dramatic decrease in urine flow rate at higher rates of vasopressin infusion with minimal change in osmolar clearance

Data from Atherton et al. 3. From Knepper and Stephenson 111
Figure 2. Figure 2.

Relation between relative medullary thickness and maximum urinary osmolality. Long loops of Henle are required to make highly concentrated urine

Data from Schmidt‐Nielsen and O'Dell 179. From Burg and Stephenson 20
Figure 3. Figure 3.

Prototype three‐section glomerular nephron. With its straight configuration, solute and water transport from different segments is uncoupled. In the proximal tubule (PT) with a large effective hydraulic permeability, Lp, absorption is nearly isotonic; in the diluting segment (DIL) a smaller Lp limits water absorption relative to solute absorption, and tubule fluid is diluted; in final segment, hydraulic permeability is under control of antidiuretic hormone (ADH). In the absence of ADH tubule fluid remains dilute, and in the presence of ADH it equilibrates osmotically with the surrounding interstitium. Thus final urine can vary from markedly hypotonic to isotonic relative to plasma. Black arrows indicate solute movement; white arrows, volume flow.

Reproduced, with permission, from Stephenson 201, Annual Review of Biophysics and Bioengineering, Vol. 7. © 1978 by Annual Reviews, Inc
Figure 4. Figure 4.

The S‐shaped mammalian nephron. The essential functional feature that allows concentration of urine is juxtaposition of the diluting segment (DIL; now ascending limb of Henle's loop) with the water‐permeable descending limb and antidiuretic hormone (ADH)‐sensitive collecting duct. This permits solute supplied by ascending limb to extract water from collecting duct and descending limb. Black arrows indicate solute movement; white arrows, volume flow; PT, proximal tubule.

Reproduced, with permission, from Stephenson 201, Annual Review of Biophysics and Bioengineering, Vol. 7. © 1978 by Annual Reviews, Inc
Figure 5. Figure 5.

Initial stage of basic scheme for concentrating by permselective membranes. A very large volume of phenol solution at concentration Co is separated by membrane P permeable to phenol but not to water (rubber in experiments of Kuhn and Ryffel) from a large volume of sucrose solution, also at concentration Co. The sucrose solution is separated by membrane W permeable to water but not to sucrose or to phenol (copper ferrocyanide) from smaller volumes of phenol and sucrose solutions, also at concentration Co. As indicated by arrows, phenol diffuses into sucrose solution, increasing its osmolality and causing water to be osmotically extracted from the smaller volume solutions.

Adapted from Kuhn and Ryffel 122
Figure 6. Figure 6.

Basic scheme for concentrating by permselective membranes at equilibrium. Phenol has been added to the original sucrose solution so that its concentration has approximately doubled. Water has been extracted from the smaller volumes of sucrose and phenol so that their volumes are halved and concentrations doubled.

Adapted from Kuhn and Ryffel 122
Figure 7. Figure 7.

Arrangement for parallel processing of counterflowing sucrose and phenol solutions to yield progressive concentration of the sucrose solution. In the first apparatus on the left, sucrose and phenol, both at concentration Co, interact through a phenol‐permeable membrane P to give a combined solution of concentration 2Co. This interacts with counterflowing sucrose solution in the first apparatus on the right through a water‐permeable membrane W to give a sucrose solution of concentration 2Co. The phenol is then removed by diffusion into a concurrent flow of any solution at concentration Co through a second membrane P. As the solution is processed from left to right, there is progressive concentration of the sucrose solution. In further stages to the right, phenol reservoirs are not shown in order to simplify the diagram. Kuhn and Ryffel suggested that this scheme might mimic the operation of the kidney.

Adapted from Kuhn and Ryffel 122
Figure 8. Figure 8.

Actual scheme used by Kuhn and Ryffel. See text for explanation.

Adapted from Kuhn and Ryffel 122
Figure 9. Figure 9.

Hargitay and Kuhn countercurrent multiplier without withdrawal. Solution at concentration C1 enters at the left and reverses flow through hairpin constriction at the right to return to the left. The two counterflowing solutions are separated by a waterpermeable and solute‐impermeable membrane. Mass balance requires that at any position x along the tube C1 = C2 and F1 = ‐ F2. This means that the pressure difference Δp, the single effect, is unapposed by any osmotic force and will drive water from the right‐flowing to the left‐flowing stream at a rate that depends on the permeability of the membrane, thus concentrating the right‐flowing and diluting the left‐flowing stream.

Adapted from Hargitay and Kuhn 65
Figure 10. Figure 10.

Hargitay and Kuhn multiplier with withdrawal. A fraction fu of the combined descending limb (DL) and collecting duct (CD) flow is withdrawn at the loop. Mass balance now requires that C1 = C3C2. This creates an osmotic force opposing the single effect and reduces the concentration. The degree of the reduction depends on the ratio of the single effect to the withdrawal. AL, ascending limb.

Figure 11. Figure 11.

Comparison of classic representation of concentration build‐up in solute cycling multiplier (A) with calculated profiles (B). Multiplication by counterflow is apparent, but the concentration difference between counterflowing limbs varies with both time and position. “Pump” and “leak” were adjusted to give a limiting “single effect” of 200 mOsm/liter.

Reprinted with permission from Garner et al. 47, Bulletin of Mathematical Biology, Vol. 40, “Transient behaviour of the single loop solute cycling model of the renal medulla.” © 1978, Pergamon Press
Figure 12. Figure 12.

Relative osmolalities of tubular fluid in slices from kidneys of hydropenic rats. Values are given as percent of maximum. O.Z, outer zone; I.Z, inner zone (as modified for ref. 111 from ref. 260).

From Knepper and Stephenson 111
Figure 13. Figure 13.

Composition of renal medulla and urine during antidiuresis and water diuresis (adapted from Hai and Thomas 63 for 20.

From Knepper and Stephenson 111
Figure 14. Figure 14.

Comparison of osmolalities of fluids collected by micropuncture near the tip of the inner medulla in nine hamsters, one kangaroo rat, and one Psammomys.

From Gottschalk and Mylle 58
Figure 15. Figure 15.

Osmolality ratio (tubule fluid/plasma) of fluid collected by micropuncture from seven rats during antidiuresis. Early tubule fluid was dilute even though urine was concentrated. Different symbols refer to different rats.

From Gottschalk and Mylle 58
Figure 16. Figure 16.

Countercurrent exchanger consisting of ascending vas rectum (AVR) and descending vas rectum (DVR). Solute is added to surrounding interstitium at rate J per unit length. Solute entering AVR from interstitium is recycled diffusively to DVR, with resulting build‐up of concentration at the loop.

Figure 17. Figure 17.

Diagram of two‐loop counterflow system with parallel Henle's loop and vas rectum.

Reprinted by permission from Stephenson 190, Nature, Vol. 206, pp. 1215–1219. © 1965 Macmillan Magazines Ltd
Figure 18. Figure 18.

A model of a vasa recta bundle. Direction of flow is indicated by arrows.

From Marsh and Segel 138
Figure 19. Figure 19.

Schematic representation of the intricate counterflow system of the renal medulla. Note that in this figure AVR denotes arterial vas rectum (or descending vas rectum; VVR, venous vas rectum (or ascending vas rectum); DL, descending limb of Henle's loop; AL, ascending limb of Henle's loop; CD, collecting duct; CAP, capillary.

From Kriz and Lever 119
Figure 20. Figure 20.

a: Central core model of the renal medulla. Black arrows indicate salt movement; open arrows, water movement; hatched arrows, urea movement. With sufficiently large solute and water permeabilities of ascending vasa recta (AVR) and descending vasa recta (DVR), concentrations in AVR, DVR, and interstitium become nearly identical at a given medullary level. Solute removal is then given by the product of the concentration and the volume flow difference in AVR and DVR. Functionally, AVR, DVR, and interstitium are merged into a single tube closed at the papillary tip and open at the junction of medulla and cortex. CD, collecting duct; AHL, ascending limb of Henle's loop; DHL, descending limb of Henle's loop. b: shows various core components being hypothetically merged to give the final cross‐sectional configuration shown in c.

Reprinted from Stephenson 193, Kidney International, with permission
Figure 21. Figure 21.

Skeletonized central core model. Black arrows indicate solute movement; white arrows, volume flow. AHL, ascending limb; DHL, descending limb, CD, collecting duct.

From Stephenson 206
Figure 22. Figure 22.

Single osmotic effect in two modes of operation. Mode I, active transport (upper titles); mode II, mixing (bracketed); initial (a) and final (b) equilibrium states are shown. Open arrow indicates water movement; closed arrow, salt movement.

Reprinted from Stephenson 193, Kidney International, with permission
Figure 23. Figure 23.

Configuration for Peskin‐Layton single nephron model. By assumption, C1(γ) = C3(γ) = C(γ) = J2s/Jv.

From Layton 124
Figure 24. Figure 24.

Structure of model 1 of Wexler, Kalaba, and Marsh. Top: cross‐section showing connections between tubules, blood vessels, and interstitial space. Bottom: vertical aspect. DVR, descending vasa recta; AVR, ascending vasa recta; DHL, descending limb of Henle's loop; AHL, ascending limb of Henle's loop; DT, distal tubule; CD, collecting duct; I, interstitium. OM, outer medulla; IM, inner medulla.

From Wexler et al. 252
Figure 25. Figure 25.

Structure of model 3 of Wexler, Kalaba, and Marsh. Top: cross‐section showing connections between tubules, blood vessels, and interstitial space. Bottom: vertical aspect. DVR, descending vasa recta; AVR, ascending vasa recta; DHL, descending limb of Henle's loop; AHL, ascending limb of Henle's loop; DT, distal tubule; CD, collecting duct; I, interstitium; OM, outer medulla; IM, inner medulla.

From Wexler et al. 252
Figure 26. Figure 26.

Interstitial concentrations of sodium chloride and urea predicted by model 3 of Wexler, Kalaba, and Marsh.

From Wexler et al. 252
Figure 27. Figure 27.

Configuration of two‐nephron, central core, electrolyte model of renal medulla of Stephenson, Zhang, and Tewarson. DHL, descending limb of Henle's loop; AHL, ascending limb of Henle's loop; CD, collecting duct; DN, distal nephron.

From Stephenson et al. 213
Figure 28. Figure 28.

Core osmolalities for various modifications of rabbit and hamster data. Top: A, unmodified; B, urea permeability of descending limb reduced; C, urea permeability of ascending limb reduced; D, urea permeability of both descending and ascending limbs reduced. Bottom: E, unmodified; F, urea permeability of both descending and ascending limbs reduced; G, urea permeabilities as in F, descending limb electrolyte permeabilities also reduced.

From Stephenson et al. 213
Figure 29. Figure 29.

Effect of varying urea permeability of thin descending limb of Henle's loop on core osmolality at papilla and at junction of inner and outer medullae.

From Stephenson et al. 213
Figure 30. Figure 30.

Model of Chandhoke, Saidel, and Knepper. Renal cortex is divided into a medullary ray and a labyrinth. Renal arterial blood (afferents) flows to four groups of glomeruli in cortical labyrinth: those of superficial nephrons (G1), midcortical short nephrons (G2), midcortical long nephrons (G3), and juxtamedullary nephrons (G4). G1 and G2 give rise to all short nephrons (SN), whereas G3 and G4 yield all long nephrons (LN). Efferent blood from G1, G2, and G3 flows to capillary plexuses of both labyrinth and medullary ray. Capillary blood of medullary ray drains into labyrinth. Efferent blood from G4 that flows to descending vasa recta (DVR) is sole blood supply to medulla. Proximal convoluted tubules of short nephrons (PCTSN) give rise to their straight counterparts (PCTSN) found within the medullary ray. However, proximal straight tubules of long nephrons (PSTLN) enter medulla without traversing medullary ray. Transition of proximal straight tubules to descending limbs of Henle (DLSN and DLLN, respectively) occurs at outer stripe‐inner stripe border. All DLSN turn to form medullary thick ascending limbs of short nephrons (MALSN) at inner stripe‐inner zone border; descending limbs of long nephrons (DLLN) turn at various depths along inner zone to form thin ascending limbs of Henle (TNAL.) At inner‐outer zone border, thin ascending limbs become medullary thick ascending limbs of long nephrons (MALLN). In medullary ray, medullary thick ascending limbs of short nephrons become cortical thick ascending of short nephrons (CALSN), which lead to distal convoluted tubules of short nephrons (DTSN) found in labyrinth. Flow from MALLN is assumed to go directly into distal convoluted tubules of long nephrons (DTLN). DTLN fluid feeds into connecting tubules (CT) that rise as arcades within labyrinth and, together with fluid delivered from DTSN, empties into cortical collecting ducts (CCD) of medullary ray. In the medulla, collecting ducts become outer medullary ducts (OMCD) and then inner medullary collecting ducts (IMCD). IMCD are assumed to have a converging structure. Blood from descending vasa recta (DVR) enters medullary capillary plexus and then flows into ascending vasa recta (AVR). At cortico‐medullary junction, the blood in AVR constitutes the medullary venous return, and the blood from cortical labyrinth capillary plexus, the cortical venous return.

From Chandhoke et al. 26
Figure 31. Figure 31.

Simulated tubule fluid/plasma (TF/P) sodium chloride in fluid of the medullary collecting ducts for no active sodium chloride transport (SO), uniform active sodium chloride transport (SU), and nonuniform active sodium chloride transport (SN).

From Chandhoke et al. 26
Figure 32. Figure 32.

Simulated tubule fluid/plasma (TF/P) urea in fluid along the medullary collecting ducts for different cases of collecting duct active sodium chloride transport: none (SO), uniform (SU), and nonuniform (SN).

From Chandhoke et al. 26
Figure 33. Figure 33.

Simulated tubule fluid/plasma (TF/P) osmolarity in fluid along the medullary collecting ducts for different cases of collecting duct active sodium chloride transport: none (SO), uniform (SU), and non‐uniform (SN).

From Chandhoke et al. 26
Figure 34. Figure 34.

Concentration ratio of three‐tube model, consisting of descending thin limb coupled to ascending and descending vasa recta, as a function of exchange efficiency K, for both an approximate analytical solution (—) and a numerical solution (•) 211.

Figure 35. Figure 35.

Effect of vasa recta plasma flow rate on solute accumulation in inner medulla. Fraction of total solute load entering medulla via descending limb lost through solute washout arising from imperfect countercurrent exchange is given in parentheses.

From Moore and Marsh 147
Figure 36. Figure 36.

Time course of distal tubule and collecting duct water permeability changes used in model of Moore, Marsh, and Martin to cause transition from steady‐state diuresis to antidiuresis. Initial 30 min were used to allow system to achieve a steady state. DT, distal tubule; CD, collecting duct

From Moore et al. 149
Figure 37. Figure 37.

Predicted change in solute concentrations in descending limb (DLH) tubular fluid following antidiuretic hormone. Zero time corresponds to 30 min in Figure 36. Concentrations expressed as fractional increase in concentration difference. Left: predicted sodium chloride concentrations. Right: predicted urea concentrations.

From Moore et al. 149


Figure 1.

Response of whole kidney to varying rates of vasopressin infusion in rats. Note the dramatic decrease in urine flow rate at higher rates of vasopressin infusion with minimal change in osmolar clearance

Data from Atherton et al. 3. From Knepper and Stephenson 111


Figure 2.

Relation between relative medullary thickness and maximum urinary osmolality. Long loops of Henle are required to make highly concentrated urine

Data from Schmidt‐Nielsen and O'Dell 179. From Burg and Stephenson 20


Figure 3.

Prototype three‐section glomerular nephron. With its straight configuration, solute and water transport from different segments is uncoupled. In the proximal tubule (PT) with a large effective hydraulic permeability, Lp, absorption is nearly isotonic; in the diluting segment (DIL) a smaller Lp limits water absorption relative to solute absorption, and tubule fluid is diluted; in final segment, hydraulic permeability is under control of antidiuretic hormone (ADH). In the absence of ADH tubule fluid remains dilute, and in the presence of ADH it equilibrates osmotically with the surrounding interstitium. Thus final urine can vary from markedly hypotonic to isotonic relative to plasma. Black arrows indicate solute movement; white arrows, volume flow.

Reproduced, with permission, from Stephenson 201, Annual Review of Biophysics and Bioengineering, Vol. 7. © 1978 by Annual Reviews, Inc


Figure 4.

The S‐shaped mammalian nephron. The essential functional feature that allows concentration of urine is juxtaposition of the diluting segment (DIL; now ascending limb of Henle's loop) with the water‐permeable descending limb and antidiuretic hormone (ADH)‐sensitive collecting duct. This permits solute supplied by ascending limb to extract water from collecting duct and descending limb. Black arrows indicate solute movement; white arrows, volume flow; PT, proximal tubule.

Reproduced, with permission, from Stephenson 201, Annual Review of Biophysics and Bioengineering, Vol. 7. © 1978 by Annual Reviews, Inc


Figure 5.

Initial stage of basic scheme for concentrating by permselective membranes. A very large volume of phenol solution at concentration Co is separated by membrane P permeable to phenol but not to water (rubber in experiments of Kuhn and Ryffel) from a large volume of sucrose solution, also at concentration Co. The sucrose solution is separated by membrane W permeable to water but not to sucrose or to phenol (copper ferrocyanide) from smaller volumes of phenol and sucrose solutions, also at concentration Co. As indicated by arrows, phenol diffuses into sucrose solution, increasing its osmolality and causing water to be osmotically extracted from the smaller volume solutions.

Adapted from Kuhn and Ryffel 122


Figure 6.

Basic scheme for concentrating by permselective membranes at equilibrium. Phenol has been added to the original sucrose solution so that its concentration has approximately doubled. Water has been extracted from the smaller volumes of sucrose and phenol so that their volumes are halved and concentrations doubled.

Adapted from Kuhn and Ryffel 122


Figure 7.

Arrangement for parallel processing of counterflowing sucrose and phenol solutions to yield progressive concentration of the sucrose solution. In the first apparatus on the left, sucrose and phenol, both at concentration Co, interact through a phenol‐permeable membrane P to give a combined solution of concentration 2Co. This interacts with counterflowing sucrose solution in the first apparatus on the right through a water‐permeable membrane W to give a sucrose solution of concentration 2Co. The phenol is then removed by diffusion into a concurrent flow of any solution at concentration Co through a second membrane P. As the solution is processed from left to right, there is progressive concentration of the sucrose solution. In further stages to the right, phenol reservoirs are not shown in order to simplify the diagram. Kuhn and Ryffel suggested that this scheme might mimic the operation of the kidney.

Adapted from Kuhn and Ryffel 122


Figure 8.

Actual scheme used by Kuhn and Ryffel. See text for explanation.

Adapted from Kuhn and Ryffel 122


Figure 9.

Hargitay and Kuhn countercurrent multiplier without withdrawal. Solution at concentration C1 enters at the left and reverses flow through hairpin constriction at the right to return to the left. The two counterflowing solutions are separated by a waterpermeable and solute‐impermeable membrane. Mass balance requires that at any position x along the tube C1 = C2 and F1 = ‐ F2. This means that the pressure difference Δp, the single effect, is unapposed by any osmotic force and will drive water from the right‐flowing to the left‐flowing stream at a rate that depends on the permeability of the membrane, thus concentrating the right‐flowing and diluting the left‐flowing stream.

Adapted from Hargitay and Kuhn 65


Figure 10.

Hargitay and Kuhn multiplier with withdrawal. A fraction fu of the combined descending limb (DL) and collecting duct (CD) flow is withdrawn at the loop. Mass balance now requires that C1 = C3C2. This creates an osmotic force opposing the single effect and reduces the concentration. The degree of the reduction depends on the ratio of the single effect to the withdrawal. AL, ascending limb.



Figure 11.

Comparison of classic representation of concentration build‐up in solute cycling multiplier (A) with calculated profiles (B). Multiplication by counterflow is apparent, but the concentration difference between counterflowing limbs varies with both time and position. “Pump” and “leak” were adjusted to give a limiting “single effect” of 200 mOsm/liter.

Reprinted with permission from Garner et al. 47, Bulletin of Mathematical Biology, Vol. 40, “Transient behaviour of the single loop solute cycling model of the renal medulla.” © 1978, Pergamon Press


Figure 12.

Relative osmolalities of tubular fluid in slices from kidneys of hydropenic rats. Values are given as percent of maximum. O.Z, outer zone; I.Z, inner zone (as modified for ref. 111 from ref. 260).

From Knepper and Stephenson 111


Figure 13.

Composition of renal medulla and urine during antidiuresis and water diuresis (adapted from Hai and Thomas 63 for 20.

From Knepper and Stephenson 111


Figure 14.

Comparison of osmolalities of fluids collected by micropuncture near the tip of the inner medulla in nine hamsters, one kangaroo rat, and one Psammomys.

From Gottschalk and Mylle 58


Figure 15.

Osmolality ratio (tubule fluid/plasma) of fluid collected by micropuncture from seven rats during antidiuresis. Early tubule fluid was dilute even though urine was concentrated. Different symbols refer to different rats.

From Gottschalk and Mylle 58


Figure 16.

Countercurrent exchanger consisting of ascending vas rectum (AVR) and descending vas rectum (DVR). Solute is added to surrounding interstitium at rate J per unit length. Solute entering AVR from interstitium is recycled diffusively to DVR, with resulting build‐up of concentration at the loop.



Figure 17.

Diagram of two‐loop counterflow system with parallel Henle's loop and vas rectum.

Reprinted by permission from Stephenson 190, Nature, Vol. 206, pp. 1215–1219. © 1965 Macmillan Magazines Ltd


Figure 18.

A model of a vasa recta bundle. Direction of flow is indicated by arrows.

From Marsh and Segel 138


Figure 19.

Schematic representation of the intricate counterflow system of the renal medulla. Note that in this figure AVR denotes arterial vas rectum (or descending vas rectum; VVR, venous vas rectum (or ascending vas rectum); DL, descending limb of Henle's loop; AL, ascending limb of Henle's loop; CD, collecting duct; CAP, capillary.

From Kriz and Lever 119


Figure 20.

a: Central core model of the renal medulla. Black arrows indicate salt movement; open arrows, water movement; hatched arrows, urea movement. With sufficiently large solute and water permeabilities of ascending vasa recta (AVR) and descending vasa recta (DVR), concentrations in AVR, DVR, and interstitium become nearly identical at a given medullary level. Solute removal is then given by the product of the concentration and the volume flow difference in AVR and DVR. Functionally, AVR, DVR, and interstitium are merged into a single tube closed at the papillary tip and open at the junction of medulla and cortex. CD, collecting duct; AHL, ascending limb of Henle's loop; DHL, descending limb of Henle's loop. b: shows various core components being hypothetically merged to give the final cross‐sectional configuration shown in c.

Reprinted from Stephenson 193, Kidney International, with permission


Figure 21.

Skeletonized central core model. Black arrows indicate solute movement; white arrows, volume flow. AHL, ascending limb; DHL, descending limb, CD, collecting duct.

From Stephenson 206


Figure 22.

Single osmotic effect in two modes of operation. Mode I, active transport (upper titles); mode II, mixing (bracketed); initial (a) and final (b) equilibrium states are shown. Open arrow indicates water movement; closed arrow, salt movement.

Reprinted from Stephenson 193, Kidney International, with permission


Figure 23.

Configuration for Peskin‐Layton single nephron model. By assumption, C1(γ) = C3(γ) = C(γ) = J2s/Jv.

From Layton 124


Figure 24.

Structure of model 1 of Wexler, Kalaba, and Marsh. Top: cross‐section showing connections between tubules, blood vessels, and interstitial space. Bottom: vertical aspect. DVR, descending vasa recta; AVR, ascending vasa recta; DHL, descending limb of Henle's loop; AHL, ascending limb of Henle's loop; DT, distal tubule; CD, collecting duct; I, interstitium. OM, outer medulla; IM, inner medulla.

From Wexler et al. 252


Figure 25.

Structure of model 3 of Wexler, Kalaba, and Marsh. Top: cross‐section showing connections between tubules, blood vessels, and interstitial space. Bottom: vertical aspect. DVR, descending vasa recta; AVR, ascending vasa recta; DHL, descending limb of Henle's loop; AHL, ascending limb of Henle's loop; DT, distal tubule; CD, collecting duct; I, interstitium; OM, outer medulla; IM, inner medulla.

From Wexler et al. 252


Figure 26.

Interstitial concentrations of sodium chloride and urea predicted by model 3 of Wexler, Kalaba, and Marsh.

From Wexler et al. 252


Figure 27.

Configuration of two‐nephron, central core, electrolyte model of renal medulla of Stephenson, Zhang, and Tewarson. DHL, descending limb of Henle's loop; AHL, ascending limb of Henle's loop; CD, collecting duct; DN, distal nephron.

From Stephenson et al. 213


Figure 28.

Core osmolalities for various modifications of rabbit and hamster data. Top: A, unmodified; B, urea permeability of descending limb reduced; C, urea permeability of ascending limb reduced; D, urea permeability of both descending and ascending limbs reduced. Bottom: E, unmodified; F, urea permeability of both descending and ascending limbs reduced; G, urea permeabilities as in F, descending limb electrolyte permeabilities also reduced.

From Stephenson et al. 213


Figure 29.

Effect of varying urea permeability of thin descending limb of Henle's loop on core osmolality at papilla and at junction of inner and outer medullae.

From Stephenson et al. 213


Figure 30.

Model of Chandhoke, Saidel, and Knepper. Renal cortex is divided into a medullary ray and a labyrinth. Renal arterial blood (afferents) flows to four groups of glomeruli in cortical labyrinth: those of superficial nephrons (G1), midcortical short nephrons (G2), midcortical long nephrons (G3), and juxtamedullary nephrons (G4). G1 and G2 give rise to all short nephrons (SN), whereas G3 and G4 yield all long nephrons (LN). Efferent blood from G1, G2, and G3 flows to capillary plexuses of both labyrinth and medullary ray. Capillary blood of medullary ray drains into labyrinth. Efferent blood from G4 that flows to descending vasa recta (DVR) is sole blood supply to medulla. Proximal convoluted tubules of short nephrons (PCTSN) give rise to their straight counterparts (PCTSN) found within the medullary ray. However, proximal straight tubules of long nephrons (PSTLN) enter medulla without traversing medullary ray. Transition of proximal straight tubules to descending limbs of Henle (DLSN and DLLN, respectively) occurs at outer stripe‐inner stripe border. All DLSN turn to form medullary thick ascending limbs of short nephrons (MALSN) at inner stripe‐inner zone border; descending limbs of long nephrons (DLLN) turn at various depths along inner zone to form thin ascending limbs of Henle (TNAL.) At inner‐outer zone border, thin ascending limbs become medullary thick ascending limbs of long nephrons (MALLN). In medullary ray, medullary thick ascending limbs of short nephrons become cortical thick ascending of short nephrons (CALSN), which lead to distal convoluted tubules of short nephrons (DTSN) found in labyrinth. Flow from MALLN is assumed to go directly into distal convoluted tubules of long nephrons (DTLN). DTLN fluid feeds into connecting tubules (CT) that rise as arcades within labyrinth and, together with fluid delivered from DTSN, empties into cortical collecting ducts (CCD) of medullary ray. In the medulla, collecting ducts become outer medullary ducts (OMCD) and then inner medullary collecting ducts (IMCD). IMCD are assumed to have a converging structure. Blood from descending vasa recta (DVR) enters medullary capillary plexus and then flows into ascending vasa recta (AVR). At cortico‐medullary junction, the blood in AVR constitutes the medullary venous return, and the blood from cortical labyrinth capillary plexus, the cortical venous return.

From Chandhoke et al. 26


Figure 31.

Simulated tubule fluid/plasma (TF/P) sodium chloride in fluid of the medullary collecting ducts for no active sodium chloride transport (SO), uniform active sodium chloride transport (SU), and nonuniform active sodium chloride transport (SN).

From Chandhoke et al. 26


Figure 32.

Simulated tubule fluid/plasma (TF/P) urea in fluid along the medullary collecting ducts for different cases of collecting duct active sodium chloride transport: none (SO), uniform (SU), and nonuniform (SN).

From Chandhoke et al. 26


Figure 33.

Simulated tubule fluid/plasma (TF/P) osmolarity in fluid along the medullary collecting ducts for different cases of collecting duct active sodium chloride transport: none (SO), uniform (SU), and non‐uniform (SN).

From Chandhoke et al. 26


Figure 34.

Concentration ratio of three‐tube model, consisting of descending thin limb coupled to ascending and descending vasa recta, as a function of exchange efficiency K, for both an approximate analytical solution (—) and a numerical solution (•) 211.



Figure 35.

Effect of vasa recta plasma flow rate on solute accumulation in inner medulla. Fraction of total solute load entering medulla via descending limb lost through solute washout arising from imperfect countercurrent exchange is given in parentheses.

From Moore and Marsh 147


Figure 36.

Time course of distal tubule and collecting duct water permeability changes used in model of Moore, Marsh, and Martin to cause transition from steady‐state diuresis to antidiuresis. Initial 30 min were used to allow system to achieve a steady state. DT, distal tubule; CD, collecting duct

From Moore et al. 149


Figure 37.

Predicted change in solute concentrations in descending limb (DLH) tubular fluid following antidiuretic hormone. Zero time corresponds to 30 min in Figure 36. Concentrations expressed as fractional increase in concentration difference. Left: predicted sodium chloride concentrations. Right: predicted urea concentrations.

From Moore et al. 149
References
 1. Ang, P. G. P., H. D. Landahl, and E. Bartoli. Transient and steady state simulation of the renal countercurrent mechanism. Comput. Biol. Med. 7: 87‐111, 1977.
 2. Atherton, J. C. Lability of renal papillary tissue composition in the rat. J. Physiol. (Lond.) 274: 323‐328, 1978.
 3. Atherton, J. C., R. Green, and S. Thomas. Influence of lysine‐vasopressin dosage on the time course of changes in renal tissue and urinary composition in the conscious rat. J. Physiol. (Lond.) 213: 291‐309, 1971.
 4. Bankir, L., and C. de Rouffignac. Urinary concentrating ability: insights from comparative anatomy. Am. J. Physiol. 249 (Regulatory Integrative Comp. Physiol. 18): R643‐R666, 1985.
 5. Barfuss, D. W., and J. A. Schafer. Differences in active and passive glucose transport along the proximal nephron. Am. J. Physiol. 240 (Renal Fluid Electrolyte Physiol. 9): F322‐F332, 1981.
 6. Bargman, J., S. L. Leonard, E. McNeely, C. Robertson, and R. L. Jamison. Examination of transepithelial exchange of water and solute in the rat renal pelvis. J. Clin. Invest. 74: 1860‐1870, 1984.
 7. Barrett, G. L., and J. S. Packer. Dynamic simulation of the renal medulla. Med. Biol. Eng. Comput. 21: 324‐332, 1983.
 8. Barrett, G. L., J. S. Packer, and J. M. Davis. Sodium chloride, water and urea handling in the rat renal medulla: a computer simulation. Renal Physiol. 9: 223‐240, 1986.
 9. Bassingthwaite, J. B., and C. A. Goresky. Modeling in the analysis of solute and water exchange in the microvascularure. In: Handbook of Physiology. The Cardiovascular System. Microcirculation, edited by E. M. Renkin and C. C. Michel. Bethesda, MD: Am. Physiol. Soc., 1984, sect. 2, vol. IV, pt. 1, chap. 13, p. 549‐626.
 10. Bellman, R. E., and R. E. Kalaba. Quasilinearization and Nonlinear Value Boundary Problems. New York: Elsevier, 1965.
 11. Berliner, R. W. Formation of concentrated urine. In: Renal Physiology: People and Ideas, edited by R. W. Berliner, G. Giebisch, and C. W. Gottschalk. Washington, DC: Am. Physiol Soc., 1987, p. 247‐276.
 12. Berliner, R. W., N. G. Levinsky, D. G. Davidson, and M. Eden. Dilution and concentration of the urine and the action of antidiuretic hormone. Am. J. Med. 24: 730‐744, 1958.
 13. Bonventre, J. V., and C. Lechene. Renal medullary concentrating process: an integrative hypothesis. Am. J. Physiol. 239 (Renal Fluid Electrolyte Physiol. 8): F578‐F588, 1980.
 14. Bonventre, J. V., R. J. Roman, and C. Lechene. Effect of urea concentration of pelvic fluid on renal concentrating ability. Am. J. Physiol. 239 (Renal Fluid Electrolyte Physiol. 8): F609‐F618, 1980.
 15. Brodsky, W. A., W. S. Rehm, and W. H. Dennis. Osmotic gradients across cellular membranes. Science 124: 221‐222, 1956.
 16. Brodsky, W. A., W. S. Rehm, W. H. Dennis, and D. G. Miller. Thermodynamic analysis of the intracellular osmotic gradient hypothesis of active water support. Science 121: 302‐303, 1955.
 17. Broyden, C. G. A class of methods for solving nonlinear simultaneous equations. Math. Comput. 19: 577‐593, 1965.
 18. Burg, M. B. Thick ascending limb of Henle's loop. Kidney Int. 22: 454‐464, 1982.
 19. Burg, M. B., and N. Green. Function of the thick ascending limb of Henle's loop. Am. J. Physiol. 224: 659‐668, 1973.
 20. Burg, M. B., and J. L. Stephenson. Transport characteristics of the loop of Henle. In: Physiology of Membrane Disorders, edited by T. E. Andreoli, J. F. Hoffman, and D. D. Fanestil. New York: Plenum, 1978, p. 661‐679.
 21. Burgen, A. S. V. A theoretical treatment of glucose reabsorption in the kidney. Can. J. Biochem. Physiol. 34: 466‐474, 1956.
 22. Cage, P. E., E. R. Carson, and K. E. Britton. A model of the human renal medulla. Comput. Biomed. Res. 10: 561‐584, 1977.
 23. Capek, K., G. Fuchs, G. Rumrich, and K. J. Ullrich. Harnstoffpermeabilität der corticalen Tubulusabschnitte von Ratten in Antidiurese und Wasserdiurese. Pflugers Arch. 290: 237‐249, 1966.
 24. Caplan, S. R., and A. Essig. Bioenergetics and Linear Nonequilibrium Thermodynamics—The Steady State. London: Harvard University Press, 1983.
 25. Chandhoke, P. S., and G. M. Saidel. Mathematical model of mass transport throughout the kidney: effects of nephron heterogeneity and tubular‐vascular organization. Ann. Biomed. Eng. 9: 263‐301, 1981.
 26. Chandhoke, P. S., G. M. Saidel, and M. A. Knepper. Role of inner medullary collecting duct NaCl transport in urinary concentration. Am. J. Physiol. 249 (Renal Fluid Electrolyte Physiol. 18): F688‐F697, 1985.
 27. Deen, W. M., C. R. Robertson, and B. M. Brenner. A model of glomerular ultrafiltration in the rat. Am. J. Physiol. 223: 1178‐1183, 1972.
 28. Deen, W. M., C. R. Robertson, and B. M. Brenner. A model of peritubular capillary control of isotonic fluid reabsorption by the renal proximal tubule. Biophys. J. 13: 340‐358, 1973.
 29. Dennis, J. E., and R. B. Schnabel. Numerical Methods for Unconstrained Optimization and Nonlinear Equations. Englewood Cliffs, NJ: Prentice Hall Inc., 1983.
 30. de Rouffignac, C, and F. Morel. Micropuncture study of water, electrolytes, and urea movements along the loops of Henle in Psammomys. J. Clin. Invest. 48: 474‐486, 1969.
 31. Diamond, J. M., and Bossert, W. H. Standing‐gradient osmotic flow: a mechanism for coupling of water and solute transport in epithelia. J. Gen. Physiol. 50: 2061‐2083, 1967.
 32. Dole, V. P. Back‐diffusion of urea in the mammalian kidney. Am. J. Physiol. 139: 504‐513, 1943.
 33. Dongarra, J. J. Performance of various computers using standard linear equations software in a Fortran environment. Argonne, IL: Mathematics and Computer Science Division, Argonne National Laboratory, Technical Memorandum No. 23, August 1984.
 34. Du Bois, R., A. Verniory, and M. Abramow. Computation of the osmotic water permeability of perfused tubule segments. Kidney Int. 10: 478‐479, 1976.
 35. Eason, G. The central core model of the renal medulla: an approximate solution. Math. Biosci. 79: 107‐116, 1986.
 36. Eason, G. On the metabolic pump in the renal medulla. Math. Biosci. 84: 155‐158, 1987.
 37. Eason, G. The central core model of the renal medulla: a particular solution. Math. Biosci. 85: 1‐12, 1987.
 38. Farahzad, P. Analysis of the equations of renal network flows. Math. Biosci. 40: 233‐242, 1978.
 39. Farahzad, P., and R. P. Tewarson. An efficient numerical method for solving the differential equations of renal counterflow systems. Comput. Biol. Med. 8: 57‐64, 1978.
 40. Farahzad, P., and R. P. Tewarson. Numerical continuation method for a system of parallel flow tubes. Comput. Biol. Med. 9: 21‐27, 1979.
 41. Foster, D. M., and J. A. Jacquez. Comparison using central core model of renal medulla of the rabbit and rat. Am. J. Physiol. 234 (Renal Fluid Electrolyte Physiol. 7): F402‐F414, 1978.
 42. Foster, D., A. Jacquez, and E. Daniels. Solute concentration in the kidney—II: input‐output studies on a central core model. Math. Biosci. 32: 337‐360, 1976.
 43. Franck, J., and J. E. Mayer. An osmotic diffusion pump. Arch. Biochem. 14: 297‐313, 1947.
 44. Franke, H., W. Niesel, and H. Röskenbleck. Differenzierung der Konzentrierungsprozesse des äusseren und des inneren Marks durch Ermittelung der Konzentrationsprofile von Natrium, Kalium und Harnstoff bei unterschiedlichen Funktionszuständen. Pflugers Arch. 315: 321‐335, 1970.
 45. Friedlander, S. K., and M. Walser. Some aspects of flow and diffusion in the proximal tubule of the kidney. J. Theor. Biol. 8: 87‐96, 1965.
 46. Furukawa, T., S. Takasugi, M. Inoue, H. Inada, F. Kajiya, and H. Abe. A digital computer model of the renal medullary countercurrent system. Comput. Biomed. Res. 7: 213‐229, 1974.
 47. Garner, J. B., K. S. Crump, and J. L. Stephenson. Transient behaviour of the single loop solute cycling model of the renal medulla. Bull. Math. Biol. 40: 273‐300, 1978.
 48. Garner, J. B., and R. B. Kellogg. A one tube flow problem arising in physiology. Bull. Math. Biol. 42: 295‐304, 1980.
 49. Garner, J. B., and R. B. Kellogg. The diffusion‐convection equation with pressure. J. Math. Anal. Appl. 79: 58‐70, 1981.
 50. Garner, J. B., and R. B. Kellogg. Diffusion and convection in a family of tubes. J. Math. Anal. Appl. 85: 461‐472, 1982.
 51. Garner, J. B., R. B. Kellogg, and J. L. Stephenson. Mathematical analysis of a model for the renal concentrating mechanism. Math. Biosci. 65: 125‐150, 1983.
 52. Gertz, K. H. Transtubuläre Natriumchloridflüsse und Permeabilität für Nichtelektrolyte im proximalen und distalen Konvolut der Rattenniere. Pflugers Arch. 276: 336‐356, 1963.
 53. Gertz, K. H., M. Brandis, G. Braun‐Schubert, and J. W. Boylan. The effect of saline infusion and hemorrhage on glomerular filtration pressure and single nephron filtration rate. Pflugers Arch. 310: 193‐205, 1969.
 54. Gertz, K. H., J. A. Mangos, G. Braun, and H. D. Pagel. On the glomerular tubular balance in the rat kidney. Pflugers Arch. 285: 360‐372, 1965.
 55. Gertz, K. H., B. Schmidt‐Nielsen, and D. Pagel. Exchange of water, urea and salt between the mammalian renal papilla and the surrounding urine. Federation Proc. 25: 327, 1966.
 56. Gottschalk, C. W. Osmotic concentration and dilution of the urine. Am. J. Med. 36: 670‐685, 1964.
 57. Gottschalk, C. W., W. E. Lassiter, M. Mylle, K. J. Ullrich, B. Schmidt‐Nielsen, R. O'Dell, and G. Pehling. Micropuncture study of composition of loop of Henle fluid in desert rodents. Am. J. Physiol. 204: 532‐535, 1963.
 58. Gottschalk, C. W., and M. Mylle. Micropuncture study of the mammalian urinary concentrating mechanism: evidence for the countercurrent hypothesis. Am. J. Physiol. 196: 927‐936, 1959.
 59. Greenwald, L. The significance of renal relative medullary thickness. Physiol. Zool. 62: 1005‐1014, 1989.
 60. Greenwald, L., and D. Stetson. Urine concentration and the length of the renal papilla. NIPS 3: 46‐49, 1988.
 61. Greger, R. Ion transport mechanisms in thick ascending limb of Henle's loop of mammalian nephron. Physiol. Rev. 65: 760‐797, 1985.
 62. Hagstrom, T., and R. P. Tewarson. Partioning and parallel algorithms for kidney models. Math. Mod. 11: 847‐849, 1988.
 63. Hai, M. A., and S. Thomas. Influence of prehydration on the changes in renal tissue composition induced by water diuresis in the rat. J. Physiol. (Land.) 205: 599‐618, 1969.
 64. Handler, J. S., and J. Orloff. Antidiuretic hormone. Annu. Rev. Physiol. 43: 611‐624, 1981.
 65. Hargitay, B., and W. Kuhn. Das Multiplikationsprinzip als Grundlage der Harnkonzentrierung in der Niere. Z. Elektrochem. Angew. Phys. Chem. 55: 539‐558, 1951.
 66. Hearon, J. Z. The steady state kinetics of some biological systems: IV. Thermodynamic aspects. Bull Math. Biophys. 12: 85‐106, 1950.
 67. Heinz, E. Electrical Potentials in Biological Membrane Transport. New York: Springer‐Verlag, 1981.
 68. Homer, L. D., and P. K. Weathersby. Transient solutions of equations for countercurrent capillary exchange. Am. J. Physiol. 245 (Regulatory Integrative Comp. Physiol. 14): R534‐R540, 1983.
 69. Horster, M. F., A. Gilg, and P. Lory. Determinants of axial osmotic gradients in the differentiating countercurrent system. Am. J. Physiol. 246 (Renal Fluid Electrolyte Physiol. 15): F124‐F132, 1984.
 70. Hubbard, B. E. Computing transient solutions for certain renal counterflow systems. In: Proc. Summer Comput. Simulation Conf., Washington, DC, 1976. Lajolla, CA: Simulation Councils, Inc., 1976.
 71. Huss, R. E., and J. L. Stephenson. A mathematical model of proximal tubule absorption. J. Membr. Biol. 47: 377‐399, 1979.
 72. Imai, M. Function of the thin ascending limb of Henle of rats and hamsters perfused in vitro. Am. J. Physiol. 232 (Renal Fluid Electrolyte Physiol. 1): F201‐F209, 1977.
 73. Imai, M. Functional heterogeneity of the descending limbs of Henle's loop. II. Interspecies differences among rabbits, rats, and hamsters. Pflugers Arch. 402: 393‐401, 1984.
 74. Imai, M., M. Hayashi, and M. Araki. Functional heterogeneity of the descending limbs of Henle's loop. I. Internephron heterogeneity in the hamster kidney. Pflugers Arch. 402: 385‐392, 1984.
 75. Imai, M., and J. P. Kokko. Sodium chloride, urea, and water transport in the thin ascending limb of Henle: generation of osmotic gradients by passive diffusion of solutes. J. Clin. Invest. 53: 393‐402, 1974.
 76. Imai, M., J. Taniguchi, and K. Tabei. Function of thin loops of Henle. Kidney Int. 31: 565‐579, 1987.
 77. Jacquez, J. A. Compartmental Analysis in Biology and Medicine (2nd ed). Ann Arbor: The University of Michigan Press, 1985.
 78. Jacquez, J. A., B. Carnahan, and P. Abbrecht. A model of the renal cortex and medulla. Math. Biosci. 1: 227‐261, 1967.
 79. Jacquez, J. A., D. Foster, and E. Daniels. Solute concentration in the kidney—I. A model of the renal medulla and its limit cases. Math. Biosci. 32: 307‐335, 1976.
 80. Jamison, R. L. Micropuncture study of segments of thin loop of Henle in the rat. Am. J. Physiol. 215: 236‐242, 1968.
 81. Jamison, R. L. Short and long loop nephrons. Kidney Int. 31: 597‐605, 1987.
 82. Jamison, R. L., C. M. Bennett, and R. W. Berliner. Countercurrent multiplication by the thin loops of Henle. Am. J. Physiol. 212: 357‐366, 1967.
 83. Jamison, R. L., and W. Kriz. Urinary Concentrating Mechanism: Structure and Function. New York: Oxford University Press, 1982.
 84. Jamison, R. L., F. B. Lacy, J. P. Pennell, and V. M. Sanjana. Potassium secretion by the descending limb or pars recta of the juxtamedullary nephron in vivo. Kidney Int. 9: 323‐332, 1976.
 85. Jamison, R. L., J. Work, and J. A. Schafer. New pathways for potassium transport in the kidney. Am. J. Physiol. 242 (Renal Fluid Electrolyte Physiol. 11): F297‐F312, 1982.
 86. Johnson, A. M., and K. S. Crump. Transient solution of a solute cycling model of the renal medulla by using Laplace transforms. In: Proc. Summer Comput. Simulation Conf., Washington, DC, 1976. Lajolla, CA: Simulation Councils, Inc., 1976, p 460‐463.
 87. Johnston, P. A., C. A. Battilana, F. B. Lacy, and R. L. Jamison. Evidence for a concentration gradient favoring outward movement of sodium from the thin loop of Henle. J. Clin. Invest. 59: 234‐240, 1977.
 88. Kaimal, R., B. Kellogg, and J. L. Stephenson. A mathematical model for the transport of PAH analogues in the kidney. Math. Biosci. 69: 103‐129, 1984.
 89. Katchalsky, A., and P. F. Curran. Nonequilibrium Thermodynamics in Biophysics. Cambridge, MA: Harvard University Press, 1965.
 90. Kawamura, S., and J. P. Kokko. Urea secretion by the straight segment of the proximal tubule. J. Clin. Invest. 58: 604‐612, 1976.
 91. Kedem, O., and A. Katchalsky. Thermodynamic analysis of the permeability of biological membranes to non‐electrolytes. Biochim. Biophys. Acta 27: 229‐246, 1958.
 92. Kedem, O., and A. Katchalsky. A physical interpretation of the phenomenological coefficients of membrane permeability. J. Gen. Physiol 45: 143‐179, 1961.
 93. Kellogg, R. B. Osmotic Flow in a Tube with Stagnant Points. College Park, MD: University of Maryland, 1975, Tech. Note BN‐818, IFDAM.
 94. Kellogg, R. B. A priori bounds for renal network flows. In: Proc. Summer Comput. Simulation Conf., Washington, DC, 1976. Lajolla, CA: Simulation Councils, Inc., 1976, p. 456‐459.
 95. Kellogg, R. B. Uniqueness in the Schauder fixed point theorem. Proc. Am. Math. Soc. 60: 207‐210, 1976.
 96. Kellogg, R. B. Difference approximation for a singular perturbation problem with turning points. In: Analytical and Numerical Approaches to Asymptotic Problems in Analysis, edited by S. Axelsson, L. S. Frank and A. van der Sluis. Amsterdam: North Holland, 1981, p. 133‐139.
 97. Kellogg, R. B. Some singular pertubation problems in renal models. J. Math. Anal. Appl. 128: 214‐240, 1987.
 98. Kelman, R. B. A theoretical note on exponential flow in the proximal part of the mammalian nephron. Bull. Math. Biophys. 24: 303‐317, 1962.
 99. Kelman, R. B. Mathematical analysis of sodium reabsorption in proximal part of nephron in presence of nonreabsorbed solute. J. Theor. Biol. 8: 22‐26, 1965.
 100. Kelman, R. B., D. J. Marsh and H. C. Howard. Nonmonotonicity of solutions of linear differential equations occurring in the theory of urine formation. SIAM Rev. 8: 463‐478, 1966.
 101. Kien, G. A., and E. Koushanpour. Digital computer simulation of the nephron: the action of osmotic diuretics. J. Pharmacol. Exp. Ther. 163: 198‐209, 1968.
 102. Kill, F. The mechanism of concentration of urine: an alternative to the countercurrent multiplier hypothesis. J. Oslo City Hosp. 10: 261‐264, 1960.
 103. Kill, F., and K. Aukland. The role of urea in the renal concentration mechanism. Scand.J. Clin. Lab. Invest. 12: 290‐299, 1960.
 104. Kislyakov, Y. Y., and Y. Y. Bagrov. Mathematical modeling of the process of glomerular filtration. Biofizika 18: 897‐901, 1973.
 105. Knepper, M. A. Measurement of osmolality in kidney slices using vapor pressure osmometry. Kidney Int. 21: 653‐655, 1982.
 106. Knepper, M. A. Urea transport in isolated thick ascending limbs and collecting ducts from rats. Am. J. Physiol. 245 (Renal Fluid Electrolyte Physiol. 14): F634‐F639, 1983.
 107. Knepper, M. A. Urea transport in nephron segments from medullary rays of rabbits. Am. J. Physiol. 244 (Renal Fluid Electrolyte Physiol. 13): F622‐F627, 1983.
 108. Knepper, M., and M. Burg. Organization of nephron function. Am. J. Physiol. 244 (Renal Fluid Electrolyte Physiol. 13): F579‐F589, 1983.
 109. Knepper, M. A., R. A. Danielson, G. M. Saidel, and R. S. Post. Quantitative analysis of renal medullary anatomy in rats and rabbits. Kidney Int. 12: 313‐323, 1977.
 110. Knepper, M. A., J. M. Sands, and C.‐L. Chou. Independence of urea and water transport in rat inner medullary collecting duct. Am. J. Physiol. 256 (Renal Fluid Electrolyte Physiol. 25): F610‐F621, 1989.
 111. Knepper, M. A., and J. L. Stephenson. Urinary concentrating and diluting processes. In: Physiology of Membrane Disorders (2nd ed.), edited by T. E. Andreoli, J. F. Hoffman, D. D. Fanestil, and S. G. Schultz. New York: Plenum, 1986, p. 713‐726.
 112. Kokko, J. P. Sodium chloride and water transport in the descending limb of Henle. J. Clin. Invest. 49: 1838‐1846, 1970.
 113. Kokko, J. P. Urea transport in the proximal tubule and the descending limb of Henle. J. Clin. Invest. 51: 1999‐2008, 1972.
 114. Kokko, J. P. Transport characteristics of the thin limbs of Henle. Kidney Int. 22: 449‐453, 1982.
 115. Kokko, J. P., and F. C. Rector, Jr. Countercurrent multiplication system without active transport in inner medulla. Kidney Int. 2: 214‐223, 1972.
 116. Koushanpour, E., R. R. Tarica, and W. F. Stevens. Mathematical simulation of normal nephron function in rat and man. J. Theoret. Biol. 31: 177‐214, 1971.
 117. Kriz, W. Structural organization of the renal medullary counterflow system. Federation Proc. 42: 2379‐2385, 1983.
 118. Kriz, W., and B. Kaissling. Structural organization of the mammalian kidney. In: The Kidney, edited by D. W. Seldin and G. Giebisch. New York: Raven, 1985, vol. 1, p. 265‐306.
 119. Kriz, W., and A. F. Lever. Renal countercurrent mechanisms: structure and function. Am. Heart J. 78: 101‐118, 1969.
 120. Kuhn, W., and A. Ramel. Aktiver Salztransport als möglicher (und wahrscheinlicher) Einzeleffekt bei der Harnkonzentrierung in der Niere. Helv. Chim. Acta 42: 628‐660, 1959.
 121. Kuhn, W., and A. Ramel. Zweierlei Gleichgewichtspotentiale an ionendurchlässigen, Ionen‐intra‐nichtpermutierenden Membranen (Ruhe‐ und Aktionspotential). Helv. Chim. Acta 42: 293‐305, 1959.
 122. Kuhn, W., and K. Ryffel. Herstellung konzentrierter Lösungen aus verdünnten durch blosse Membranwirkung. Ein Modellversuch zur Funktion der Niere. Hoppe‐Seylers Z. Physiol. Chem. 276: 145‐178, 1942.
 123. Lassiter, W. E., C. W. Gottschalk, and M. Mylle. Micropuncture study of net transtubular movement of water and urea in nondiuretic mammalian kidney. Am. J. Physiol. 200: 1139‐1146, 1961.
 124. Layton, H. E. Distribution of Henle's loops may enhance urine concentrating capability. Biophys. J. 49: 1033‐1040, 1986.
 125. Layton, H. E. Concentrating urine in the inner medulla of the kidney. Comments Theor. Biol. 1: 179‐196, 1989.
 126. Layton, H. E. Urea transport in a distributed loop model of the urine concentrating mechanism. Am. J. Physiol. 258 (Renal Fluid Electrolyte Physiol. 27): F1110‐F1124, 1990.
 127. Lemley, K. V., and W. Kriz. Cycles and separations: the histotopography of the urinary concentrating process. Kidney Int. 31: 538‐548, 1987.
 128. Lever, A. F. The vasa recta and countercurrent multiplication. Acta Med. Scand. 178 (Suppl. 434): 1‐43, 1965.
 129. Lin, C. C., and L. A. Segel. Mathematics Applied to Deterministic Problems in the Natural Sciences. New York: Macmillan, 1974, p. 244‐276.
 130. Lory, P. Numerical solution of a kidney model by multiple shooting. Math. Biosci. 50: 117‐128, 1980.
 131. Lory, P. Effectiveness of a salt transport cascade in the renal medulla: computer simulations. Am. J. Physiol. 252 (Renal Fluid Electrolyte Physiol. 21): F1095‐F1102, 1987.
 132. Lory, P., A. Gilg, and M. Horster. Renal countercurrent system: role of collecting duct convergence and pelvic urea predicted from a mathematical model. J. Math. Biol. 16: 281‐304, 1983.
 133. Macey, R. I. Pressure flow patterns in a cylinder with reabsorbing walls. Bull. Math. Biophys. 25: 1‐9, 1963.
 134. Marsh, D. J. Solute and water flows in thin limbs of Henle's loop in the hamster kidney. Am. J. Physiol. 218: 824‐831, 1970.
 135. Marsh, D. J. Osmotic concentration and dilution of the urine. In: The Kidney 3: Morphology, Biochemistry, Physiology, edited by C. Rouiller and A. Muller. New York: Academic, 1971, p. 71‐126.
 136. Marsh, D. J. Computer simulation of renal countercurrent systems. Federation Proc. 42: 2398‐2404, 1983.
 137. Marsh, D. J., R. B. Kelman, and H. C. Howard. The theory of urine formation in water diuresis with implications for antidiuresis. Bull. Math. Biophys. 29: 67‐89, 1967.
 138. Marsh, D. J., and L. A. Segel. Analysis of countercurrent diffusion exchange in blood vessels of the renal medulla. Am. J. Physiol. 221: 817‐828, 1971.
 139. Marumo, F., Y. Yoshikawa, and S. Koshikawa. A study on the concentration mechanism of the renal medulla by mathematical model. Jpn. Circ. J. 31: 1309‐1317, 1967.
 140. Mejia, R. CONKUB: a conversational path‐follower for systems of nonlinear equations. J. Comput. Physics. 63: 67‐84, 1986.
 141. Mejia, R., R. B. Kellogg, and J. L. Stephenson. Comparison of numerical methods for renal network flows. J. Comput. Physics 23: 53‐62, 1977.
 142. Mejia, R., J. M. Sands, J. L. Stephenson, and M. A. Knepper. Renal actions of atrial natriuretic factor: a mathematical modeling study. Am. J. Physiol. 257 (Renal Fluid Electrolyte Physiol. 26): F1146‐F1157, 1989.
 143. Mejia, R., and J. L. Stephenson. Numerical solution of a central core model of the renal medulla. In: Proc. Comput. Simulation Conf., Montreal, 1973. Lajolla, CA: Simulation Councils, Inc., 1973, vol. 2, p. 806‐810.
 144. Mejia, R., and J. L. Stephenson. Numerical solution of multinephron kidney equations. J. Comput. Physics. 32: 235‐246, 1979.
 145. Mejia, R., and J. L. Stephenson. Solution of a multinephron, multisolute model of the mammalian kidney by Newton and continuation methods. Math. Biosci. 68: 279‐298, 1984.
 146. Mejia, R., J. L. Stephenson, and R. J. LeVeque. A test problem for kidney models. Math. Biosci. 50: 129‐131, 1980.
 147. Moore, L. C., and D. J. Marsh. How descending limb of Henle's loop permeability affects hypertonic urine formation. Am. J. Physiol. 239 (Renal Fluid Electrolyte Physiol. 8): F57‐F71, 1980.
 148. Moore, L. C., D. J. Marsh, and R. E. Kalaba. A simulation study of the mode of osmotic equilibration by limbs of Henle's loop in the kidney medulla. In: Proc. Summer Computer Simulation Conf., Washington, DC, 1976. Lajolla, CA: Simulation Councils, Inc., 1976, p. 515‐517.
 149. Moore, L. C., D. J. Marsh, and C. M. Martin. Loop of Henle during the water‐to‐antidiuresis transition in Brattleboro rats. Am. J. Physiol. 239 (Renal Fluid Electrolyte Physiol. 8): F72‐F83, 1980.
 150. Morel, F., and C. de Rouffignac. Micropuncture study of urea medullary recycling in desert rodents. In: Proc. Int. Congr. Urea and the Kidney, Sarasota, FL, 1968. Amsterdam: Excerpta Med., 1968, p. 401‐413. (Int. Congr. Ser. 195.)
 151. Morel, F., M. Imbert‐Teboul, and D. Chabardes. Distribution of hormone‐dependent adenylate cyclase in the nephron and its physiological significance. Annu. Rev. Physiol. 43: 569‐581, 1981.
 152. Niesel, W., and H. Röskenbleck. Die Bedeutung der Stromgeschwindigkeiten in den (Gefäss‐systemen der Niere und der Schwimmblase für die Aufrechterhaltung von Konzentrationsgradienten. Pflugers Arch. 277: 302‐315, 1963.
 153. Niesel, W., and H. Röskenbleck. Möglichkeiten der Konzentrierung von Stoffen in biologischen Gegenstromsystemen. Pflugers Arch. 276: 555‐567, 1963.
 154. Niesel, W., and H. Röskenbleck. Konzentrierung von Lösungen unterschiedlicher Zusammensetzung durch alleinige Gegenstromdiffusion und Gegenstromosmose als möglicher Mechanismus der Harnkonzentrierung. Pflugers Arch. 283: 230‐241, 1965.
 155. Niesel, W., H. Röskenbleck, P. Hanke, N. Specht, and L. Heuer. Die gegenseitige Beeinflussung von Harnstoff, NaCl, KCI und Harnfluss bei der Bildung eines maximal konzentrierten Harns. Pflugers Arch. 315: 308‐320, 1970.
 156. Ortega, J. M., and W. C. Rheinboldt. Iterative Solution of Nonlinear Equations in Several Variables. New York: Academic, 1970.
 157. Packer, J. S., and J. E. Packer. An analogue‐computer simulation of the facultative water‐reabsorption process in the human kidney—a vascular role for a.d.h. Med. Biol. Eng. 11: 310‐318, 1973.
 158. Packer, J. S., and J. E. Packer. Recycling of urea in the rat kidney: a dynamic self regulating analogue computer simulation. Med. Biol. Eng. 12: 633‐646, 1974.
 159. Packer, J. S., and J. E. Packer. Medullary sodium depletion during diuresis—a digital computer simulation. Med. Biol. Eng. Comput. 15: 134‐139, 1977.
 160. Palatt, P. J. Hydrodynamic effects in a capillary, counter‐current system. Proc. Summer Computer Simulation Conf., Washington, DC, 1976. LaJolla, CA: Simulation Councils, Inc., 1976, p. 464‐467.
 161. Palatt, P. J., and G. M. Saidel. An analysis of countercurrent exchange with emphasis on renal function. Bull Math. Biol. 35: 275‐286, 1973.
 162. Palatt, P. J., and G. M. Saidel. Countercurrent exchange in the inner renal medulla: vasa recta‐descending limb system. Bull Math. Biol. 35: 431‐447, 1973.
 163. Palatt, P. J., G. M. Saidel, and M. Macklin. Transport processes in the renal cortex. J. Theor. Biol. 29: 251‐274, 1970.
 164. Pallone, T. L., Y. Yagil, and R. L. Jamison. Effect of small‐solute gradients on transcapillary fluid movement in renal inner medulla. Am. J. Physiol. 257 (Renal Fluid Electrolyte Physiol. 26): F547‐F553, 1989.
 165. Pennell, J. P., F. B. Lacy, and R. L. Jamison. An in vivo study of the concentrating process in the descending limb of Henle's loop. Kidney Int 5: 337‐347, 1974.
 166. Pennell, J. P., V. Sanjana, N. R. Frey, and R. L. Jamison. The effect of urea infusion on the urinary concentrating mechanism in protein‐depleted rats. J. Clin. Invest. 55: 399‐409, 1975.
 167. Persson, E. Water permeability in rat distal tubules. Acta Physiol. Scand. 78: 364‐375, 1970.
 168. Pinter, G. G., and J. L. Shohet. Origin of sodium concentration profile in the renal medulla. Nature 200: 955‐958, 1963.
 169. Rocha, A. S., J. P. Kokko. Sodium chloride and water transport in the medullary thick ascending limb of Henle: evidence for active chloride transport. J. Clin. Invest. 52: 612‐623, 1973.
 170. Röcha, A. S., and L. H. Kudo. Water, urea, sodium, chloride, and potassium transport in the in vitro isolated perfused papillary collecting duct. Kidney Int. 22: 485‐491, 1982.
 171. Röskenbleck, H., and W. Niesel. Gekoppelte Gegenstrom‐systeme als Modell de konzentrierenden Niere. Pflugers Arch. 277: 316‐324, 1963.
 172. Saidel, G. M., P. S. Chandhoke, and M. A. Knepper. Spatially discrete models of counter‐current mass transport for application to the kidney. Math. Comput. Simul. 20: 259‐270, 1978.
 173. Salane, D. E., and R. P. Tewarson. A unified derivation of symmetric quasi‐Newton update formulas. J. Inst. Math. Applic. 25: 29‐36, 1980.
 174. Sands, J. M., and M. A. Knepper. Urea permeability of mammalian inner medullary collecting duct system and papillary surface epithelium. J. Clin. Invest. 79: 138‐147, 1987.
 175. Sanjana, V. M., P. A. Johnston, C. R. Robertson, and R. L. Jamison. An examination of transcapillary water flux in renal inner medulla. Am. J. Physiol. 231: 313‐318, 1976.
 176. Sanjana, V. M., C. R. Robertson, and R. L. Jamison. Water extraction from the inner medullary collecting tubule system: a role for urea. Kidney Int. 10: 139‐146, 1976.
 177. Sasaki, S., and M. Imai. Effects of vasopressin on water and NaCl transport across the in vitro perfused medullary thick ascending limb of Henle's loop of mouse, rat, and rabbit kidneys. Pflugers Arch. 383: 215‐221, 1980.
 178. Schafer, J. A., C. S. Patlak, and T. E. Andreoli. Osmosis in cortical collecting tubules: a theoretical and experimental analysis of the osmotic transient phenomenon. J. Gen. Physiol. 64: 201‐227, 1974.
 179. Schmidt‐Nielsen, B., and R. O'Dell. Structure and concentrating mechanism in the mammalian kidney. Am. J. Physiol. 200: 1119‐1124, 1961.
 180. Scholander, P. F. Secretion of gases against high pressures in the swim‐bladder of deep sea fishes. II. The rete mirabile. Biol. Bull. 107: 260‐277, 1954.
 181. Scholander, P. F. The wonderful net. Sci. Am. 196: 96‐107, 1957.
 182. Scholander, P. F., and J. Krog. Countercurrent heat exchange and vascular bundles in sloths. J. Appl. Physiol. 10: 405‐411, 1957.
 183. Schutz, W., and J. Schermann. Pelvic urine composition as a determinant of inner medullary solute concentration and urine osmolarity. Pflugers Arch. 334: 154‐166, 1972.
 184. Segel, L. A. Standing‐gradient flows driven by active solute transport. J. Theor. Biol. 29: 233‐250, 1970.
 185. Shohet, J. L., and G. G. Pinter. Derivation of the partial differential equations utilized in a model describing the Na concentration profile in the renal medulla. Nature 204: 689‐690, 1964.
 186. Smith, S. W. The Kidney: Structure and Function in Health and Disease. New York: Oxford University Press, 1951.
 187. Sonnenberg, H. Medullary collecting‐duct function in antidiuretic and in salt‐ or water‐diuretic rats. Am. J. Physiol. 226: 501‐506, 1974.
 188. Staverman, A. J. The theory of measurement of osmotic pressure. Rec. Trav. Chim. Paysbas 70: 344‐352, 1951.
 189. Staverman, A. J. Non‐equilibrium thermodynamics of membrane processes. Trans. Faraday Soc. 48: 176‐185, 1952.
 190. Stephenson, J. L. Ability of counterflow systems to concentrate. Nature 206: 1215‐1219, 1965.
 191. Stephenson, J. L. Concentration in renal counterflow systems. Biophys. J. 6: 539‐551, 1966.
 192. Stephenson, J. L. Consistency of equations for solute and water movement in the renal medulla. Biophys. J. 11: 277a, 1971.
 193. Stephenson, J. L. Concentration of urine in a central core model of the renal counterflow system. Kidney Int. 2: 85‐94, 1972.
 194. Stephenson, J. L. Concentrating engines and the kidney. I. Central core model of the renal medulla. Biophys. J. 13: 512‐545, 1973.
 195. Stephenson, J. L. Concentrating engines and the kidney. II. Multisolute central core systems. Biophys. J. 13: 546‐567, 1973.
 196. Stephenson, J. L. The mathematical theory of renal function. In: Engineering Principles in Physiology, edited by J. H. V. Brown and D. S. Gann. New York: Academic, 1973, vol. 2, p. 283‐320.
 197. Stephenson, J. L. Transient behavior of the single loop solute cycling countercurrent multiplier. Bull. Math. Biol. 35: 183‐194, 1973.
 198. Stephenson, J. L. Free‐energy balance in renal counterflow systems. Math. Biosci. 21: 299‐310, 1974.
 199. Stephenson, J. L. Concentrating engines and the kidney. III. Canonical mass balance equation for multinephron models of the renal medulla. Biophys. J. 16: 1273‐1286, 1976.
 200. Stephenson, J. L. Analysis of the transient behavior of kidney models. Bull. Math. Biol. 40: 211‐221, 1978.
 201. Stephenson, J. L. Countercurrent transport in the kidney. Annu. Rev. Biophys. Bioeng. 7: 315‐339, 1978.
 202. Stephenson, J. L. Case studies in renal and epithelial physiology. In: Mathematical Aspects of Physiology. Lectures in Applied Mathematics, edited by F. C. Hoppensteadt. Providence, RI: American Mathematical Society, 1981, vol. 19, p. 171‐212.
 203. Stephenson, J. L. Concentrating engines and the kidney. IV. Mass balance in a single stage of a multistage model of the renal medulla. Math. Biosci. 55: 265‐278, 1981.
 204. Stephenson, J. L. Renal concentrating mechanisms: introduction. Federation Proc. 42: 2377‐2378, 1983.
 205. Stephenson, J. L. Renal concentrating mechanism: fundamental theoretical concepts. Federation Proc. 42: 2386‐2391, 1983.
 206. Stephenson, J. L. Models of the urinary concentrating mechanism. Kidney Int. 31: 648‐661, 1987.
 207. Stephenson, J. L., and R. W. Berliner. Renal concentrating mechanism: summary. Federation Proc. 42: 2405, 1983.
 208. Stephenson, J. L., and J. F. Jen. Do osmolytes help to drive countercurrent multiplication in thin limbs of Henle's loops? Kidney Int. 37: 572, 1990.
 209. Stephenson, J. L., J. F. Jen, H. Wang, and R. P. Tewarson. Convective uphill transport of NaCl from ascending limb of Henle's loop. J. Am. Soc. Nephrol. 2: 727, 1991.
 210. Stephenson, J. L., R. Mejia, and R. P. Tewarson. Model of solute and water movement in the kidney. Proc. Natl. Acad. Sci. USA 73: 252‐256, 1976.
 211. Stephenson, J. L., R. P. Tewarson, and R. Mejia. Quantitative analysis of mass and energy balance in non‐ideal models of the renal counterflow system. Proc. Natl. Acad. Sci. USA 71: 1618‐1622, 1974.
 212. Stephenson, J. L., Y. Zhang, A. Eftekhari, and R. Tewarson. Electrolyte transport in a central core model of the renal medulla. Am. J. Physiol. 253 (Renal Fluid Electrolyte Physiol. 22): F982‐F997, 1987.
 213. Stephenson, J. L., Y. Zhang, and R. Tewarson. Electrolyte, urea, and water transport in a two‐nephron central core model of the renal medulla. Am. J. Physiol. 257 (Renal Fluid Electrolyte Physiol. 26): F399‐F413, 1989.
 214. Stewart, J. Urea handling by the renal countercurrent system: insights from computer simulation. Pflugers Arch. 356: 133‐151, 1975.
 215. Stewart, J., M. E. Luggen, H. Valtin. A computer model of the renal countercurrent system. Kidney Int. 2: 253‐263, 1972.
 216. Stewart, J., and H. Valtin. Computer simulation of osmotic gradient without active transport in renal inner medulla. Kidney Int. 2: 264‐270, 1972.
 217. Tabei, K., and M. Imai. K+ transport in upper portion of descending limbs of long‐loop nephron from hamster. Am. J. Physiol. 252 (Renal Fluid Electrolyte Physiol. 21): F387‐F392, 1987.
 218. Taniguchi, J., K. Tabei, and M. Imai. Profiles of water and solute transport along long‐loop descending limb: analysis by mathematical model. Am. J. Physiol. 252 (Renal Fluid Electrolyte Physiol. 21): F393‐F402, 1987.
 219. Tewarson, R. P. Sparse matrix methods and mathematical models of the renal concentrating mechanism. Proc. Summer Comput. Simulation Conf., Washington, DC, 1976. LaJolla, CA: Simulation Councils, Inc., 1976, p. 500‐501.
 220. Tewarson, R. P. Use of smoothing and damping techniques in the solution of nonlinear equations. SIAM Rev. 19: 35‐45, 1977.
 221. Tewarson, R. P. A unified derivation of quasi‐Newton methods for solving non‐sparse and sparse nonlinear equations. Computing 21: 113‐125, 1979.
 222. Tewarson, R. P. On the use of Simpson's rule in renal models. Math Biosci. 55: 1‐5, 1981.
 223. Tewarson, R. P. A seventh‐order numerical method for solving boundary value nonlinear ordinary differential equations. Int. J. Num. Methods Eng. 18: 1313‐1319, 1982.
 224. Tewarson, R. P. On the solution of sparse non‐linear equations and some applications. In: Sparsity and Applications, edited by D. J. Evans. Cambridge: Cambridge University Press, 1984, p. 137‐152.
 225. Tewarson, R. P. A review of computational techniques in flow network models. Mausam 36: 441‐446, 1985.
 226. Tewarson, R. P., and P. Farahzad. On the numerical solution of differential equations for renal counterflow systems. Comput. Biomed. Res. 11: 381‐391, 1978.
 227. Tewarson, R. P., and S. Gupta. A sparse matrix method for renal models. Math. Biosci. 61: 191‐203, 1982.
 228. Tewarson, R. P., and N. S. Huslak. An adaptive implementation of interpolation methods for boundary value ordinary differential equations. BIT 23: 382‐387, 1983.
 229. Tewarson, R. P., S. Kim, and J. L. Stephenson. Using quasi‐Newton methods for kidney modeling equations. Appl. Math. Lett. 3: 93‐96, 1989.
 230. Tewarson, R. P., A. Kydes, J. L. Stephenson, and R. Mejia. Use of sparse matrix techniques in numerical solution of differential equations for renal counterflow systems. Comput. Biomed. Res. 9: 507‐520, 1976.
 231. Tewarson, R. P., and J. L. Stephenson. Using quasi‐Newton methods for kidney modeling equations. Presented as part of an Invited Paper at the Seventh International Conference on Computer and Mathematical Modeling, August 2‐5, 1989, Chicago, II.
 232. Tewarson, R. P., J. L. Stephenson, M. Garcia, and Y. Zhang. On the solution of equations for renal counterflow models. Comput. Biol. Med. 15: 287‐295, 1985.
 233. Tewarson, R. P., J. L. Stephenson, and L. L. Juang. A note on solution of large sparse systems of nonlinear equations. J. Math. Anal. Appl. 63: 439‐445, 1978.
 234. Tewarson, R. P., and Y. Zhang. Solution of two‐point boundary value problems using splines. Int. J. Num. Methods Eng. 23: 707‐710, 1986.
 235. Tewarson, R. P., and Y. Zhang. Sparse quasi‐Newton LDU updates. Int. J. Num. Methods Eng. 24: 1093‐1100, 1987.
 236. Ullrich, K. J. Permeability characteristics of the mammalian nephron. With Appendix: Sauer, F. Nonequilibrium thermodynamics of kidney tubule transport. In: Handbook of Physiology, Renal Physiology, edited by J. Orloff, R. W. Berliner. Washington, DC: Am. Physiol. Soc., 1973, Sect. 8, Ch. 12, p. 377‐414.
 237. Ullrich, K. J., K. H. Jarausch, and W. Overbeck. Verteilung von Na, K, Ca, Mg, CI, PO4 und Harnstoffin Rinde und Mark der Hundeniere bei Verschiedenen Funktionszustanden. Ber. Ges. Physiol. 180: 131‐132, 1956.
 238. Ullrich, K. J., K. Kramer, and J. Boylan. Present knowledge of the counter‐current system in the mammalian kidney. In: Heart, Kidney and Electrolytes, edited by C. K. Friedberg. New York: Grune & Stratton, 1962, p. 1‐37.
 239. Ullrich, K. J., B. Schmidt‐Nielsen, R. O'Dell, G. Pehling, C. W. Gottschalk, W. E. Lassiter, and M. Mylle. Micropuncture study of composition of proximal and distal tubular fluid in rat kidney. Am. J. Physiol. 204: 527‐531, 1963.
 240. Walser, M. Mathematical aspects of renal function: the dependence of solute reabsorption on water reabsorprion, and the mechanism of osmotic natriuresis. J. Theor. Biol. 10: 307‐326, 1966.
 241. Weinstein, A. M. Thermodynamic relations in a system of parallel flow tubes. Math. Biosci. 36: 1‐14, 1977.
 242. Weinstein, A. M. Nonequilibrium thermodynamic model of the rat proximal tubule epithelium. Biophys. J. 44: 153‐170, 1983.
 243. Weinstein, A. M. An equation for flow in the renal proximal tubule. Bull. Math. Biol. 48: 29‐57, 1986.
 244. Weinstein, A. M. A mathematical model of the rat proximal tubule. Am. J. Physiol. 250 (Renal Fluid Electrolyte Physiol. 19): F860‐F873, 1986.
 245. Weinstein, A. M. Osmotic diuresis in a mathematical model of the rat proximal tubule. Am. J. Physiol. 250 (Renal Fluid Electrolyte Physiol. 19): F874‐F884, 1986.
 246. Weinstein, A. M., and J. L. Stephenson. Electrolyte transport across a simple epithelium: steady‐state and transient analysis. Biophys. J. 27: 165‐186, 1979.
 247. Weinstein, A. M., J. L. Stephenson, and K. R. Spring. The coupled transport of water. In: Membrane Transport, New Comprehensive Biochemistry, edited by S. L. Bonting and J. J. H. H. M. de Pont. Amsterdam: Elsevier/North‐Holland Biomedical Press, 1981, vol. 2, p. 311‐351.
 248. Wesson, L. G., Jr. A theoretical analysis of urea excretion by the mammalian kidney. Am. J. Physiol. 179: 364‐371, 1954.
 249. Wexler, A. S. Automatic evaluation of derivatives. Appl. Math Comput. 24: 19‐46, 1987.
 250. Wexler, A. S. Solution of nonlinear boundary value problems coupled to a system of algebraic equations using quasilinearization. Nonlinear Anal. Theory Methods Appl. 11: 691‐696, 1987.
 251. Wexler, A. S., R. E. Kalaba, and D. J. Marsh. Automatic derivative evaluation in solving boundary value problems: the renal medulla. Am. J. Physiol. 251 (Renal Fluid Electrolyte Physiol. 20): F358‐F378, 1986.
 252. Wexler, A. S., R. E. Kalaba, and D. J. Marsh. Passive, one‐dimensional countercurrent models do not simulate hypertonic urine formation. Am. J. Physiol. 253 (Renal Fluid Electrolyte Physiol. 22): F1020‐F1030, 1987.
 253. Wexler, A. S., R. E. Kalaba, and D. J. Marsh. Three‐dimensional anatomy and renal concentrating mechanism. I. Modeling results. Am. J. Physiol. 260 (Renal Fluid Electrolyte Physiol. 29): F368‐F383, 1991.
 254. Wexler, A. S., R. E. Kalaba, and D. J. Marsh. Three‐dimensional anatomy and renal concentrating mechanism. II. Sensitivity results. Am. J. Physiol. 260 (Renal Fluid Electrolyte Physiol. 29): F384‐F394, 1991.
 255. Wirz, V. H. Der osmotische Druck des Blutes in der Nierenpapille. Helv. Physiol. Pharmacol. Acta 11: 20‐29, 1953.
 256. Wirz, V. H. Druckmessung in Kapillaren und Tubuli der Niere durch Mikropunktion. Helv. Physiol. Pharmacol. Acta 13: 42‐49, 1955.
 257. Wirz, V. H. Der osmotische Druck in den corticalen Tubuli der Rattenniere. Helv. Physiol. Pharmacol. 14: 353‐362, 1956.
 258. Wirz, H. The location of antidiuretic action in the mammalian kidney. In: The Neurohypophysis, edited by H. Heller. New York: Academic, 1957, p. 157‐169.
 259. Wirz, H., and R. Dirix. Urinary concentration and dilution. In: Handbook of Physiology, Renal Physiology, edited by J. Orloff and R. W. Berliner. Washington, DC: Am. Physiol. Soc., Sect. 8, Chap. 13, 1973, p. 415‐430.
 260. Wirz, H., B. Hrgitay, and W. Kuhn. Lokalisation des Konzentrierungsprozesses in der Niere durch direkte Kryoskopie. Helv. Physiol. Pharmacol. Acta 9: 196‐207, 1951.
 261. Yancey, P. H., and M. Burg. Distribution of major organic osmolytes in rabbit kidneys in diuresis and antidiuresis. Am. J. Physiol. 257 (Renal Fluid Electrolyte Physiol. 26): F602‐F607, 1989.
 262. Zhang, Y., and R. P. Tewarson. Least change updates to Cholesky factor subject to the nonlinear quasi‐Newton condition. IMAA J. Num. Anal. 7: 509‐521, 1987.

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

John L. Stephenson. Urinary Concentration and Dilution: Models. Compr Physiol 2011, Supplement 25: Handbook of Physiology, Renal Physiology: 1349-1408. First published in print 1992. doi: 10.1002/cphy.cp080230