Comprehensive Physiology Wiley Online Library

Basic Principles of Transport

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Thermodynamics
1.1 Maximal Work Is Attained on Reversible Paths
1.2 Changes in Free Energy Are Obtained from Chemical Potentials: Free Energy per Mole
1.3 Solutes in Phase Equilibrium
1.4 Osmotic Equilibrium
1.5 Chemiosmotic Coupling
2 Diffusion
2.1 Flux Is Proportional to the Concentration Gradient: Fick's First Law
2.2 Conservation of Matter: Fick's Second Law
2.3 Progress of Diffusion: Mean Square Displacement = 2 Dt
2.4 Randomly Diffusing Molecules Spread Out in a Normal Distribution
2.5 The Diffusion Front, Where Solute Depletion Ends and Accumulation Begins, Moves as 2nDt
2.6 Diffusion Transients in Thin Membranes Are Very Rapid
2.7 Permeation through Membranes Takes Place in At Least Three Steps
2.8 The Exchange Time for Filling or Emptying of Cells Equals V/PA
2.9 In Simple Exponential Processes, the Time Constant = the Mean Residence Time
2.10 Cytoplasmic Diffusion Transients Are Rapid
2.11 Unstirred Layers Can Be a Significant Barrier
2.12 Membrane Diffusion Is Assumed to Be Rate‐Limiting for Plasma Membranes
2.13 Selective Permeability of Lipid Bilayers Is Determined Primarily by Solubility
3 Water Transport
3.1 Water Transport Can Be Driven by Three Different Gradients
3.2 In Lipid Membranes Water Transport Occurs by Solubility‐Diffusion Mechanism, with Pf = Posm = Pd
3.3 Osmotic Gradients Generate Hydraulic Pressure Gradients in Aqueous Channels
3.4 In Narrow Channels the Ratio Posmotic/Pdiffusion = the Number of Water Molecules Contained within the Channel
3.5 Coupling of Solute and Solvent Transport Is Described by the Kedem‐Katchalsky Equations
4 Ionic Diffusion
4.1 Diffusion with Superimposed Drift Due to External Forces
4.2 Ions Transported by Simple Diffusion Follow the Ussing Flux Ratio Relation
4.3 Bulk Solutions Carry No Net Charge
4.4 The Constant Field Is a Convenient Idealization
4.5 Conductance Depends on Ionic Concentrations
4.6 Permeability Ratios Can Be Measured by Changes in Membrane Potential
4.7 An Electrogenic Pump Contributes to Ψm
4.8 Channels Can Be Incorporated into the Nernst‐Planck Formulation
5 Energy Barriers
5.1 The Born Energy Estimates the Work Required to Transfer an Ion from One Medium to Another
5.2 Born Energy, Image Forces, Dipole Potentials, and Hydrophobic Interactions Contribute to the Energy Barriers of Lipid Bilayers
5.3 Solvation Energies Are Important Determinants of Channel Accessibility
5.4 Surface Potentials Modify the Transmembrane Potential as Well as Local Ion Concentrations
5.5 Transport across Energy Barriers
5.6 Eyring Rate Theory: Rate Constants Depend on Ψ
5.7 Kinetic Approaches
6 Channels
6.1 Single‐Occupancy Channels with Binding Sites Show Saturation Kinetics
6.2 Competition, Unidirectional Flux, and the Ussing Flux Ratio
6.3 Single‐Occupancy Channels: Voltage Dependence in Symmetric Channels
6.4 Single‐Occupancy‐Channel Results Can Be Generalized to N Sites
6.5 Multiple Occupancy
7 Simple Carriers
7.1 Net Flux
7.2 Unidirectional Flux
7.3 Equilibrium at the Boundaries
7.4 Rate‐Limiting Steps at the Boundaries
7.5 Energy‐driven Simple Carrier Systems
8 Cotransport
8.1 Thermodynamics: Cotransport Can Move Solutes “Uphill”
8.2 Kinetic Description
8.3 General Net Flux
8.4 Unidirectional Fluxes
8.5 Kinetics of Simultaneous Binding of mA and nB Resembles Simple Carrier Kinetics When Carrier Concentrations Are Replaced with the Product AmBn
8.6 Relations between the Net Flux Equation Parameters and the Michaelis‐Menten—type Parameters for 1:1 Stoichiometry
9 Countertransport
9.1 Thermodynamics
9.2 Kinetic Models
9.3 Ping‐Pong Model
9.4 Sequential Model
9.5 Generalization to m‐n Stoichiometry for Simultaneous Binding
10 Fluctuating Barriers: Channels and Carriers
10.1 Channel Transport Properties Depend on the Rate of Transition between the Conformations
10.2 Channels with Fluctuating Barriers Do Not Show Michaelis‐Menten Kinetics
10.3 Channels with Fluctuating Barriers Can Show Carrier Kinetics
Figure 1. Figure 1.

A linear spring attached to a weightless pan containing sand rises a distance h when sand is removed.

Figure 2. Figure 2.

Two ways of depicting a cyclic reaction. After reacting with membrane carrier Y on the inner surface of the membrane, solute C is pumped out of the cell. Energy is supplied by the substrate S, which is degraded to P.

Figure 3. Figure 3.

Lattice model of diffusion. The solute at position xi is “caged” by surrounding water molecules. It may advance to position xi+1 when it acquires sufficient thermal energy to break out of its cage. Energy barriers separating the solute from adjacent positions are illustrated, with average distance between activation energy peaks of height ΔGo given by Λ. J represents net flux from xi to xi+1.

Figure 4. Figure 4.

Common types of symmetry encountered in diffusion problems: A. spherical symmetry; B. cylindrical symmetry; C. planar symmetry.

Figure 5. Figure 5.

Concentration profiles, described by Equation 71, show the fate of solute molecules that originate (time t = to = 0) within the thin shaded rectangle as they diffuse through the solution. The y axis represents the concentration at each position. As time progresses (e.g., at t = t1) the solutes diffuse in both directions, depleting the rectangle and creating a bell‐shaped (Gaussian) curve. The curve flattens more and more with time (e.g., t = t2) until at t = ∞ the concentration profile becomes uniformly flat.

Figure 6. Figure 6.

Plots illustrating time course of diffusion from a source originally located at x = 0. The top plot is similar to Figure 5. The second and third plots illustrate the corresponding flux and solute accumulation respectively (n=1; see text). The shaded area distinguishes regions of solute accumulation from regions of depletion.

Figure 7. Figure 7.

Concentration profiles within a homogeneous membrane following sudden addition of solute to side “in” on the left. The concentration on side “o” at x > d is maintained at zero. The y axis represents concentration, the x axis distance. At time t = 0 solute is added at concentration C = Cin to the left‐hand side. At that instant there is no solute within the membrane: the concentration profile is represented by the black line. As time progresses, the membrane fills with solute and the concentration profile quickly approaches the steady state shown by the straight line at t = ∞. Intermediate profiles are represented by the curves labeled t1 and t2. For most practical purposes, the steady state is reached “instantly” (see text).

Figure 8. Figure 8.

A symmetric homogeneous membrane is represented by a series of energy barriers. Interfacial barriers (barriers 0 and m + 1) are encountered by the solute upon entering or leaving the membrane. Internal barriers (1 to m) are encountered while passing through the membrane.

Figure 9. Figure 9.

Effects of unstirred layers on permeability. The illustration, consisting of three “membranes” in series, shows the concentration profiles through each of these. Concentration is plotted on the y axis, distance on the x axis. P1 and P2 represent the permeabilities of the two unstirred layers; Pm represents the permeability of the membrane.

Figure 10. Figure 10.

A plot of nonelectrolyte permeability in egg phosphati‐dylcholine‐decane bilayers as a function of their partition coefficient in hexadecane redrawn from the paper of Walter and Gutknecht 138. Numbers in the figure correspond to the solutes as follows: 1. water; 2. hydrofluoric acid; 3. ammonia; 4. hydrochloric acid; 5. formic acid; 6. methylamine; 7. formamide; 8. nitric acid; 9. urea; 10. thiocyanic acid; 11. acetic acid; 12. ethylamine; 13. ethanediol; 14. acetamide, 15. propionic acid; 16. 1,2‐propanediol; 17. glycerol; 18. butyric acid; 19. 1,4‐butanediol; 20. benzoic acid; 21. hexanoic acid; 22. salicylic acid; 23. codeine.

Figure 11. Figure 11.

Three ways of measuring water flow through a membrane. A. A hydraulic piston pushing on side “in” forces water through the membrane to side “o.” A permeable mechanical support is provided to prevent mechanical deformation of the membrane. B. An impermeable solute on side “o” draws the water toward it. C. Flux of labeled water placed on side “in.”

Figure 12. Figure 12.

A membrane water channel forms a continuum bridging the two bathing solutions. Both side “in” and the channel contain pure water, while side “o” contains an impermeable solute. Rapid equilibrium at the interfaces requires continuity of the chemical potential for water, but the impermeable solute on side “o” creates a discontinuity in its concentration at the “o” interface. These boundary conditions are satisfied when there is an equal and opposite discontinuity in hydrostatic pressure at the “o” interface. Panel A: Plot of hydrostatic pressure profile through the membrane. The pressure drop across the membrane creates a hydraulic flow of water through the membrane even though both bathing media are at atmospheric pressure. Panel B: Equilibrium (J = 0) ensues when the cryptic pressure drop within the membrane is wiped out by the application of a hydrostatic pressure (PPatm) = ΔII = RTΣCi to side “o.”

Figure 13. Figure 13.

Efflux of solute out of an elementary volume where motion is constrained to the x direction. If the solute velocity = dx/dt, then all solutes contained within the box defined by dxdydz at time t will have passed through the shaded plane in time t + dt.

Figure 14. Figure 14.

Concentration and electrical potential profiles through a membrane.

Figure 15. Figure 15.

Normalized I vs. Ψm plot obtained from Equation 182 for various values of Co while Cin is held fixed. Values for Co/Cin are indicated on the graph. Note that the slope; (conductances) of all curves approach the same limiting value (determined by Equation 186) at infinite Ψm.

Figure 16. Figure 16.

A model channel illustrating the continuum approach with a macroscopic dielectric constant. Channel cross‐sectional area, a(x), varies from the wide‐mouthed vestibule on the left to a narrow, negatively charged, aperture that approximates the size of an ion on the right. The dotted lines show equipotential lines generated by the channel charge; the solid lines show lines of force.

Redrawn from Dani and Levitt 30
Figure 17. Figure 17.

Born energy (dashed lines) for three different ionic radii are obtained from Equation 219. Corresponding Born energy corrected for “image forces” and shown as solid lines are obtained from Equation 221 using the first 20 terms of the series.

Figure 18. Figure 18.

Lipid bilayer energy barriers for a hydrophobic ion with radius = 4.2 Å. Top panel: The upper trace shows a composite consisting of the Born image plus neutral hydrophobic energies. The two lower curves of plot A, “ + Dipole” and “—Dipole,” are plots of dipole energies for cations and anions respectively. Bottom panel shows the composite sum of these energies, one for anions, the other for cations. The negative dipole potential lowers the energy barrier for anions, making them very permeable.

Redrawn from Flewelling and Hubbel 41
Figure 19. Figure 19.

Energy‐barrier profile of gramicidin channel constructed from molecular dynamics simulations by Roux and Karplus. Simulations were accomplished by dividing the channel into four sections labeled I to IV (lower figure), corresponding to sections separated by open circles in energy profile (upper figure). Significance of the different sections: II—transition from bulk water to single file solvation; I—the beginning of single‐file transport; III—single‐file transport; IV—intermolecular region.

Redrawn from Roux and Karplus 120
Figure 20. Figure 20.

Electrical potential profiles in the presence of surface charge. Surface‐charge density on inner surface at left is assumed larger than the corresponding charge density on outer surface at the right. The fast rise in the potential on entering the membrane from either side “in” or “o” is due to a small dipole potential. Notice that the internal transmembrane potential difference is larger than the measured membrane potential.

Figure 21. Figure 21.

Illustration of a simple channel with two binding sites. Each channel can exist in three states: an empty channel Yoo, a channel with C on the inside Yco, and a channel with C on the outside Yoc. The upper triangular sketch shows the transitions among the states. The lower “reaction” scheme shows the reversible transport of Cin to Co.

Figure 22. Figure 22.

Reaction scheme illustrating a single‐occupancy channel and two binding sites, where two solutes, C and B, compete. See Fig. 21 for details.

Figure 23. Figure 23.

Normalized plot of I vs. Ψm (Equation 263) with Cin = Co = K. Values of λ/k−in are indicated in the graph. Channel saturation becomes more evident at lower voltages the larger the λ/k−in (translocation rate compared to dissociation rate). Dotted straight lines show the corresponding linear conductance predicted by corresponding constant‐field equations.

Figure 24. Figure 24.

Reaction scheme illustrating a single‐occupancy channel with n + 2 binding sites. See Figure 21 for details.

Figure 25. Figure 25.

Illustration of a simple carrier where a mobile component Y within the membrane facilitates permeation. Beginning on side “in” the solute Cin combines with an empty reactive carrier site that is exposed on side “in” and is designated as Yoin. The reaction forms a carrier‐solute complex Ycin. Ycin then undergoes a translocation step (conformation change) that exposes the reactive site to side “o,” where it is designated as Yco. The complex Yco then dissociates, releasing Co to side “o.” The free carrier Yoo with the reactive site still facing side “o” will then complete the cycle by translocating its reactive site to side “in,” where it again becomes Yoin.

Figure 26. Figure 26.

Cotransport random carrier (m = n = 1). The cycle begins on side “in” where Ain or Bin combine with its reactive site on component Yooin, forming the complex YAoin or YoBin. This complex reacts with the other solute, forming the ternary complex YABin. YABin undergoes a translocation step, exposing the reactive sites to side “o,” and becomes YABo. YABo dissociates, releasing in a random order Ao or Bo to side “o,” going through the intermediate complexes YAoo or YoBo, before becoming totally in the undissociated form Yooo. Yooo will then translocate them to side “in,” completing the cycle. The association‐dissociation reactions at each boundary are assumed to be in equilibrium. At these boundaries only three out of the four cyclic dissociation constants, Keq, are independent. This is reflected in the terms containing α1 and α2 (see discussion of Equations 331, 332, and 333).

Figure 27. Figure 27.

Reaction cycle used to derive Equation 335. See text.

Figure 28. Figure 28.

Cotransport random: Families of Lineweaver‐Burk plots of 1/JA‐influx versus 1/Ao for increasing values of Bo indicated by the curved arrow. All plots intersect in the second or third quadrant, depending on the relative values of α1 and the product (1+k−1X) (1+k2X). To interpret plots, recall that for each value of Bo, the plot shows a straight line with a y‐axis intercept equal to 1/JmaxA‐influx and an x‐axis intercept equal to −1/KMA‐influx a: α1 < (1 + k−1X) (1+k2X). From Equations 352 and 353, yA = yB > 0 and the intersection lies in the second quadrant… From the changes in the x and y intercepts, it follows that JmaxA‐influx increases with Bo while KMA‐influx decreases. b: If α1 > (1+k−1X) (1+k2X) then yA = yB < 0. The intersection lies below the x axis, in the third quadrant. Both JmaxA‐influx and KMA‐influx will increase with Bo. c: If α1 = (1+k−1X) (1+k2X) then yA = yB = 0. The intersection lies on the x axis. In addition, JmaxA‐influx increases with increasing Bo, while KMA‐influx remains constant, independent of Bo.

Figure 29. Figure 29.

Cotransport sequential. Families of Lineweaver‐Burk plots illustrating graphical criteria for an ordered sequence mechanism for influx with Ao always binding first. a: Plot of 1/JA‐influx versus 1/Ao for increasing values of Bo indicated by the curved arrow. As Bo → ∞, the slope → 0. b: Plot of 1/JB‐influx versus 1/Bo for increasing values of Ao indicated by the curved arrow. All lines intersect on the y axis.

Figure 30. Figure 30.

Countertransport carrier ping‐pong mechanism. Beginning on side “in” the solute Ain or Bin combine with the reactive site of component Yooin. The complex YAin or YBin will then undergo a translocation step, moving the reactive site to side “o.” The complex then dissociates, releasing Ao or Bo to side “o.” The undissociated carrier with the reactive site still facing side “o” will then either translocate its reactive site in the free (Yooo) or bound form (YAo or YBo) to side “in.” Reactions of the solute A and B with the reactive site of component Y are assumed to be in equilibrium.

Figure 31. Figure 31.

Countertransport sequential model with m = n = 1. When A binds to oinYo or ooYin, it forms the complex AinYo or AoYin. This complex will then bind B on the empty reactive site located on the opposite side of the membrane, forming AinYBo or AoYBin. On the other hand, if B binds first, then the complex oinYBo or ooYBin would be formed before AoYBo or AoYBin. Only three out of the four cyclic dissociation constants, Keq, are independent (see Equations 48 and 49). For each step, the dissociation constant is located closest to the dissociation arrow.

Figure 32. Figure 32.

Countertransport ping‐pong. Lineweaver‐Burk plot for the influx of a solute A for different concentrations of B. a: Plot of stimulation of the influx of A by the counter solute (Bin) on the trans side of the membrane where it binds to the free carrier when there is no slippage, b: Plot of a competitive inhibition. Influx of A is decreased by Bo, which competes for the same carrier site on the same side “o” (cis side) of the membrane.

Figure 33. Figure 33.

Reaction scheme for single‐site channel with energy barriers that can fluctuate between two conformations, Y and Y*. The dashed arrows are redundant; they have been drawn to facilitate an overview of all transport pathways.

Figure 34. Figure 34.

Two plots of flux vs. concentration for singly occupied dynamic channels, with different Po obtained from Equation 394. Values of the parameters are shown on each curve. Neither curve is a rectangular hyperbola. Note the maximum in the upper plot.

Figure 35. Figure 35.

Operation of cycle (Equation 390) shows how channel with very large fluctuating barriers behaves like conventional carrier.



Figure 1.

A linear spring attached to a weightless pan containing sand rises a distance h when sand is removed.



Figure 2.

Two ways of depicting a cyclic reaction. After reacting with membrane carrier Y on the inner surface of the membrane, solute C is pumped out of the cell. Energy is supplied by the substrate S, which is degraded to P.



Figure 3.

Lattice model of diffusion. The solute at position xi is “caged” by surrounding water molecules. It may advance to position xi+1 when it acquires sufficient thermal energy to break out of its cage. Energy barriers separating the solute from adjacent positions are illustrated, with average distance between activation energy peaks of height ΔGo given by Λ. J represents net flux from xi to xi+1.



Figure 4.

Common types of symmetry encountered in diffusion problems: A. spherical symmetry; B. cylindrical symmetry; C. planar symmetry.



Figure 5.

Concentration profiles, described by Equation 71, show the fate of solute molecules that originate (time t = to = 0) within the thin shaded rectangle as they diffuse through the solution. The y axis represents the concentration at each position. As time progresses (e.g., at t = t1) the solutes diffuse in both directions, depleting the rectangle and creating a bell‐shaped (Gaussian) curve. The curve flattens more and more with time (e.g., t = t2) until at t = ∞ the concentration profile becomes uniformly flat.



Figure 6.

Plots illustrating time course of diffusion from a source originally located at x = 0. The top plot is similar to Figure 5. The second and third plots illustrate the corresponding flux and solute accumulation respectively (n=1; see text). The shaded area distinguishes regions of solute accumulation from regions of depletion.



Figure 7.

Concentration profiles within a homogeneous membrane following sudden addition of solute to side “in” on the left. The concentration on side “o” at x > d is maintained at zero. The y axis represents concentration, the x axis distance. At time t = 0 solute is added at concentration C = Cin to the left‐hand side. At that instant there is no solute within the membrane: the concentration profile is represented by the black line. As time progresses, the membrane fills with solute and the concentration profile quickly approaches the steady state shown by the straight line at t = ∞. Intermediate profiles are represented by the curves labeled t1 and t2. For most practical purposes, the steady state is reached “instantly” (see text).



Figure 8.

A symmetric homogeneous membrane is represented by a series of energy barriers. Interfacial barriers (barriers 0 and m + 1) are encountered by the solute upon entering or leaving the membrane. Internal barriers (1 to m) are encountered while passing through the membrane.



Figure 9.

Effects of unstirred layers on permeability. The illustration, consisting of three “membranes” in series, shows the concentration profiles through each of these. Concentration is plotted on the y axis, distance on the x axis. P1 and P2 represent the permeabilities of the two unstirred layers; Pm represents the permeability of the membrane.



Figure 10.

A plot of nonelectrolyte permeability in egg phosphati‐dylcholine‐decane bilayers as a function of their partition coefficient in hexadecane redrawn from the paper of Walter and Gutknecht 138. Numbers in the figure correspond to the solutes as follows: 1. water; 2. hydrofluoric acid; 3. ammonia; 4. hydrochloric acid; 5. formic acid; 6. methylamine; 7. formamide; 8. nitric acid; 9. urea; 10. thiocyanic acid; 11. acetic acid; 12. ethylamine; 13. ethanediol; 14. acetamide, 15. propionic acid; 16. 1,2‐propanediol; 17. glycerol; 18. butyric acid; 19. 1,4‐butanediol; 20. benzoic acid; 21. hexanoic acid; 22. salicylic acid; 23. codeine.



Figure 11.

Three ways of measuring water flow through a membrane. A. A hydraulic piston pushing on side “in” forces water through the membrane to side “o.” A permeable mechanical support is provided to prevent mechanical deformation of the membrane. B. An impermeable solute on side “o” draws the water toward it. C. Flux of labeled water placed on side “in.”



Figure 12.

A membrane water channel forms a continuum bridging the two bathing solutions. Both side “in” and the channel contain pure water, while side “o” contains an impermeable solute. Rapid equilibrium at the interfaces requires continuity of the chemical potential for water, but the impermeable solute on side “o” creates a discontinuity in its concentration at the “o” interface. These boundary conditions are satisfied when there is an equal and opposite discontinuity in hydrostatic pressure at the “o” interface. Panel A: Plot of hydrostatic pressure profile through the membrane. The pressure drop across the membrane creates a hydraulic flow of water through the membrane even though both bathing media are at atmospheric pressure. Panel B: Equilibrium (J = 0) ensues when the cryptic pressure drop within the membrane is wiped out by the application of a hydrostatic pressure (PPatm) = ΔII = RTΣCi to side “o.”



Figure 13.

Efflux of solute out of an elementary volume where motion is constrained to the x direction. If the solute velocity = dx/dt, then all solutes contained within the box defined by dxdydz at time t will have passed through the shaded plane in time t + dt.



Figure 14.

Concentration and electrical potential profiles through a membrane.



Figure 15.

Normalized I vs. Ψm plot obtained from Equation 182 for various values of Co while Cin is held fixed. Values for Co/Cin are indicated on the graph. Note that the slope; (conductances) of all curves approach the same limiting value (determined by Equation 186) at infinite Ψm.



Figure 16.

A model channel illustrating the continuum approach with a macroscopic dielectric constant. Channel cross‐sectional area, a(x), varies from the wide‐mouthed vestibule on the left to a narrow, negatively charged, aperture that approximates the size of an ion on the right. The dotted lines show equipotential lines generated by the channel charge; the solid lines show lines of force.

Redrawn from Dani and Levitt 30


Figure 17.

Born energy (dashed lines) for three different ionic radii are obtained from Equation 219. Corresponding Born energy corrected for “image forces” and shown as solid lines are obtained from Equation 221 using the first 20 terms of the series.



Figure 18.

Lipid bilayer energy barriers for a hydrophobic ion with radius = 4.2 Å. Top panel: The upper trace shows a composite consisting of the Born image plus neutral hydrophobic energies. The two lower curves of plot A, “ + Dipole” and “—Dipole,” are plots of dipole energies for cations and anions respectively. Bottom panel shows the composite sum of these energies, one for anions, the other for cations. The negative dipole potential lowers the energy barrier for anions, making them very permeable.

Redrawn from Flewelling and Hubbel 41


Figure 19.

Energy‐barrier profile of gramicidin channel constructed from molecular dynamics simulations by Roux and Karplus. Simulations were accomplished by dividing the channel into four sections labeled I to IV (lower figure), corresponding to sections separated by open circles in energy profile (upper figure). Significance of the different sections: II—transition from bulk water to single file solvation; I—the beginning of single‐file transport; III—single‐file transport; IV—intermolecular region.

Redrawn from Roux and Karplus 120


Figure 20.

Electrical potential profiles in the presence of surface charge. Surface‐charge density on inner surface at left is assumed larger than the corresponding charge density on outer surface at the right. The fast rise in the potential on entering the membrane from either side “in” or “o” is due to a small dipole potential. Notice that the internal transmembrane potential difference is larger than the measured membrane potential.



Figure 21.

Illustration of a simple channel with two binding sites. Each channel can exist in three states: an empty channel Yoo, a channel with C on the inside Yco, and a channel with C on the outside Yoc. The upper triangular sketch shows the transitions among the states. The lower “reaction” scheme shows the reversible transport of Cin to Co.



Figure 22.

Reaction scheme illustrating a single‐occupancy channel and two binding sites, where two solutes, C and B, compete. See Fig. 21 for details.



Figure 23.

Normalized plot of I vs. Ψm (Equation 263) with Cin = Co = K. Values of λ/k−in are indicated in the graph. Channel saturation becomes more evident at lower voltages the larger the λ/k−in (translocation rate compared to dissociation rate). Dotted straight lines show the corresponding linear conductance predicted by corresponding constant‐field equations.



Figure 24.

Reaction scheme illustrating a single‐occupancy channel with n + 2 binding sites. See Figure 21 for details.



Figure 25.

Illustration of a simple carrier where a mobile component Y within the membrane facilitates permeation. Beginning on side “in” the solute Cin combines with an empty reactive carrier site that is exposed on side “in” and is designated as Yoin. The reaction forms a carrier‐solute complex Ycin. Ycin then undergoes a translocation step (conformation change) that exposes the reactive site to side “o,” where it is designated as Yco. The complex Yco then dissociates, releasing Co to side “o.” The free carrier Yoo with the reactive site still facing side “o” will then complete the cycle by translocating its reactive site to side “in,” where it again becomes Yoin.



Figure 26.

Cotransport random carrier (m = n = 1). The cycle begins on side “in” where Ain or Bin combine with its reactive site on component Yooin, forming the complex YAoin or YoBin. This complex reacts with the other solute, forming the ternary complex YABin. YABin undergoes a translocation step, exposing the reactive sites to side “o,” and becomes YABo. YABo dissociates, releasing in a random order Ao or Bo to side “o,” going through the intermediate complexes YAoo or YoBo, before becoming totally in the undissociated form Yooo. Yooo will then translocate them to side “in,” completing the cycle. The association‐dissociation reactions at each boundary are assumed to be in equilibrium. At these boundaries only three out of the four cyclic dissociation constants, Keq, are independent. This is reflected in the terms containing α1 and α2 (see discussion of Equations 331, 332, and 333).



Figure 27.

Reaction cycle used to derive Equation 335. See text.



Figure 28.

Cotransport random: Families of Lineweaver‐Burk plots of 1/JA‐influx versus 1/Ao for increasing values of Bo indicated by the curved arrow. All plots intersect in the second or third quadrant, depending on the relative values of α1 and the product (1+k−1X) (1+k2X). To interpret plots, recall that for each value of Bo, the plot shows a straight line with a y‐axis intercept equal to 1/JmaxA‐influx and an x‐axis intercept equal to −1/KMA‐influx a: α1 < (1 + k−1X) (1+k2X). From Equations 352 and 353, yA = yB > 0 and the intersection lies in the second quadrant… From the changes in the x and y intercepts, it follows that JmaxA‐influx increases with Bo while KMA‐influx decreases. b: If α1 > (1+k−1X) (1+k2X) then yA = yB < 0. The intersection lies below the x axis, in the third quadrant. Both JmaxA‐influx and KMA‐influx will increase with Bo. c: If α1 = (1+k−1X) (1+k2X) then yA = yB = 0. The intersection lies on the x axis. In addition, JmaxA‐influx increases with increasing Bo, while KMA‐influx remains constant, independent of Bo.



Figure 29.

Cotransport sequential. Families of Lineweaver‐Burk plots illustrating graphical criteria for an ordered sequence mechanism for influx with Ao always binding first. a: Plot of 1/JA‐influx versus 1/Ao for increasing values of Bo indicated by the curved arrow. As Bo → ∞, the slope → 0. b: Plot of 1/JB‐influx versus 1/Bo for increasing values of Ao indicated by the curved arrow. All lines intersect on the y axis.



Figure 30.

Countertransport carrier ping‐pong mechanism. Beginning on side “in” the solute Ain or Bin combine with the reactive site of component Yooin. The complex YAin or YBin will then undergo a translocation step, moving the reactive site to side “o.” The complex then dissociates, releasing Ao or Bo to side “o.” The undissociated carrier with the reactive site still facing side “o” will then either translocate its reactive site in the free (Yooo) or bound form (YAo or YBo) to side “in.” Reactions of the solute A and B with the reactive site of component Y are assumed to be in equilibrium.



Figure 31.

Countertransport sequential model with m = n = 1. When A binds to oinYo or ooYin, it forms the complex AinYo or AoYin. This complex will then bind B on the empty reactive site located on the opposite side of the membrane, forming AinYBo or AoYBin. On the other hand, if B binds first, then the complex oinYBo or ooYBin would be formed before AoYBo or AoYBin. Only three out of the four cyclic dissociation constants, Keq, are independent (see Equations 48 and 49). For each step, the dissociation constant is located closest to the dissociation arrow.



Figure 32.

Countertransport ping‐pong. Lineweaver‐Burk plot for the influx of a solute A for different concentrations of B. a: Plot of stimulation of the influx of A by the counter solute (Bin) on the trans side of the membrane where it binds to the free carrier when there is no slippage, b: Plot of a competitive inhibition. Influx of A is decreased by Bo, which competes for the same carrier site on the same side “o” (cis side) of the membrane.



Figure 33.

Reaction scheme for single‐site channel with energy barriers that can fluctuate between two conformations, Y and Y*. The dashed arrows are redundant; they have been drawn to facilitate an overview of all transport pathways.



Figure 34.

Two plots of flux vs. concentration for singly occupied dynamic channels, with different Po obtained from Equation 394. Values of the parameters are shown on each curve. Neither curve is a rectangular hyperbola. Note the maximum in the upper plot.



Figure 35.

Operation of cycle (Equation 390) shows how channel with very large fluctuating barriers behaves like conventional carrier.

References
 1. Abramowitz, M., and Stegun, I. A. Handbook of Mathematical Functions with Formulas Graphs and Mathematical Tables. National Bureau of Standards Applied Mathematics Series 55, 1964.
 2. Agin, D. Electroneutrality and electrodiffusion in the squid axon. Proc. Natl. Acad. Sci. USA 57: 1232–1238, 1967.
 3. Andersen, O. S. Permeability properties of unmodified lipid bilayer membranes. In: Membrane Transport in Biology, edited by G. Giebisch, D. C. Tosteson, and H. H. Ussing,. Berlin: Springer‐Verlag, 1978, p. 369–446.
 4. Andersen, O. S. Kinetics of ion movements mediated by carriers and channels. Methods Enzymol. 171: 62–112, 1989.
 5. Andersen, O. S., A. Finkelstein, I. Katz, and A. Cass. Effect of phloretin on the permeability of thin lipid membranes. J. Gen. Physiol. 67: 749–771, 1976.
 6. Andersen, O. S., and M. Fuchs. Potential energy barriers to ion transport within lipid bilayers: studies with tetraphenylborate. Biophys. J. 15: 795–830, 1975.
 7. Anderson, J. L., and D. M. Malone. Mechanism of osmotic flow in porous membranes. Biophys. J. 14: 957–982, 1974.
 8. Aronson, P. S. Kinetic properties of the plasma membrane Na+‐H+ exchanger. Annu. Rev. Physiol. 47: 546–560, 1985.
 9. Aronson, P. S., and P. Igarashi. Molecular properties and physiological roles of the renal Na+‐H+ exchanger. Curr. Top. Membr. Transp. 26: 57–75, 1986.
 10. Aveyard, R., and D. A. Haydon. An Introduction to the Principles of Surface Chemistry. Cambridge: Cambridge University Press, 1973.
 11. Barr, L. Membrane potential profiles and the Goldman equation. J. Theor. Biol. 9: 351–359, 1965.
 12. Barry, P. H., and J. M. Diamond. Effects of unstirred layers on membrane phenomena. Physiol. Rev. 64: 763–872, 1984.
 13. Begenisich, T., and C. Smith. Multi‐ion nature of potassium channels in squid axons. Curr. Top. Membr. Transp. 22: 353–369, 1984.
 14. Benedek, G. B., and F. M. H. Villars. Physics with Illustrative examples from Biology and Medicine. Reading, MA: Addison‐Wesley, 1979.
 15. Berg, H. C. Random Walks in Biology. Princeton: Princeton University Press, 1983.
 16. Blaustein, M. P. Sodium‐calcium exchange in cardiac, smooth, and skeletal muscle: Key to contractility. Curr. Top. Membr. Transp. 34: 289–330, 1989.
 17. Bockris, J. O., and A. K. N. Reddy. Modern Electrochemistry. New York: Plenum Press, 1970.
 18. Cai, M., and P. C. Jordan. How does vestibule surface charge affect ion conduction and toxin binding in a sodium channel? Biophys. J. 57: 883–891, 1990.
 19. Carslaw, H. S., and J. C. Jaeger. Conduction of Heat in Solids. London: Oxford University Press, 1959.
 20. Chiu, S. W., and E. Jakobsson. Stochastic theory of singly occupied ion channels. II. Effects of access resistance and potential gradients extending into the bath. Biophys. J. 55: 147–157, 1989.
 21. Cole, K. S. Electrodiffusion models for the membrane of squid giant axon. Physiol. Rev. 45: 340–379, 1965.
 22. Colquhoun, D., and A. G. Hawkes. The principles of the stochastic interpretation of ion‐channel mechanisms. In: Single Channel Recording, edited by B. Sakmann and E. Neher,. New York: Plenum, 1983.
 23. Cooper, K., E. Jakobsson, and P. Wolynes. The theory of ion transport through membrane channels. Prog. Biophys. Mol. Biol. 46: 51–96, 1985.
 24. Cooper, K. E., P. Y. Gates, and R. S. Eisenberg. Diffusion theory and discrete rate constants in ion permeation. J. Membr. Biol. 106: 95–105, 1988a.
 25. Cooper, K. E., P. Y. Gates, and R. S. Eisenberg. Surmounting barriers in ionic channels. Q. Rev. Biophys. 21: 331–364, 1988b.
 26. Cornish‐Bowden, A., and C. W. Wharton. Enzyme Kinetics. Oxford: IRL Press, 1988.
 27. Crank, J. The Mathematics of Diffusion. Oxford: Clarendon Press, 1975.
 28. Dainty, J., and C. R. House. “Unstirred layers” in frog skin. J. Physiol. (Lond.) 182: 66–78, 1966.
 29. Dani, J. A., and D. G. Levitt. Water transport and ion‐water interaction in the gramicidin channel. Biophys. J. 35: 501–508, 1981.
 30. Dani, J. A., and D. G. Levitt. Diffusion and kinetic approaches to describe permeation in ionic channels. J. Theor. Biol. 146: 289–301, 1990.
 31. Diamond, J. M., G. Szabo, and Y. Katz. Theory of nonelectrolyte permeation in a generalized membrane. J. Membr. Biol. 17: 148–152, 1974.
 32. Diamond, J. M., and E. M. Wright. Molecular forces governing non‐electrolyte permeation through cell membranes. Proc. R. Soc. Lond. Biol. 172: 273–316, 1969.
 33. DiFrancesco, D., and D. Noble. A model of cardiac electrical activity incorporating ionic pumps and concentration changes. Philos. Trans. R. Soc. Lond. Biol. 307: 353–98, 1985.
 34. Eisenberg, D., and D. Carothers. Physical Chemistry. Menlo Park, CA.: Benjamin Cummings, 1979.
 35. Ferreira, H. G., and M. W. Marshall. The Biophysical Basis of Excitability. New York: Cambridge University Press, 1985.
 36. Finkelstein, A. Water Movement Through Lipid Bilayer, Pores, and Plasma Membranes. New York: John Wiley & Sons, 1987.
 37. Finkelstein, A., and A. Cass. Permeability and electrical properties of thin lipid membranes. J. Gen. Physiol. 52: 145s, 1968.
 38. Finkelstein, A. and A. Mauro. Physical principles and formalisms of electrical excitability. In: Handbook of Physiology. The Nervous System, edited by Eric R. Kandel, Baltimore, Maryland: Am. Physiol. Soc., sect 1, vol. 1, 1977, p. 161–213.
 39. Finkelstein, A., and P. A. Rosenberg. Single file transport: Implications for ion and water movement through gramidicin A channels. In: Membrane Transport Processes, edited by C. F. Stevens and R. W. Tsien,. New York: Raven, 1979, p. 73–88.
 40. Flewelling, R. F., and W. L. Hubbell. Hydrophobic ion interactions with membranes. Biophys. J. 49: 531–540, 1986a.
 41. Flewelling, R. F., and W. L. Hubbell. The membrane dipole potential in a total membrane potential model. Biophys. J. 49: 541–552, 1986b.
 42. Freedman, J. C., and J. F. Hoffman. Ionic and osmotic equilibria of human red blood cells treated with nystatin. J. Gen. Physiol. 74: 157–185, 1979.
 43. Garrahan, P. J., and I. M. Glynn. Driving the sodium pump backwards to form adenosine triphosphate. Nature 211: 1414–1415, 1966.
 44. Glasstone, S. Thermodynamics for Chemists. Huntington, NY: R. E. Krieger, 1972.
 45. Glasstone, S., K. J. Laidler, and H. Eyring. The Theory of Rate Processes. New York: McGraw‐Hill, 1941.
 46. Glynn, I. M. Sodium and potassium movements in human red cells. J. Physiol. 34: 278–310, 1956.
 47. Glynn, I. M. The action of cardiac glycosides on sodium and potassium movements in the human red cells. J. Physiol. 136 (1): 148–173, 1957.
 48. Goldman, D. Potential, impedance, and rectification in membranes. J. Gen. Physiol. 27: 37–60, 1944.
 49. Hardt, S. L. The diffusion transit time: a simple derivation. Bull. Math. Biol. 41: 89–99, 1981.
 50. Helfferich, F. Ion Exchange. New York: McGraw‐Hill, 1962.
 51. Hille, B. Ionic Channels of Excitable Membranes. Sunderland, MA: Sinauer Associates, 1992.
 52. Hille, B., and W. Schwarz. Potassium channels as multi‐ion single file pores. J. Gen. Physiol. 72: 409–442, 1978.
 53. Hodgkin, A. L., and A. F. Huxley. A quantitative description of membrane current and its application to conduction and excitation in nerve. J. Physiol. 117: 500–544, 1952.
 54. Hodgkin, A. L., and R. D. Keynes. The potassium permeability of a giant nerve fibre. J. Physiol. 128: 61–68, 1955.
 55. Horowicz, P., P. W. Gage, and R. S. Eisenberg. The role of the electrochemical gradient in determining potassium fluxes in frog striated muscle. J. Gen. Physiol. 51: 193s–203s, 1968.
 56. Jacquez, J. A. A general relation between membrane potential, ion activities, and pump fluxes for nonsymmetric cells in a steady state. Math. Biosci. 53: 53–57, 1981.
 57. Jakobsson, E., and S. W. Chiu. Stochastic theory of ion movement in channels with single‐ion occupancy. Biophys. J. 52: 33–45, 1987.
 58. Jordan, P. C. Electrostatic modeling of ion pores. II. Effects attributable to the membrane dipole potential. Biophys. J. 41: 189–195, 1983.
 59. Jordan, P. C. Effect of pore structure on energy barriers and applied voltage profiles. I. Symmetrical channels. Biophys. J. 45: 1091–1100, 1984a.
 60. Jordan, P. C. Effect of pore structure on energy barriers and applied voltage profiles. II. Unsymmetrical channels. Biophys. J. 45: 1101–1107, 1984b.
 61. Jordan, P. C. The total eletrostatic potential in a gramicidin channel. J. Membr. Biol. 78: 91–102, 1984c.
 62. Jordan, P. C. Ion channel electrostatics and the shapes of channel proteins. In: Ion Channel Reconstitution, edited by C. Miller,. New York: Plenum Press, 1986, p. 37–55.
 63. Jordan, P. C. How pore mouth charge distributions alter the permeability of transmembrane ionic channels. Biophys. J. 51: 297–311, 1987.
 64. Jordan, P. C., R. J. Bacquet, J. A. McCammon, and P. Tran. How electrolyte shielding influences the electrical potential in transmembrane ion channels. Biophys. J. 55: 1041–1052, 1989.
 65. Katchalsky, A. and P. F. Curran. Nonequilibrium Thermodynamics in Biophysics. Cambridge, MA: Harvard University Press, 1965.
 66. Katz, M. A., and E. H. Bresler. Osmosis. In: Edema, edited by N. C. Staub and A. E. Taylor, New York: Raven Press, 1984, p. 39–60.
 67. Kedem, O., and A. Katchalsky. Thermodynamic analysis of the permeability of biological membranes to non‐electrolytes. Biochim. Biophys. Acta 27: 229–246, 1958.
 68. Kedem, O., and A. Katchalsky. A physical interpretation of the phenomenological coefficients of membrane permeability. J. Gen. Physiol. 45 (1): 143–179, 1961.
 69. Ketterer, B., B. Neumcke, and P. Lauger. Transport mechanism of hydrophobic ions through lipid bilayer membranes. J. Membr. Biol. 5: 225–245, 1971.
 70. Kinne, R. K. H. Selectivity and direction: plasma membranes in renal transport. Am. J. Physiol. 260 (Renal Fluid Electrolyte Physiol. 29): F153–F162, 1991.
 71. Klingenberg, M. The ADP‐ATP translocation in mitocondria, a membrane potential controlled transport. J. Membr. Biol. 56: 97–105, 1980.
 72. Kohler, H. H., and K. Heckmann. Unidirectional fluxes in saturated single‐file pores of biological and artificial membranes. I. Pores containing no more than one vacancy. J. Theor. Biol. 79: 381–401, 1979.
 73. Kotyk, A., K. Janacek, and J. Koryta. Biophysical Chemistry of Membrane Functions. New York: John Wiley & Sons, 1988.
 74. Latorre, R., P. Labarca, and D. Naranjo. Surface charge effects on ion conduction in ion channels. Methods Enzymol. 207: 471–501, 1992.
 75. Lauger, P. Ion transport through pores: a rate theory analysis. Biochim. Biophys. Acta 311: 423–441, 1973.
 76. Lauger, P. Diffusion‐limited ion flow through pores. Biochim. Biophys. Acta 455: 493–509, 1976.
 77. Lauger, P. Kinetic properties of ion carriers and channels. J. Membr. Biol. 57: 163–178, 1980.
 78. Lauger, P. Conformational transitions of ionic channels. In: Single Channel Recording, edited by B. Sakmann and E. Neher,. New York: Plenum Press, 1983.
 79. Lauger, P. Dynamics of ion transport systems in membranes. Physiol. Rev. 67: 1296–1331, 1987.
 80. Lauger, P. Electrogenic Ion Pumps. Sunderland, MA: Sinauer Associates, 1991.
 81. Lauger, P., W. Stephan, and E. Frehland. Fluctuation of barrier structure in ionic channels. Biochim. Biophys. Acta 602: 167–180, 1980.
 82. Leo, A. Hydrophobic parameter: measurement and calculation. Methods Enzymol. 202: 544–591, 1991.
 83. Levie, R. Mathematical modeling of transport of lipid soluble ions and ion carrier complexes through lipid bilayer membranes. Adv. Chem. Phys. 37: 99–137, 1978.
 84. Levie, R., and H. Moreira. Transport of ions of one kind through thin membranes. J. Membr. Biol. 9: 241–260, 1972.
 85. Levitt, D. G. A new theory of transport for cell membrane pores. I. General theory and application to red cell. Biochim. Biophys. Acta 373: 115–131, 1974.
 86. Levitt, D. G. Electrostatic calculations for an ion channel. I. Energy and potential profiles and interactions between ions. Biophys. J. 22: 209–219, 1978a.
 87. Levitt, D. G. Electrostatic calculations for an ion channel. II. Kinetic behavior of the gramicidin A channel. Biophys. J. 22: 221–248, 1978b.
 88. Levitt, D. G. Comparison of Nernst‐Plank and reaction rate models for multiply occupied channels. Biophys. J. 37 (3): 575–587, 1982.
 89. Levitt, D. G. Kinetics of movement in narrow channels. Curr. Topic. Membr. Transp. 21: 182–197, 1984.
 90. Levitt, D. G. Interpretation of biological ion channel flux data—Reaction‐rate versus continuum theory. Annu. Rev. Biophys. Biophys. Chem. 15: 29–57, 1986.
 91. Levitt, D. G. General continuum theory for multiion channel. I. Theory. Biophys. J. 59: 271–277, 1991a.
 92. Levitt, D. G. General continuum theory for multiion channel. II. Application to acetylcholine channel. Biophys. J. 59: 278–288, 1991b.
 93. Lew, V. L., H. G. Ferreira, and T. Moura. The behaviour of transporting epithelial cells. I. Computer analysis of a basic model. Proc. R. Soc. Lond [Biol.] 206: 53–83, 1979.
 94. Lew, V. L., C. J. Freeman, O. E. Ortiz, and R. M. Bookchin. A mathematical model of the volume, pH, and ion content regulation in reticulocytes. Application to the pathophysiology of sickle cell dehydration. J. Clin. Invest. 87: 100–112, 1991.
 95. Lieb, W. R. A kinetic approach to transport studies. In: Red Cell Membranes—A Methodological Approach, edited by J. C. Ellory and J. D. Young,. New York: Academic Press, 1982, p. 135–154.
 96. Lieb, W. R., and W. D. Stein. Non‐Stokesian nature of transverse diffusion within human red cell membranes. J. Membr. Biol. 92: 111–119, 1986.
 97. Lipschutz, S. Schaum's Outline of Theory and Problems of Linear Algebra. New York: McGraw‐Hill, 1991.
 98. Macey, R. I. Mathematical models of membrane transport processes. In: Physiology of Membrane Disorders, edited by T. E. Andreoli, J. F. Hoffman, D. D. Fanestil, and S. G. Schultz,. New York: Plenum, 1986, p. 111–131.
 99. Macey, R. I., and L. W. Yousef. Osmotic stability of red cells in renal circulation requires rapid urea transport. Am. J. Physiol. 254 (Cell Physiol 23): C699–C674, 1988.
 100. Macgillivary, A. D., and D. Hare. Applicability of Goldman's constant field assumption to biological systems. J. Theor. Biol. 25: 113–126, 1969.
 101. Mauro, A. Some properties of ionic and nonionic semipermeable membranes. Circulation 21: 845–854, 1960.
 102. Mauro, A. The role of negative pressure in osmotic equilibrium and osmotic flow. In: Water Transport across Epithelia, Alfred Benzon Symposium 15, edited by H. H. Ussing, N. A. Lassen, and O. Sten Knudsen. Copenhagen: Munksgaard, 1981, p. 107–119.
 103. McLaughlin, S. Electrostatic potentials at membrane‐solution interfaces. Curr. Top. Membr. Transp. 9: 71–95, 1977.
 104. McLaughlin, S. The electrostatic properties of membranes. Annu. Rev. Biophys. Biophys. Chem. 18: 113–136, 1989.
 105. Moura, T. F., R. I. Macey, D. Y. Chien, D. M. Karan, and H. Santos. Thermodynamics of all‐or‐none water channel closure in red cells. J. Membr. Biol. 81: 105–111, 1984.
 106. Mullins, L. J. An electrogenic saga: Consequences of sodium‐calcium exchange in cardiac muscle. In: Electrogenic Transport: Fundamental Principles of Physiological Implications, edited by M. P. Blaustein and M. Lieberman,. New York: Raven Press, 1984, p. 161–179.
 107. Mullins, L. J., and K. Noda. The influence of sodium‐free solutions on the membrane potential of frog muscle fibers. J. Gen. Physiol. 47: 117–132, 1963.
 108. Naranjo, D., R. Latorre, D. Cherbavaz, P. McGill, and M. F. Schumaker. A simple model for surface charge on ion channel proteins. Biophys. J. 66: 59–70, 1994.
 109. Neumcke, B., and P. Lauger. Nonlinear electrical effects in lipid bilayer membranes. II. Integration of the generalized Nernst‐Planck equations. Biophys. J. 9: 1160–1170, 1969.
 110. Noble, D., and G. Bett. Reconstructing the heart: a challenge for integrative physiology, Cardiovasc. Res. 27: 1701–1712, 1993.
 111. Patlak, C. S. Contributions to the theory of active transport. Bull. Math. Biophys. 18: 271–315, 1956.
 112. Patlak, C. S. Derivation of an equation for the diffusion potential. Nature 188 (4754): 944–945, 1960.
 113. Patlak, C. S., D. A. Goldstein, and J. F. Hoffman. The flow of solute and solvent across a two‐membrane system. J. Theor. Biol. 5: 426–442, 1963.
 114. Patlak, C. S., and S. I. Rapoport. Use of transient and steady‐state measurements of the unidirectional flux ratio for the determination of the free energy change of chemical reactions and active transport systems. Bull. Math. Biol. 42: 529–537, 1980.
 115. Purcell, E. M. Electricity and Magnetism. New York: McGraw‐Hill, 1985.
 116. Reif, F. Fundamentals of Statistical and Thermal Physics. New York: McGraw‐Hill, 1965.
 117. Robinson, R. A., and R. H. Stokes. Electrolyte Solutions. London: Butterworth, 1965.
 118. Roux, B., and M. Karplus. Ion transport in a gramicidin‐like channel: dynamics and mobility. J. Phys. Chem. 95: 4856–4868, 1991a.
 119. Roux, B., and M. Karplus. Ion transport in a model gramicidin channel. Structure and thermodynamics. Biophys. J. 59: 961–981, 1991b.
 120. Roux, B., and M. Karplus. Ion transport in the gramicidin channel: Free energy of the solvated right‐handed dimer in a model membrane. J. Am. Chem. Soc. 115: 3250–3262, 1993.
 121. Roux, B., and M. Karplus. Molecular dynamics simulations of the gramicidin channel. Annu. Rev. Biophys. Biomol. Struct. 23: 731–761, 1994.
 122. Salhany, J. M. 1990. Erythrocyte Band 3 Protein. Boca Raton, FL: CRC Press, 1990.
 123. Schultz, S. G. Basic Principles of Membrane Transport. London: Cambridge University Press, 1980.
 124. Spangler, S. G. Expansion of the constant field equation to include both divalent and monovalent ions. Alabama J. Med. Sci. 9: 218–223, 1972.
 125. Stein, W. D. Transport and Diffusion across Cell Membranes. San Diego: Academic Press, 1986.
 126. Stein, W. D. Channels, Carriers, and Pumps: An Introduction to Membrane Transport. San Diego: Academic Press, 1990.
 127. Sten‐Knudsen, O. Passive transport processes. In: Membrane Transport in Biology, edited by G. Giebisch, D. C. Tosteson, and H. H. Ussing,. Berlin: Springer‐Verlag, 1978, p. 5–114.
 128. Strieter, J., J. L. Stephenson, G. Giebisch, and A. M. Weinstein. A mathematical model of the rabbit cortical collecting tubule. Am. J. Physiol. 263 (Renal Fluid Electrolyte Physiol. 32): F1063–F1075, 1992.
 129. Strieter, J., J. L. Stephenson, L. G. Palmer, and A. M. Weinstein. Volume‐activated chloride permeability can mediate cell volume regulation in a mathematical model of a tight epithelium. J. Gen. Physiol. 96 (2): 319–344, 1990.
 130. Stryer, L. Biochemistry. New York: W. H. Freeman, 1988.
 131. Teorell, T. Transport processes and electrical phenomena in ionic membranes. Prog. Biophys. Biophys. Chem. 3: 305–369, 1953.
 132. Turner, R. J., and A. Moran. Futher studies of proximal tubular brush border membrane. D‐glucose transport heterogeneity. J. Membr. Biol. 70: 37–45, 1982a.
 133. Turner, R. J. and A. Moran. Heterogeneity of sodium dependent D‐glucose transport sites along the proximal tubule: evidence from vesicle studies. Am. J. Physiol 242 (Renal Fluid Electrolyte Physiol. 11): F406–F414, 1982b.
 134. Tyrrell, H. J. V., and K. R. Harris. Diffusion in Liquids: A Theoretical and Experimental Study. London, Butterworths, 1984.
 135. Urban, B. W., and S. B. Hladky. Ion transport in the simplest single file pore. Biochim. Biophys. Acta 554: 410–429, 1979.
 136. Ussing, H. H. Distinction by means of tracers between active transport and diffusion. Acta Physiol. Scand. 19: 43–56, 1949.
 137. Wall, F. T. Chemical Thermodynamics: A Course of Study. San Francisco: W. H. Freeman, 1974.
 138. Walter, A., and J. Gutknecht. Permeability of small nonelectrolytes through lipid bilayer membranes. J. Membr. Biol. 90: 207–217, 1986.
 139. Weinstein, A. M. Chloride transport in a mathematical model of the rat proximal tubule. Am. J. Physiol. 263 (Renal Fluid Electrolyte Physiol. 32): F784–F798, 1992.
 140. 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.
 141. Wieth, J. O., O. S. Andersen, J. Brahm, P. J. Bjerrum, and J. R. Borders. Chloride‐bicarbonate exchange in red blood cells: physiology of transport and chemical modification of binding sites. Philos. Trans. R. Soc. Lond. Biol. 299: 383–399, 1982.
 142. Wright, E. M., and N. Bindslev. Thermodynamic analysis of nonelectrolyte permeation across the toad urinary bladder. J. Membr. Biol. 29: 289–312, 1976.
 143. Wright, E. M., and J. M. Diamond. Patterns of nonelectrolyte permeability. Proc. R. Soc. Lond. [Biol.] 172: 227–271, 1969.
 144. Wright, E. M., A. P. Smulders, and J. M. Tormey. The role of the lateral intracellular spaces and solute polarization effect in the passive flow of water across the rabbit gallbladder. J. Membr. Biol. 7: 198–219, 1972.
 145. Young, H. D. University Physics. Reading, MA: Addison‐Wesley, 1992.
 146. Yousef, L. W. and R. I. Macey. A method to distinguish between pore and carrier kinetics applied to urea transport across the erythrocyte membrane. Biochim. Biophys. Acta 984: 281–288, 1989.
 147. Zemansky, M. W., and R. H. Dittman. Heat and Thermodynamics: An Intermediate Textbook. New York: McGraw‐Hill, 1981.

Contact Editor

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

* Required Field

How to Cite

Robert I. Macey, Teresa F. Moura. Basic Principles of Transport. Compr Physiol 2011, Supplement 31: Handbook of Physiology, Cell Physiology: 181-259. First published in print 1997. doi: 10.1002/cphy.cp140106