Comprehensive Physiology Wiley Online Library

Physical Chemistry of Bile

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Structural and Surface Properties of Bile Salts
1.1 Molecular Structure
1.2 Crystal Structure
1.3 Surface Chemistry
1.4 Hydrophobic‐Hydrophilic Balance
2 Aqueous Solution Properties of Bile Salts
2.1 Solubility in Water
2.2 Solubility in Organic Solvents
2.3 Ionization Behavior
2.4 Aggregation Behavior
3 Bile Salt–Lecithin Interactions
3.1 Composition of Lecithins in Bile
3.2 Bile Salt–Phospholipid–Water Systems
4 Cholesterol
4.1 Solubility in Aqueous Media
4.2 Cholesterol‐Bile Salt Interactions
4.3 Cholesterol‐Phospholipid Interactions
4.4 Quaternary System: Bile Salt‐Phospholipid‐Cholesterol‐Water
5 Bile
5.1 Normal Bile
5.2 Metastable Bile
5.3 Cholesterol Crystal Formation in Bile
5.4 Secretion of Bile from the Hepatocyte
Figure 1. Figure 1.

Molecular structure of common bile acids showing common steroid ring and side‐chain structure. Hydroxyl group(s) location and orientation are given for each bile acid.

Figure 2. Figure 2.

Crystal structure of bile acids and salts. A: stereoview of the cell of lithocholic acid: α‐axis projection with c‐axis vertical. Oxygen molecules are blackened. B: stereoview of chenodeoxycholic acid; b‐axis projection. C: stereoview of unit cell of deoxycholic acid: H2O, showing channel filled with H2O molecules. Oxygens of H2O molecules are blackened. D: crystal packing of DCA: Rb viewed along b‐axis, showing planar wavy bilayer patterns. Broken lines, hydrogen bonds; large closed circles, Rb+, smaller closed circles, oxygen from H2O; small open circles, oxygen from the bile acid. E: crystal packing of unit cell of Na‐cholate‐monohydrate b‐axis projection. Wavy bilayer pattern is illustrated. Broken lines, hydrogen bonds; larger closed circles, Na+; smaller closed circles, carbon atoms; open circles, oxygen atoms. F: stereoview of bilayer packing of calcium cholate chloride heptahydrate. Closed circles of cholate molecules represent oxygen atoms of hydroxyl and carboxyl groups. Closed circles between cholate molecules represent Ca2+. G: stereoview of bilayer packing of sodium cholate monohydrate. Closed circles of cholate molecules represent oxygen atoms of hydroxyl and carboxyl groups. Closed circles between molecules represent H2O. H: stereoview of bilayer packing of rubidium deoxycholate. Closed circles of deoxycholate molecules represent oxygen atoms of hydroxyl and carboxyl groups. Closed circles between molecules represent H2O.

A from Arora et al. 4; B from Lindley et al. 68; C from Tang et al. 128; D from Coiro et al. 30; E from Cobbledick and Einstein 29; F–H from Hogan et al. 48a)
Figure 3. Figure 3.

Schematic of surface configuration of cholanic acid and hydroxylate derivatives. Maximum area per molecule was taken from first inflection point of compression isotherms; minimum area was taken from collapse point. Pressure at film collapse and state of film are given. Collapse pressure for chenodeoxycholic acid in 3 M NaCl low pH was determined to be ∼27 dyn/cm, similar to that of deoxycholic acid.

Data from Carey 20 and Small 113 and references therein
Figure 4. Figure 4.

Reverse‐phase high‐performance liquid chromatography (HPLC) chromatographs of mixture of bile salts. For a given conjugation state, mobility decreases in the order UDCA > CA > CDCA > DCA. Column used was an Ultrasphere ODS 250 X 4.6 mm, with a mobile phase of 75% MeOH‐25% 0.005 M KH2PO4/H3PO3, pH 5.0. TUDC, tauroursodexoycholate; TC, taurocholate; TCDC, taurochenodeoxycholate; TDC, taurodeoxycholate; GUDC, glycoursodeoxycholate; GC, glycocholate; GCDC, glycochenodeoxycholate; GDC, glycodeoxycholate; UDC, ursodeoxycholate; C, cholate; CDC, chenodeoxycholate; DC, deoxycholate. OD, optical density.

From Armstrong and Carey 3
Figure 5. Figure 5.

Hypothetical titration curve for solutions of free bile salts or for glycine conjugates. W, first equivalence point where titration of bile salt with hydrochloric acid commences. Y, last point where bile salt solution is in thermodynamic equilibrium as a single aqueous phase. T, Tyndall effect noted in this region of titration curve. X, point where precipitation of bile acid crystals commences. X', equilibrium pH at point of bile acid precipitation. Z, second equivalence point where titration of bile salt with hydrochloric acid is complete. TOT, total amount of acid required to complete the titration. HA, amount of acid added from first equivalence point W to point Y, which represents maximum solubility of bile acid (HA) in bile salt solution (A‐).

From Small 113
Figure 6. Figure 6.

Titration curves for cholic acid (CA) in several molecular environments. The pH is changed by adding HCl or NaOH, and chemical shift of carboxyl carbon (C‐24) is monitored by 13C NMR. Aqueous monomeric CA (.2 mM below its CMC and solubility limit). Aqueous micellar CA 116 mM). CA‐sodium taurocholate mixed micelles 1:7 mol/mol). CA‐BSA (bovine serum albumin) complexes 2:1 mol/mol). CA–egg yolk lecithin vesicles 1:20 mol/mol), major peak from CA in outer monolayer of vesicle, minor peak from CA in inner monolayer of vesicle.

Adapted from Cabral et al. 15
Figure 7. Figure 7.

Schematic of structure of bile salt micelles. Molecules of primary micelles associate via hydrophobic regions with aggregation numbers from 2 to 10. At higher concentrations some bile salts form larger aggregates, secondary micelles, stabilized by hydrophobic associations and hydrogen bonding.

From Small 112, printed with permission from Advances in Chemistry Series, copyright 1968, American Chemical Society
Figure 8. Figure 8.

Lecithin–sodium cholate–water diagram. W, H2O; L, lecithin; NaC, sodium cholate. Along side W‐L, numbers indicate H2O wt percentage; along side L‐NaC, numbers indicate lecithin wt percentage; along side NaC‐W, numbers indicate NaC wt percentage. Solid lines, well‐defined boundaries (+L2%); broken lines, less well‐defined frontiers. I, Zone of neat phase. II, Zone of cubic phase. III, Zone of middle phase. IV, Zone of isotropic phase. 1, Zone of separation of neat phase and NaC crystals. 2, Zone of separation of cubic phase and NaC crystals. 3, Zone of separation of middle phase and isotropic solution. 4, Zone of separation of neat phase and cubic phase. 5, Zone of separation of cubic phase and middle phase. 6, Zone of separation of neat phase and middle phase. 7, Zone of separation of neat phase and isotropic phase. 8, Zone of separation of isotropic solution and NaC crystals, a–e, Zones of separation of 3 phases whose compositions for each zone are indicated by apices of the triangles. Points A–H, J, K, M, O–Q are discussed in text. Right side of diagram, molar ratios of lecithin to NaC.

Figure 9. Figure 9.

Lecithin–sodium cholate water diagram at high‐H2O concentrations. NaC, sodium cholate; W, H2O. Left: larger numbers, wt%, smaller numbers, concentrations (mM). Parts of single‐phase zones I, III, and IV are shown, as well as 2‐phase regions 3, 6, and 7. Approximate tie lines in regions 6 and 7 and 3‐phase region are illustrated. Right, Bragg spacings (D) from lamellar and hexagonal lattices. These lattice parameters are given in angstroms along high‐H2O phase boundaries of lamellar phase and hexagonal phase. Dashed line from P (critical micellar concentration of NaC in H2O) to O, separates micelles formed by dilution of hexagonal phase (below) from micelles formed by dissolution of NaC crystals (above).

Figure 10. Figure 10.

Lecithin–sodium cholate system at >90% H2O, illustrating different techniques for preparing mixtures. NaC, sodium cholate; W, H2O. NaC‐W axis: larger numbers, wt% NaC from 0% to 10%; smaller numbers, concentration (mM). Point P, critical micelle concentration of NaC in H2O (.52% or ∼12 mM). Lecithin axis: large numbers, wt% lecithin; small numbers, concentrations (mM). Zone IV has been divided into 2 regions: IVa and IVb. Line separating 2 regions is a straight line running from point P to point O (see Fig. 9 and indicates a change in micellar structure. Narrow 2‐phase region 3 and larger 2‐phase region 7 are spearated by 3‐phase region a. Heavy lines, various methods of preparing mixtures. Any mixture may be made up by coprecipitation method (e.g., points 1 on 10% solid line and 5% solid line; TS, total solutes). A mixture made up by coprecipitation may then be diluted with H2O or buffer and will follow line 2. If a coprecipitated mixture is diluted with a specific concentration of bile salt, e.g., 2%, it will follow line 3. A coprecipitated mixture such as that plotted on 5% line at 40 mM lecithin and 45 mM NaC may be placed in a dialysis bag and dialyzed against H2O, buffer, or a bile salt solution. Because lecithin does not escape from the dialysis bag and its concentration remains unchanged, the composition will move toward W‐lecithin side along a constant 40‐mM line, line 4. Concentration of dialysate (not shown) will be found along W‐NaC axis. Finally, any concentration of lecithin, eg., a 6% suspension (either coarse liposomes or unilamellar vesicles), may be diluted with a specific concentration of bile salt, for instance at 10%, as shown on line 5. Compositions will fall on line 6, depending on amount of 10% bile salt solution added.

Figure 11. Figure 11.

A: mean hydrodynamic radii () of sodium taurocholate‐lecithin solutions (L/NaTC) at various total lipid concentrations (.625%–10% at 24°C in 0.15 M NaCl). diverges as lecithinbile salt ratio approaches micellar‐phase limits, and divergence limits are very concentration dependent. Dashed line, radius of simple discoid micelle. Phase limits shown at bottom of each panel represent appropriate lecithin‐bile salt ratio for total lipid concentrations of a, 0.0625%; b, 1.25%; c, 2.5%; d, 5.0%; and e, 10.0%. values indicated at tops of certain curves represent size of bilayer vesicles formed in supersaturated systems. B: at 3 L‐NaTC ratios as a function of the total wt% of lipids. As the total percentage of lipids increases, decreases markedly and approaches the dimensions of the simple discoidal micelle (arrows at right) at high lipid concentrations.

A reprinted with permission from Mazer et al. 70. Copyright 1980 American Chemical Society
Figure 12. Figure 12.

Lecithin‐sodium taurocholate (NaTC) system at high H2O (W) content. A: divergence limit (see Fig. 11A) and coexistence boundary plotted on rectangular coordinates. B‐C: system plotted on triangular coordinates (note difference in scale). Dashed line, phase boundary between micellar phase and zone 3. Phase boundary is virtually same as divergence limit line. Point P (critical micellar concentration of NaTC in H2O) has a value of 3.2 mM; 2‐phase regions 3 and 7 are shown. Tie lines in region 7 and the 3‐phase region a are also shown.

A from Mazer et al. 70; B‐C, coexistence and divergence boundaries from Mazer et al. 70
Figure 13. Figure 13.

A: mean hydrodynamic radius . B: polydispersity index V of a mixed micellar stock solution sodium glycocholatelecithin [lecithin/sodium glycocholate = 0.85; total concentration (Ctot) = 5 g/dl; temperature = 20°C] as a function of dilution. Dilution factor of 1 corresponds to stock solution Ctot, and dilution factor of n corresponds to Ctot divided by n.

From Schurtenberger et al. 108. Reprinted with permission from Journal of Physical Chemistry, copyright 1985, American Chemical Society
Figure 14. Figure 14.

Lecithin‐sodium glycocholate‐water system in dilute region of zone 7. L, lecithin; NaGC, sodium glycocholate; W, water. Closed circles, composition of fluid within dialysis bag; open circles, corresponding compositions (connected by a line) of dialysate fluid. Lines connect aqueous phase (intermicellar concentration or dialysate) with mixture composition (closed circles) for a series of vesicles of different sizes [ (Å)]. Extensions of these lines intersect L‐NaGC border (as shown by closed circles on right) and give ratio of NaGC bound to lecithin vesicle. Rh, mean hydrodynamic radii.

Figure 15. Figure 15.

Rates of dissolution of multilamellar liposomes by bile salts (BS) of different concentration. Lecithin concentrations of 1.5 mM were diluted with an equal volume of bile salts at concentrations on horizontal axis. A: time for half dissolution (T1/2 is plotted as a function of bile salt concentration; B: complex rate constant k1 91 for the process is plotted. Order of dissolution rates, chenodeoxycholate (CDC) > deoxycholate (DC) > cholate (C) > ursodeoxycholate (UDC).

Figure 16. Figure 16.

Sodium taurocholate‐lecithin‐cholesterol phase diagram. Chol, cholesterol; NaTC, sodium taurocholate; PL, phospholipid. At 10% total lipid, micellar phase is bounded by solid line. Dotted lines, micellar boundaries with decreasing concentrations of total lipids. Liquid crystal phase is a phospholipid phase with varying amounts of cholesterol and small amounts of bile salts. Crystalline phase is cholesterol monohydrate. We suggest that a 3‐phase region cannot directly abut on a 1‐phase region, except at a point where the 2‐phase region on the right extends to point B. The 3‐phase region consists of 1) cholesterol crystals, 2) liquid crystals of phospholipid, cholesterol, and bile salt, and 3) a micellar solution of composition B.

Adapted from Carey et al. 24
Figure 17. Figure 17.

Phase diagrams of aqueous taurine‐conjugated ursodeoxycholic acid (TUDC)‐, cholic acid (TC)‐, or chenodeoxycholic acid (TCDC)‐lecithin‐cholesterol monohydrate at a total lipid concentration of 10 g/100 ml (.20 M Na+, pH 7.4, 37°C). Axes are expressed in mol%. Although phase diagrams show 3‐phase region abutting 1‐phase region, there must be a 2‐phase region, however small, separating them.

From Salvioli et al. 101
Figure 18. Figure 18.

Partial phase diagram for aqueous system sodium taurocholate‐lecithin‐cholesterol 20 g/100 ml, 0.15 M NaCl, 24°C). Shaded area, metastable region; open circles, maximum equilibrium cholesterol solubility; closed circles, metastable‐labile limit.

From Carey et al. 24
Figure 19. Figure 19.

Behavior of hypothetical biles of compositions a‐f as they transform from dilute hepatic bile to concentrated gallbladder bile. Total lipids for hepatic bile are ∼1.5% and for gallbladder bile 10%. Phase boundaries are from Carey et al. 24. BS, bile salt. Point B is the same as that discussed in Fig. 16. For further explanation, see text.

Figure 20. Figure 20.

Electron micrograph of a bile canaliculus of a bile salt‐depleted rat liver that shows the presence of vesicular material within the lumen (left) (X 32,800). Higher magnification (inset) illustrates more clearly unilamellar structure of vesicle and can be compared with structure of canalicular membrane. Mean diameter of these vesicles is 56 nm 30–85 nm). Similar vesicles are also apparent in fresh bile specimen of same animal (right). Bars, 0.25 μm.

From Ulloa et al. 130, © by Am. Assoc. of the Study of Liver Diseases, 1987
Figure 21. Figure 21.

Distribution and composition of biliary lipids in bile fractionated on a continuous metrizamide density gradient. Left, T‐tube human hepatic bile; right, rat hepatic bile. Histograms are mean +L SD of percentage of each lipid in total sample of 7 human biles and 4 rat biles. Total lipid composition of rat biles was 2.4 g/dl; total composition of human T‐tube biles was 1.7 g/dl. BS, bile salt; C, cholesterol; PL, phospholipid. Compositions of each of the density cuts are plotted below on triangular coordinates with appropriate line as determined by Carey et al. 24. Open triangle, mean composition of bile; solid circle, density gradient fraction. Human bile is supersaturated with respect to 1.75% total solids line, whereas rat bile is not supersaturated. Least‐dense fraction (fraction 1) of both human and rat bile contains vesicles and has a composition very different from other density cuts.

Data from Ulloa 130


Figure 1.

Molecular structure of common bile acids showing common steroid ring and side‐chain structure. Hydroxyl group(s) location and orientation are given for each bile acid.



Figure 2.

Crystal structure of bile acids and salts. A: stereoview of the cell of lithocholic acid: α‐axis projection with c‐axis vertical. Oxygen molecules are blackened. B: stereoview of chenodeoxycholic acid; b‐axis projection. C: stereoview of unit cell of deoxycholic acid: H2O, showing channel filled with H2O molecules. Oxygens of H2O molecules are blackened. D: crystal packing of DCA: Rb viewed along b‐axis, showing planar wavy bilayer patterns. Broken lines, hydrogen bonds; large closed circles, Rb+, smaller closed circles, oxygen from H2O; small open circles, oxygen from the bile acid. E: crystal packing of unit cell of Na‐cholate‐monohydrate b‐axis projection. Wavy bilayer pattern is illustrated. Broken lines, hydrogen bonds; larger closed circles, Na+; smaller closed circles, carbon atoms; open circles, oxygen atoms. F: stereoview of bilayer packing of calcium cholate chloride heptahydrate. Closed circles of cholate molecules represent oxygen atoms of hydroxyl and carboxyl groups. Closed circles between cholate molecules represent Ca2+. G: stereoview of bilayer packing of sodium cholate monohydrate. Closed circles of cholate molecules represent oxygen atoms of hydroxyl and carboxyl groups. Closed circles between molecules represent H2O. H: stereoview of bilayer packing of rubidium deoxycholate. Closed circles of deoxycholate molecules represent oxygen atoms of hydroxyl and carboxyl groups. Closed circles between molecules represent H2O.

A from Arora et al. 4; B from Lindley et al. 68; C from Tang et al. 128; D from Coiro et al. 30; E from Cobbledick and Einstein 29; F–H from Hogan et al. 48a)


Figure 3.

Schematic of surface configuration of cholanic acid and hydroxylate derivatives. Maximum area per molecule was taken from first inflection point of compression isotherms; minimum area was taken from collapse point. Pressure at film collapse and state of film are given. Collapse pressure for chenodeoxycholic acid in 3 M NaCl low pH was determined to be ∼27 dyn/cm, similar to that of deoxycholic acid.

Data from Carey 20 and Small 113 and references therein


Figure 4.

Reverse‐phase high‐performance liquid chromatography (HPLC) chromatographs of mixture of bile salts. For a given conjugation state, mobility decreases in the order UDCA > CA > CDCA > DCA. Column used was an Ultrasphere ODS 250 X 4.6 mm, with a mobile phase of 75% MeOH‐25% 0.005 M KH2PO4/H3PO3, pH 5.0. TUDC, tauroursodexoycholate; TC, taurocholate; TCDC, taurochenodeoxycholate; TDC, taurodeoxycholate; GUDC, glycoursodeoxycholate; GC, glycocholate; GCDC, glycochenodeoxycholate; GDC, glycodeoxycholate; UDC, ursodeoxycholate; C, cholate; CDC, chenodeoxycholate; DC, deoxycholate. OD, optical density.

From Armstrong and Carey 3


Figure 5.

Hypothetical titration curve for solutions of free bile salts or for glycine conjugates. W, first equivalence point where titration of bile salt with hydrochloric acid commences. Y, last point where bile salt solution is in thermodynamic equilibrium as a single aqueous phase. T, Tyndall effect noted in this region of titration curve. X, point where precipitation of bile acid crystals commences. X', equilibrium pH at point of bile acid precipitation. Z, second equivalence point where titration of bile salt with hydrochloric acid is complete. TOT, total amount of acid required to complete the titration. HA, amount of acid added from first equivalence point W to point Y, which represents maximum solubility of bile acid (HA) in bile salt solution (A‐).

From Small 113


Figure 6.

Titration curves for cholic acid (CA) in several molecular environments. The pH is changed by adding HCl or NaOH, and chemical shift of carboxyl carbon (C‐24) is monitored by 13C NMR. Aqueous monomeric CA (.2 mM below its CMC and solubility limit). Aqueous micellar CA 116 mM). CA‐sodium taurocholate mixed micelles 1:7 mol/mol). CA‐BSA (bovine serum albumin) complexes 2:1 mol/mol). CA–egg yolk lecithin vesicles 1:20 mol/mol), major peak from CA in outer monolayer of vesicle, minor peak from CA in inner monolayer of vesicle.

Adapted from Cabral et al. 15


Figure 7.

Schematic of structure of bile salt micelles. Molecules of primary micelles associate via hydrophobic regions with aggregation numbers from 2 to 10. At higher concentrations some bile salts form larger aggregates, secondary micelles, stabilized by hydrophobic associations and hydrogen bonding.

From Small 112, printed with permission from Advances in Chemistry Series, copyright 1968, American Chemical Society


Figure 8.

Lecithin–sodium cholate–water diagram. W, H2O; L, lecithin; NaC, sodium cholate. Along side W‐L, numbers indicate H2O wt percentage; along side L‐NaC, numbers indicate lecithin wt percentage; along side NaC‐W, numbers indicate NaC wt percentage. Solid lines, well‐defined boundaries (+L2%); broken lines, less well‐defined frontiers. I, Zone of neat phase. II, Zone of cubic phase. III, Zone of middle phase. IV, Zone of isotropic phase. 1, Zone of separation of neat phase and NaC crystals. 2, Zone of separation of cubic phase and NaC crystals. 3, Zone of separation of middle phase and isotropic solution. 4, Zone of separation of neat phase and cubic phase. 5, Zone of separation of cubic phase and middle phase. 6, Zone of separation of neat phase and middle phase. 7, Zone of separation of neat phase and isotropic phase. 8, Zone of separation of isotropic solution and NaC crystals, a–e, Zones of separation of 3 phases whose compositions for each zone are indicated by apices of the triangles. Points A–H, J, K, M, O–Q are discussed in text. Right side of diagram, molar ratios of lecithin to NaC.



Figure 9.

Lecithin–sodium cholate water diagram at high‐H2O concentrations. NaC, sodium cholate; W, H2O. Left: larger numbers, wt%, smaller numbers, concentrations (mM). Parts of single‐phase zones I, III, and IV are shown, as well as 2‐phase regions 3, 6, and 7. Approximate tie lines in regions 6 and 7 and 3‐phase region are illustrated. Right, Bragg spacings (D) from lamellar and hexagonal lattices. These lattice parameters are given in angstroms along high‐H2O phase boundaries of lamellar phase and hexagonal phase. Dashed line from P (critical micellar concentration of NaC in H2O) to O, separates micelles formed by dilution of hexagonal phase (below) from micelles formed by dissolution of NaC crystals (above).



Figure 10.

Lecithin–sodium cholate system at >90% H2O, illustrating different techniques for preparing mixtures. NaC, sodium cholate; W, H2O. NaC‐W axis: larger numbers, wt% NaC from 0% to 10%; smaller numbers, concentration (mM). Point P, critical micelle concentration of NaC in H2O (.52% or ∼12 mM). Lecithin axis: large numbers, wt% lecithin; small numbers, concentrations (mM). Zone IV has been divided into 2 regions: IVa and IVb. Line separating 2 regions is a straight line running from point P to point O (see Fig. 9 and indicates a change in micellar structure. Narrow 2‐phase region 3 and larger 2‐phase region 7 are spearated by 3‐phase region a. Heavy lines, various methods of preparing mixtures. Any mixture may be made up by coprecipitation method (e.g., points 1 on 10% solid line and 5% solid line; TS, total solutes). A mixture made up by coprecipitation may then be diluted with H2O or buffer and will follow line 2. If a coprecipitated mixture is diluted with a specific concentration of bile salt, e.g., 2%, it will follow line 3. A coprecipitated mixture such as that plotted on 5% line at 40 mM lecithin and 45 mM NaC may be placed in a dialysis bag and dialyzed against H2O, buffer, or a bile salt solution. Because lecithin does not escape from the dialysis bag and its concentration remains unchanged, the composition will move toward W‐lecithin side along a constant 40‐mM line, line 4. Concentration of dialysate (not shown) will be found along W‐NaC axis. Finally, any concentration of lecithin, eg., a 6% suspension (either coarse liposomes or unilamellar vesicles), may be diluted with a specific concentration of bile salt, for instance at 10%, as shown on line 5. Compositions will fall on line 6, depending on amount of 10% bile salt solution added.



Figure 11.

A: mean hydrodynamic radii () of sodium taurocholate‐lecithin solutions (L/NaTC) at various total lipid concentrations (.625%–10% at 24°C in 0.15 M NaCl). diverges as lecithinbile salt ratio approaches micellar‐phase limits, and divergence limits are very concentration dependent. Dashed line, radius of simple discoid micelle. Phase limits shown at bottom of each panel represent appropriate lecithin‐bile salt ratio for total lipid concentrations of a, 0.0625%; b, 1.25%; c, 2.5%; d, 5.0%; and e, 10.0%. values indicated at tops of certain curves represent size of bilayer vesicles formed in supersaturated systems. B: at 3 L‐NaTC ratios as a function of the total wt% of lipids. As the total percentage of lipids increases, decreases markedly and approaches the dimensions of the simple discoidal micelle (arrows at right) at high lipid concentrations.

A reprinted with permission from Mazer et al. 70. Copyright 1980 American Chemical Society


Figure 12.

Lecithin‐sodium taurocholate (NaTC) system at high H2O (W) content. A: divergence limit (see Fig. 11A) and coexistence boundary plotted on rectangular coordinates. B‐C: system plotted on triangular coordinates (note difference in scale). Dashed line, phase boundary between micellar phase and zone 3. Phase boundary is virtually same as divergence limit line. Point P (critical micellar concentration of NaTC in H2O) has a value of 3.2 mM; 2‐phase regions 3 and 7 are shown. Tie lines in region 7 and the 3‐phase region a are also shown.

A from Mazer et al. 70; B‐C, coexistence and divergence boundaries from Mazer et al. 70


Figure 13.

A: mean hydrodynamic radius . B: polydispersity index V of a mixed micellar stock solution sodium glycocholatelecithin [lecithin/sodium glycocholate = 0.85; total concentration (Ctot) = 5 g/dl; temperature = 20°C] as a function of dilution. Dilution factor of 1 corresponds to stock solution Ctot, and dilution factor of n corresponds to Ctot divided by n.

From Schurtenberger et al. 108. Reprinted with permission from Journal of Physical Chemistry, copyright 1985, American Chemical Society


Figure 14.

Lecithin‐sodium glycocholate‐water system in dilute region of zone 7. L, lecithin; NaGC, sodium glycocholate; W, water. Closed circles, composition of fluid within dialysis bag; open circles, corresponding compositions (connected by a line) of dialysate fluid. Lines connect aqueous phase (intermicellar concentration or dialysate) with mixture composition (closed circles) for a series of vesicles of different sizes [ (Å)]. Extensions of these lines intersect L‐NaGC border (as shown by closed circles on right) and give ratio of NaGC bound to lecithin vesicle. Rh, mean hydrodynamic radii.



Figure 15.

Rates of dissolution of multilamellar liposomes by bile salts (BS) of different concentration. Lecithin concentrations of 1.5 mM were diluted with an equal volume of bile salts at concentrations on horizontal axis. A: time for half dissolution (T1/2 is plotted as a function of bile salt concentration; B: complex rate constant k1 91 for the process is plotted. Order of dissolution rates, chenodeoxycholate (CDC) > deoxycholate (DC) > cholate (C) > ursodeoxycholate (UDC).



Figure 16.

Sodium taurocholate‐lecithin‐cholesterol phase diagram. Chol, cholesterol; NaTC, sodium taurocholate; PL, phospholipid. At 10% total lipid, micellar phase is bounded by solid line. Dotted lines, micellar boundaries with decreasing concentrations of total lipids. Liquid crystal phase is a phospholipid phase with varying amounts of cholesterol and small amounts of bile salts. Crystalline phase is cholesterol monohydrate. We suggest that a 3‐phase region cannot directly abut on a 1‐phase region, except at a point where the 2‐phase region on the right extends to point B. The 3‐phase region consists of 1) cholesterol crystals, 2) liquid crystals of phospholipid, cholesterol, and bile salt, and 3) a micellar solution of composition B.

Adapted from Carey et al. 24


Figure 17.

Phase diagrams of aqueous taurine‐conjugated ursodeoxycholic acid (TUDC)‐, cholic acid (TC)‐, or chenodeoxycholic acid (TCDC)‐lecithin‐cholesterol monohydrate at a total lipid concentration of 10 g/100 ml (.20 M Na+, pH 7.4, 37°C). Axes are expressed in mol%. Although phase diagrams show 3‐phase region abutting 1‐phase region, there must be a 2‐phase region, however small, separating them.

From Salvioli et al. 101


Figure 18.

Partial phase diagram for aqueous system sodium taurocholate‐lecithin‐cholesterol 20 g/100 ml, 0.15 M NaCl, 24°C). Shaded area, metastable region; open circles, maximum equilibrium cholesterol solubility; closed circles, metastable‐labile limit.

From Carey et al. 24


Figure 19.

Behavior of hypothetical biles of compositions a‐f as they transform from dilute hepatic bile to concentrated gallbladder bile. Total lipids for hepatic bile are ∼1.5% and for gallbladder bile 10%. Phase boundaries are from Carey et al. 24. BS, bile salt. Point B is the same as that discussed in Fig. 16. For further explanation, see text.



Figure 20.

Electron micrograph of a bile canaliculus of a bile salt‐depleted rat liver that shows the presence of vesicular material within the lumen (left) (X 32,800). Higher magnification (inset) illustrates more clearly unilamellar structure of vesicle and can be compared with structure of canalicular membrane. Mean diameter of these vesicles is 56 nm 30–85 nm). Similar vesicles are also apparent in fresh bile specimen of same animal (right). Bars, 0.25 μm.

From Ulloa et al. 130, © by Am. Assoc. of the Study of Liver Diseases, 1987


Figure 21.

Distribution and composition of biliary lipids in bile fractionated on a continuous metrizamide density gradient. Left, T‐tube human hepatic bile; right, rat hepatic bile. Histograms are mean +L SD of percentage of each lipid in total sample of 7 human biles and 4 rat biles. Total lipid composition of rat biles was 2.4 g/dl; total composition of human T‐tube biles was 1.7 g/dl. BS, bile salt; C, cholesterol; PL, phospholipid. Compositions of each of the density cuts are plotted below on triangular coordinates with appropriate line as determined by Carey et al. 24. Open triangle, mean composition of bile; solid circle, density gradient fraction. Human bile is supersaturated with respect to 1.75% total solids line, whereas rat bile is not supersaturated. Least‐dense fraction (fraction 1) of both human and rat bile contains vesicles and has a composition very different from other density cuts.

Data from Ulloa 130
References
 1. Admirand, W. H., and D. M. Small. The physico‐chemical basis of cholesterol gallstone formation in man. J. Clin. Invest. 47: 1045–1052, 1968.
 2. Angelico, M. Subselection of phospholipids during formation of native bile: physical‐chemical coupling of bile salts and phospholipids (Abstract). Clin. Res. 33: 538A, 1985.
 3. Armstrong, M. C., and M. C. Carey. The hydrophobic‐hydrophilic balance of bile salts. Inverse correlation between reverse phase high pressure liquid chromatographic mobilities and micellar cholesterol‐solubilizing capacities. J. Lipid Res. 23: 70–80, 1982.
 4. Arora, S. K., G. Germain, and J. P. Declercq. The crystal and molecular structure of lithocholic acid. Acta Crystallogr. Sect. B. Struct. Crystallogr. Cryst. Chem. 32: 415–419, 1982.
 5. Backer, J. M., and E. A. Dawidowicz. Mechanism of cholesterol exchange between phospholipid vesicles. Biochemistry 20: 3805–3810, 1981.
 6. Barnes, S., and J. M. Geckle. High resolution nuclear magnetic resonance spectroscopy of bile salts: individual proton assignments for sodium cholate in aqueous solution at 400 MHz. J. Lipid Res. 23: 161–170, 1982.
 7. Bennett, W. S., G. Ellington, and S. Kovack. Self‐association of phenolics and bile acid derivatives. Nature Lond. 214: 776–780, 1967.
 8. Birdi, K. S. Aggregation of bile salts (Na‐deoxycholate). Finn. Chem. Lett. 6–8: 142–146, 1982.
 9. Bloch, C. A., and J. B. Watkins. Determination of conjugated bile acids in human bile and duodenal fluid by reverse phase high pressure liquid chromatography. J. Lipid Res. 19: 510–513, 1978.
 10. Borgstrom, B. Studies of the phospholipids of human bile and small intestine. Acta Chem. Scand. 11: 749, 1957.
 11. Bourges, M., D. M. Small, and D. G. Dervichian. Biophysics of lipidic associations. II. The ternary systems. Cholesterol‐lecithin‐water. Biochim. Biophys. Acta 137: 157–167, 1967.
 12. Bourges, M., D. M. Small, and D. G. Dervichian. Biophysics of lipidic associations. III. The quaternary systems. Bile salt‐lecithin‐cholesterol‐water. Biochim. Biophys. Acta 144: 189–202, 1967.
 13. Bruckdirfer, K. R., and M. K. Sherry. The solubility of cholesterol and its exchange between membranes. Biochim. Biophys. Acta 769: 187–196, 1984.
 14. Burnstein, M. J., R. G. Ilson, C. N. Petrunka, R. D. Taylor, and S. M. Strasberg. Evidence of a potential nucleating factor in the gallbladder bile of patients with cholesterol gallstones. Gastroenterology 85: 801–807, 1983.
 15. Cabral, D. J., J. A. Hamilton, and D. M. Small. The ionization behavior of bile acids in different aqueous environments. J. Lipid Res. 27: 334–343, 1986.
 16. Cabral, D. J., D. M. Small, H. S. Lilly, and J. A. Hamilton. Transbilayer movement of bile acids in model membranes. Biochemistry 26: 1801–1804, 1987.
 17. Cabral, D. J., D. M. Small, and J. A. Hamilton. Exchange behavior of bile acids in phosphatidylcholine vesicles (Abstract). Biophys. J. 51: 540a, 1987.
 18. Campanelli, A. R., D. Ferro, E. Giglio, P. Imperatori, and V. Piacente. Thermal and x‐ray study of sodium deoxycholate crystal and fiber. Thermochim. Acta 67: 223–232, 1983.
 19. Cantafora, A., A. DiBiase, D. Alvaro, M. Angelico, M. Marin, and A. F. Attili. High performance liquid chromatographic analysis of molecular species of phosphatidylcholine‐development of quantitative assay and its application to human bile. Clin. Chim. Acta 134: 281–285, 1983.
 20. Carey, M. C. Critical tables for calculating the cholesterol saturation of native bile. J. Lipid Res. 19: 945–955, 1978.
 21. Carey, M. C. Physical‐chemical properties of bile acids and their salts. In: New Comprehensive Biochemistry: Sterols and Bile Acids, edited by J. Sjovall and H. Danielson. Amsterdam: Elsevier, 1985, p. 345–403.
 22. Carey, M. C., M. J. Armstrong, N. A. Mazer, H. Igimi, and G. Salvioli. Measurement of the hydrophilic‐hydrophobic balance of bile salt: correlation with physical‐chemical interactions. In: Bile Acids and Cholesterol in Health and Disease, edited by G. Paumgartner, A. Stiehl, and W. Gerok. Lancaster, UK: MTP, 1983, p. 31–42.
 23. Carey, M. C., and D. M. Small. Micellar properties of dihydroxy and trihydroxy bile salts: effects of counterion and temperature. J. Colloid Interface Sci. 31: 382–396, 1969.
 24. Carey, M. C., and D. M. Small. The physical chemistry of cholesterol solubility in bile: relationship to gallstone formation and dissolution in man. J. Clin. Invest. 61: 998–1026, 1978.
 25. Carey, M. D., D. M. Small, and C. M. Bliss. Lipid digestion and absorption. Annu. Rev. Physiol. 45: 651–677, 1983.
 26. Castellino, F. J., and B. N. Violand. 31P‐nuclear magnetic resonance and 31P[1H] nuclear Overhauser effect analysis of mixed egg phosphatidylcholine‐sodium taurocholate vesicles and micelles. Arch. Biochem. Biophys. 193: 543–550, 1979.
 27. Chang, Y., and J. R. Cardinal. Light scattering studies on bile acid salts. II. Pattern of self‐association of sodium deoxycholate, sodium taurodeoxycholate and sodium glycodeoxycholate in aqueous electrolyte solutions. J. Pharm. Sci. 67: 994–999, 1978.
 28. Christopher, F. Textbook of Surgery (9th ed.), edited by L. Davis. Philadelphia, PA: Saunders, 1968.
 29. Cistola, D. P., D. M. Small, and J. A. Hamilton. Ionization behavior of aqueous short‐chain carboxylic acids: a carbon‐13 NMR study. J. Lipid Res. 23: 795–799, 1982.
 30. Claffey, W. J., and R. T. Holzbach. Dimorphism in bile salt/lecithin mixed micelles. Biochemistry 20: 415–418, 1981.
 31. Cobbledick, R. E., and F. W. B. Einstein. The structure of 3α,7α,12α‐trihydroxy‐5β‐cholan‐24‐oate monohydrate (sodium cholate monohydrate). Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 36: 287–292, 1980.
 32. Coiro, Y. M., E. Giglio, S. Morosette, and A. Palleschi. A monoclinic phase of the deoxycholic acid rubidium salt. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 36: 1478–1480, 1980.
 33. Conte, G., R. DiBlasi, E. Giglio, A. Parrett, and N. V. Pavel. Nuclear magnetic resonance and x‐ray studies on micellar aggregates of sodium deoxycholate. J. Phys. Chem. 88: 5720–5724, 1984.
 34. Delacroix, D. L., H. J. F. Hodgson, A. McPherson, C. Dive, and J. P. Vaerman. Selective transport of polymeric immunoglobulin A in bile. J. Clin. Invest. 70: 230–241, 1982.
 35. DeMaria, P., A. Fini, and A. Roda. Chemical properties of bile acids. I. Thermodynamic dissociation constants of some cholanic acid derivatives in 50 weight percent aqueous methanol. Gazz. Chim. Ital. 111: 95–97, 1981.
 36. Dowling, R. H., E. Mack, and D. M. Small. Effects of controlled interruption of the enterohepatic circulation of bile salts by biliary diversion and by ileal resection on bile salt secretion, synthesis and pool size in the Rhesus monkey. J. Clin. Invest. 49: 232–242, 1970.
 37. Dowling, R. J., E. Mack, and D. M. Small. Biliary lipid secretion and bile composition following acute and chronic interruption of the enterohepatic circulation in the Rhesus monkey. J. Clin. Invest. 50: 1917–1926, 1971.
 38. Dowling, R. H., and D. M. Small. The effect of pH on the solubility of varying mixtures of free and conjugated bile salts in solution. Gastroenterology 54: 1291, 1968.
 39. Ekwall, P., T. Rosendahl, and N. Lofman. Studies on bile acid salt solutions. Acta Chem. Scand. 11: 590–598, 1957.
 40. Ekwall, P., T. Rosendahl, and A. Sten. Studies on bile acid salt solutions. II. The solubility of cholic acid in sodium cholate solutions and that of deoxycholate in sodium deoxycholate solutions. Acta Chem. Scand. 12: 1622–1633, 1958.
 41. Fini, A., A. Roda, and P. DeMaria. Chemical properties of bile acids. Part 2. pKa values in H2O and aqueous methanol of some hydroxy bile acids. Eur. J. Chem. 17: 467–470, 1982.
 42. Fini, A., A. Roda, R. Fugazza, and B. Grigolo. Chemical properties of bile acids. III. Bile acid structure and solubility in water. J. Solution Chem. 14: 595–603, 1985.
 43. Fromhertz, P. Lipid‐vesicle structure: size control by edgeactive agents. Chem. Phys. Lipids 94: 259–266, 1983.
 44. Fromhertz, P., and D. Ruppel. Lipid vesicle formation: the transition from open disk to closed shells. FEBS Lett. 179: 155–159, 1985.
 45. Gallinger, S., P. R. C. Harvey, C. N. Petrunka, and S. M. Strasberg. The effect of binding of ionized calcium on the in vitro nucleation of cholesterol and calcium bilirubinate in human gallbladder bile. Gut 27: 1382–1386, 1986.
 46. Gallinger, S., P. Robert, C. Harvey, C. N. Petrunka, R. G. Ilson, and S. M. Strasberg. Biliary proteins and the nucleation defect in cholesterol cholelithiasis. Gastroenterology 91: 867–875, 1987.
 47. Gallinger, S., R. D. Taylor, P. R. C. Harvey, C. N. Petrunka, and S. M. Strasberg. Effect of mucous glycoprotein on nucleation time of human bile. Gastroenterology 89: 648–658, 1985.
 48. Gilat, T., and G. J. Sömjen. Cholesterol solubility in human bile. J. Clin. Gastroenterol. In press.
 49. Haberland, M. E., and J. A. Reynolds. Self‐association of cholesterol in aqueous solution. Proc. Natl. Acad. Sci. USA 70: 2313–2316, 1973.
 50. Halpern, Z., M. A. Dudley, A. Kibe, M. P. Lynn, A. C. Breuer, and R. T. Holzbach. Rapid vesicle formation and aggregation in abnormal human biles. A time‐lapse videoenhanced contrast microscopy study. Gastroenterology 90: 875–885, 1986.
 51. Harvey, P. R. C., C. A. Rupar, S. Gallinger, C. N. Petrunka, and S. M. Strasberg. Quantitative and qualitative comparison of gallbladder mucus glycoprotein from patients with and without gallstones. Gut 27: 374–381, 1986.
 52. Harvey, P. R. C., D. Taylor, C. N. Petrunka, A. D. Murray, and S. M. Strasberg. Quantitative analysis of major, minor and trace elements in gallbladder bile of patients with and without gallstones. Hepatology Baltimore 5: 129–132, 1985.
 53. Hegardt, F. G., and H. Dam. The solubility of cholesterol in aqueous solutions of bile salts and lecithin. Z. Ernaehrungswiss. 10: 223–233, 1971.
 54. Hepatology Baltimore 4, Suppl.: 1S–252S, 1984.
 55. Hofmann, A. F. Animal models of calcium cholelithiasis. Hepatology Baltimore 4, Suppl.: 209S–211S, 1984.
 56. Hogan, A., S. E. Ealick, C. E. Bugg, and S. Barnes. Aggregation patterns of bile salts: crystal structure of calcium cholate chloride heptahydrate. J. Lipid Res. 25: 791–798, 1984.
 57. Holan, K. R., R. T. Holzbach, R. E. Hermann, A. C. Cooperman, and W. J. Claffey. Nucleation time: a key factor in the pathogenesis of cholesterol gallstone disease. Gastroenterology 77: 611–617, 1979.
 58. Holzbach, R. T., and C. Corbusier. Liquid crystals and cholesterol nucleation during equilibration in supersaturated bile analogues. Biochim. Biophys. Acta 528: 436–444, 1978.
 59. Holzbach, R. T., A. Kibe, E. Thiel, J. H. Howell, M. Marsh, and R. E. Hermann. Biliary proteins: unique inhibitors of cholesterol crystal nucleation in human gallbladder bile. J. Clin. Invest. 73: 35–45, 1984.
 60. Holzbach, R. T., M. Marsh, M. Olsezewski, and K. Holan. Cholesterol solubility in bile: evidence that supersaturated bile is frequent in healthy man. J. Clin. Invest. 52: 1467–1479, 1973.
 61. Hunt, G. R. A. A comparison of triton X‐100 and the bile salt taurocholate as micellar ionophones or fusogenes in phospholipid vesicular membranes. A 1H NMR method using the lanthanide probe ion Pr3+. FEBS Lett. 119: 132–136, 1980.
 62. Hunt, G. R. A., and K. Jawaharlal. A 1H‐NMR investigation of the mechanism for the ionophore activity of the bile salts in phospholipid vesicular membranes and the effects of cholesterol. Biochim. Biophys. Acta 601: 678–684, 1980.
 63. Igimi, H., and M. C. Carey. pH‐solubility relations of chenodeoxycholic and ursodeoxycholic acids: physical‐chemical basis for dissimilar solution and membrane phenomena. J. Lipid Res. 21: 72–90, 1980.
 64. Iser, J. H., G. M. Murphy, and R. H. Dowling. Speed of change in biliary lipids and bile acids with chenodeoxycholic acid—is intermittent therapy feasible? Gut 18: 7–15, 1977.
 65. Kibe, A., M. A. Dudley, Z. Halpern, M. P. Lynn, A. C. Breuer, and R. T. Holzbach. Factors affecting cholesterol monohydrate crystal nucleation time in model systems of supersaturated bile. J. Lipid Res. 26: 1102–1111, 1985.
 66. Kolehmainen, E. Solubilization of aromatics in aqueous bile salts. I. Benzene and alkylbenzenes in sodium cholate: 1H NMR study. J. Colloid Interface Sci. 105: 273–277, 1985.
 67. Kramer, R. M., H. J. Hasselbach, and G. Semenza. Rapid transmembrane movement of phosphatidylcholine in small unilamellar lipid vesicles formed by detergent removal. Biochim. Biophys. Acta 643: 233–242, 1985.
 68. Kratohvil, J. P., and H. T. Dellicolli. Micellar properties of bile salts. Sodium taurodeoxycholate and sodium glycodeoxycholate. Can. J. Biochem. 46: 945–952, 1968.
 69. Kratohvil, J. P., W. P. Hsu, M. A. Jacobs, T. M. Aminabhavi, and Y. Mukunoki. Concentration‐dependent aggregation patterns of conjugated bile salts in aqueous sodium chloride solutions. Colloid Polym. Sci. 261: 781–785, 1983.
 70. Lack, L., and I. M. Weiner. Intestinal bile salt transport: structure‐activity relationships and other properties. Am. J. Physiol. 210: 1142–1152, 1966.
 71. Lake, M., and D. T. Organisciak. Determination of the composition of mixed micelles of bile salts by kinetic dialysis. Lipids 19: 553–557, 1984.
 72. LaRusso, N. F. Proteins in bile: how they get there and what they do. Am. J. Physiol. 247 (Gastrointest. Liver Physiol. 10): G199–G205, 1984.
 73. Lecuyer, H., and D. G. Dervichian. Structure of aqueous mixtures of lecithin and cholesterol. J. Mol. Biol. 45: 39–57, 1969.
 74. Lee, S. P., M. C. Carey, and J. T. LaMont. Aspirin prevention of cholesterol gallstone formation in the prairie dog. Science Wash. DC 211: 1420, 1981.
 75. Lee, S. P., J. T. LaMont, and M. C. Carey. Role of gallstone mucin hypersecretion in the evolution of cholesterol gallstones: studies in the prairie dog. J. Clin. Invest. 67: 1712–1723, 1981.
 76. Leibfritz, D., and J. D. Roberts. Nuclear magnetic resonance spectroscopy. Carbon‐13 spectra of cholic acids and hydrocarbons included in sodium deoxycholate solutions. J. Am. Chem. Soc. 95: 4996–5003, 1973.
 77. Lichtenberg, D., Y. Zilberman, P. Greenzaid, and S. Zamir. Structural and kinetic studies on the solubilization of lecithin by sodium deoxycholate. Biochemistry 18: 3517–3525, 1979.
 78. Lindheimer, M., J. C. Montet, J. Molenat, R. Bontemps, and B. Brun. Ionic self‐diffusion of various bile salts. J. Chim. Phys. 78: 447–455, 1981.
 79. Lindley, P. F., M. M. Mahmoud, F. E. Watson, and W. A. Jones. The structure of chenodeoxycholic acid, C24H40O4. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 36: 1893–1897, 1980.
 80. Lindman, B., N. Kamenka, H. Fabre, J. Ulmius, and T. Weiloch. Aggregation, aggregate composition and dynamics in aqueous sodium cholate solutions. J. Colloid Interface Sci. 73: 556–565, 1980.
 81. MacPherson, B. R., R. S. Pemsingh, and G. W. Scott. Experimental cholelithiasis in the ground squirrel. Lab. Invest. 56: 138–145, 1987.
 82. Mazer, N. A., G. B. Benedek, and M. C. Carey. Quasielastic light‐scattering studies of aqueous biliary lipid systems. Mixed micelle formation in bile salt‐lecithin solutions. Biochemistry 19: 601–615, 1980.
 83. Mazer, N. A., and M. C. Carey. Quasi‐elastic light scattering studies of aqueous biliary lipid systems. Cholesterol solubilization and precipitation in model bile solutions. Biochemistry 22: 426–442, 1983.
 84. Mazer, N. A., M. C. Carey, R. F. Kwasnick, and G. B. Benedek. Quasi‐elastic light scattering studies of aqueous biliary lipid systems. Size, shape and thermodynamics of bile salt micelles. Biochemistry 18: 3064–3075, 1979.
 85. Mazer, N. A., P. Schurtenberger, M. C. Carey, R. Presig, K. Weigand, and W. Kaniz. Quasi‐elastic light scattering studies of native bile from the dog: comparison with aggregation behavior of model biliary lipid systems. Biochemistry 23: 1994–2005, 1984.
 86. McLean, L. R., and M. C. Phillips. Mechanism of cholesterol and phospholipid exchange or transfer between unilamellar vesicles. Biochemistry 20: 2893–2900, 1981.
 87. Montet, J. C., and D. G. Dervichian. Solubilisation micellaire du cholesterol par les sels biliares et les lécithines extraits de la bile humaine. Biochimi Paris 53: 751–754, 1971.
 88. Moore, E. W. The role of calcium in the pathogenesis of gallstones: Ca++ electrode studies of model bile salt solutions and other biologic systems. Hepatology Baltimore 4, Suppl.: 228S–243S, 1984.
 89. Moore, E. W., L. Celic, and J. D. Ostrow. Interactions between ionized calcium and sodium cholate: bile salts are important buffers for prevention of calcium‐containing gallstones. Gastroenterology 83: 1079–1089, 1982.
 90. Mukidjam, E., S. Barnes, and G. A. Elgavish. NMR studies of the binding of sodium and calcium ions to the bile salts glycocholate and taurocholate in dilute solution, as probed by the paramagnetic lanthanide dysprosium. J. Am. Chem. Soc. 108: 7082–7089, 1986.
 91. Muller, K. Structural dimorphism of bile salt/lecithin mixed micelles. A possible regulatory mechanism for cholesterol solubility in bile? X‐ray structure analysis. Biochemistry 20: 404–414, 1981.
 92. Murata, Y., G. Suglhara, K. Fukushima, and M. Tanaka. Study of the micelle formation of sodium deoxycholate. Concentration dependence of carbon‐13 nuclear magnetic resonance chemical shift. J. Phys. Chem. 86: 4690–4694, 1982.
 93. Norton, D. A., and B. Haner. Crystal data (I) for some bile acid derivatives. Acta Crystallogr. Sect. B. Struct. Crystallogr. Cryst. Chem. 19: 477–479, 1965.
 94. O'Connor, C. J., B. T. Ch'ng, and R. G. Wallace. Studies in bile salt solutions. 1. Surface tension evidence for a stepwise aggregation model. J. Colloid Interface Sci. 95: 410–419, 1983.
 95. Olive, J., and D. G. Dervichian. Action d'une phospholipase sur la lécithine a l'état micellaire. Bull. Soc. Chim. Biol. 50: 1409–1418, 1968.
 96. Pal, S., A. R. Das, and S. P. Moulik. Interaction of bile salts (sodium cholate and sodium deoxycholate) with a nonionic surfactant (Triton X‐100) and polyethylene glycols. Indian J. Biochem. Biophys. 19: 295–300, 1982.
 97. Palmer, R. H. Bile salts and the liver. Prog. Liver Dis. 7: 221–242, 1982.
 98. Parris, N. A. Liquid chromatographic separation of bile acids. J. Chromatogr. 133: 273–279, 1977.
 99. Patton, G. M., S. B. Clark, J. M. Fasulo, and S. J. Robins. Utilization of individual lecithins in intestinal lipoprotein formation in the rat. J. Clin. Invest. 73: 231–240, 1984.
 100. Patton, G. M., S. J. Robins, J. M. Fasulo, and S. B. Clark. Influence of lecithin acyl chain composition on the kinetics of exchange between chylomicrons and high density lipoproteins. J. Lipid Res. 26: 1285–1293, 1985.
 101. Paul, R., M. K. Mathew, R. Narayanan, and P. Balaram. Fluorescent probe and NMR studies of the aggregation of bile salts in aqueous solutions. Chem. Phys. Lipids 25: 345–356, 1979.
 102. Plank, L., C. E. Dahl, and B. R. Ware. Effect of sterol incorporation on headgroup separation in liposomes. Chem. Phys. Lipids 36: 319–328, 1985.
 103. Rajagopalan, N., and S. Lindenbaum. The binding of Ca2+ to taurine and glycine‐conjugated bile salt micelles. Biochim. Biophys. Acta 711: 66–74, 1982.
 104. Rajagopalan, N., and S. Lindenbaum. Kinetics and thermodynamics of the formation of mixed micelles of egg phosphatidylcholine and bile salt. J. Lipid Res. 25: 135–147, 1984.
 105. Rajagopalan, N., M. Vadnere, and S. Lindenbaum. Thermodynamics of aqueous bile salt solutions: heat capacity, enthalpy and entropy of dilution. J. Solution Chem. 10: 785–801, 1981.
 106. Redinger, R. N., and D. M. Small. Bile composition, bile salt metabolism and gallstones. Arch. Intern. Med. 130: 618–630, 1972.
 107. Renshaw, P. F., A. S. Janoff, and K. W. Miller. On the nature of dilute aqueous cholesterol suspensions. J. Lipid Res. 24: 47–51, 1983.
 108. Robins, S. J., J. M. Fasulo, and G. M. Patton. Lipids of pigment gallstones. Biochim. Biophys. Acta 712: 21–25, 1982.
 109. Roda, A., A. F. Hofmann, and K. J. Mysels. The influence of bile salt structure and aggregation in aqueous solutions. J. Biol. Chem. 258: 6362–6370, 1983.
 110. Roe, J. M., and B. W. Barry. Measurement of critical micelle concentration by photon correlation spectroscopy. J. Colloid Interface Sci. 94: 580–583, 1983.
 111. Roe, J. M., and B. W. Barry. Bile salt association (cholate, deoxycholate, chenodeoxycholate and ursodeoxycholate) and interactions with aromatic alcohols (benzyl, 2‐phenylethanol, and 3‐phenylpropanol). J. Colloid Interface Sci. 107: 398–404, 1985.
 112. Ruppin, D. C., G. M. Murphy, R. H. Dowling, and the British Gallstone Study Group. Gallstone disease without gallstones—bile acid and bile lipid metabolism after complete gallstone dissolution. Gut 27: 559–566, 1986.
 113. Saad, H. Y., and W. I. Higuchi. Water solubility of cholesterol. J. Pharm. Sci. 54: 1205–1206, 1965.
 114. Saito, H., Y. Sugimoto, R. Tabeta, S. Suzuki, G. Izumi, M. Kodama, S. Toyoshima, and C. Nagata. Incorporation of bile acids of low concentration into model and biological membranes studied by 2H and 31P NMR. J. Biochem. Tokyo 94: 1877–1887, 1983.
 115. Salvioli, G., H. Igimi, and M. C. Carey. Cholesterol gallstone dissolution in bile. Dissolution kinetics of crystalline cholesterol monohydrate by conjugated chenodeoxycholatelecithin and conjugated ursodeoxycholate‐lecithin mixtures: dissimilar phase equilibria and dissolution mechanisms. J. Lipid Res. 24: 701–720, 1983.
 116. Sedaghat, S., and S. M. Grundy. Cholesterol crystals and the formation of cholesterol gallstones. N. Engl. J. Med. 302: 1274–1277, 1980.
 117. Sewell, R. B., S. J. T. Mao, T. Kawamoto, and N. F. LaRusso. Apolipoproteins of high, low, and very low density lipoproteins in human bile. J. Lipid Res. 24: 391–401, 1983.
 118. Shankland, W. The equilibrium and structure of lecithincholate mixed micelles. Chem. Phys. Lipids 4: 109–130, 1970.
 119. Shaw, R., and W. H. Elliott. Bile acids. XLVIII. Separation of conjugated bile acids by high‐pressure liquid chromatography. Anal. Biochem. 74: 273–281, 1976.
 120. Shaw, R., M. Rivetna, and W. H. Elliott. Bile acids. LXIII. Relationship between the mobility on reverse‐phase high performance liquid chromatography and the structure of bile acids. J. Chromatogr. 21: 347–361, 1980.
 121. Shurtenberger, P., and B. Lindman. Coexistence of simple and mixed bile salt micelles: an NMR self‐diffusion study. Biochemistry 24: 7161–7165, 1985.
 122. Shurtenberger, P., N. Mazer, and W. Kanzig. Static and dynamic light scattering studies of miceller growth and interactions in bile salt solutions. J. Phys. Chem. 87: 308–315, 1983.
 123. Schurtenberger, P., N. Mazer, and W. Kanzig. Micelle to vesicle transition in aqueous solutions of bile salt and lecithin. J. Phys. Chem. 89: 1042–1049, 1985.
 124. Small, D. M. Phase equilibria and structure of dry and hydrated egg lecithin. J. Lipid Res. 8: 551–557, 1967.
 125. Small, D. M. Physico‐chemical studies of cholesterol gallstone formation. Gastroenterology 52: 607–610, 1967.
 126. Small, D. M. A classification of biological lipids based upon their interaction in aqueous systems. J. Am. Oil Chem. Soc. 45: 108–119, 1968.
 127. Small, D. M. Studies on the size and structure of bile salt micelles, influences of structure, concentration, counterion concentration, pH and temperature. Adv. Chem. Ser. 84: 31–52, 1968.
 128. Small, D. M. The formation of gallstones. Adv. Intern. Med. 16: 243–264, 1970.
 129. Small, D. M. The physical chemistry of cholanic acids. In: The Bile Acids: Chemistry, Physiology and Metabolism. Chemistry, edited by P. P. Nair and D. Kritchevsky. New York: Plenum, 1971, vol. 1, p. 249–356.
 130. Small, D. M. Cholesterol nucleation and growth in gallstone formation. N. Engl. J. Med. 302: 1305–1307, 1980.
 131. Small, D. M. Nucleation and growth of cholesterol gallstones. Med. Chir. Dig. 9: 619–635, 1980.
 132. Small, D. M. The staging of cholesterol gallstones with respect to nucleation and growth. (Workshop on dissolution of gallstones. Proc. 6th Bile Acid Meet., Freiburg, W. Germany, October 9–11, 1980.) In: Bile Acids and Lipids, edited by G. Paumgartner, A. Stiehl, and W. Gerok. Lancaster, UK: MTP, 1981, pt. 1, p. 291–300.
 133. Small, D. M. The physical chemistry of lipids from alkanes to phospholipids. In: Handbook of Lipid Research, edited by D. Hanahan. New York: Plenum, 1986, vol. 4, p. 1–672.
 134. Small, D. M., and M. Bourges. Lyotropic paracrystalline phases obtained with aqueous ternary systems of amphiphilic substances in water. Mol. Cryst. Liq. Cryst. 1: 541–561, 1966.
 135. Small, D. M., M. Bourges, and D. G. Dervichian. Biophysics of lipid associations. I. The ternary systems. Lecithinbile salt‐water. Biochim. Biophys. Acta 125: 563–580, 1966.
 136. Small, D. M., M. Bourges, and D. G. Dervichian. Ternary and quaternary aqueous systems containing bile salts, lecithin and cholesterol. Nature Lond. 211: 816–818, 1966.
 137. Small, D. M., D. J. Cabral, D. P. Cistola, J. S. Parks, and J. A. Hamilton. The ionization behavior of fatty acids and bile acids in micelles and membranes. Hepatology Baltimore 4, Suppl.: 77S–79S, 1984.
 138. Small, D. M., S. A. Penkett, and D. Chapman. Studies on simple and mixed bile salt micelles by nuclear resonance spectroscopy. Biochim. Biophys. Acta 176: 178–189, 1969.
 139. Somjen, G. J., and T. Gilat. A non‐micellar mode of cholesterol transport in human bile. FEBS Lett. 156: 265–268, 1983.
 140. Somjen, G. J., Y. Marikovsky, P. Lelkes, and T. Gilat. Cholesterol‐phospholipid vesicles in human bile: an ultrastructural study. Biochim. Biophys. Acta 879: 14–21, 1986.
 141. Spink, C. H., K. Muller, and J. M. Sturtevant. Precision scanning calorimetry of bile salt‐phosphatidylcholine micelles. Biochemistry 26: 6598–6605, 1982.
 142. Stark, R. E., G. J. Gosselin, J. M. Donovan, M. C. Carey, and M. F. Roberts. Influence of dilution on the physical state of model bile systems: NMR and quasi‐elastic lightscattering investigations. Biochemistry 24: 5599–5605, 1985.
 143. Stark, R. E., J. L. Manstein, W. Curatolo, and B. Sears. Deuterium nuclear magnetic resonance studies of bile salt/phosphatidylcholine mixed micelles. Biochemistry 22: 2486–2490, 1983.
 144. Stark, R. E., and M. R. Roberts. 500 MHz 1H‐NMR studies of bile salt‐phosphatidylcholine vesicles. Evidence for differential motional restraint on bile salt and phosphatidylcholine resonances. Biochim. Biophys. Acta 770: 115–121, 1984.
 145. Strasberg, S. M., R. N. Redinger, D. M. Small, and R. H. Egdahl. The effect of elevated biliary tract pressure on biliary lipid metabolism and bile flow in nonhuman primates. J. Lab. Clin. Med. 99: 343–353, 1982.
 146. Tang, C. P., R. Popovitz‐Biro, M. Lahav, and L. Leiserowitz. The tetragonal crystal structure of 2:3 deoxycholate acid‐water complex. Isr. J. Chem. 18: 385–389, 1979.
 147. Thomas, D. C., and S. D. Christian. Micellar and surface behavior of sodium deoxycholate characterized by surface tension and ellipsometric methods. J. Colloid Interface Sci. 78: 466–478, 1980.
 148. Ulloa, N., J. Garrido, and F. Nervi. Ultracentrifugal isolation of vesicular carriers of biliary cholesterol in native human and rat bile. Hepatology Baltimore 7: 235–244, 1987.
 149. Ulmius, J., G. Lindblom, H. Wennerstrom, L. B. A. Johansson, K. Fontell, O. Soderman, and G. Arvidson. Molecular organization in liquid‐crystal phases of lecithinsodium cholate‐water systems studied by nuclear magnetic resonance. Biochemistry 21: 1553–1560, 1982.
 150. Vadnere, M., and S. Lindenbaum. Association of deoxycholic acid in organic solvents. J. Pharm. Sci. 71: 881–883, 1982.
 151. Vadnere, M., R. Natarajan, and S. Lindenbaum. Apparent molal volumes of bile salts in water and water‐d2 solution. J. Phys. Chem. 84: 1900–1903, 1980.
 152. Waterhous, D. V., S. Barnes, and D. D. Muccio. Nuclear magnetic resonance spectroscopy of bile acids. Development of two‐dimensional NMR methods for the elucidation of proton reasonance assignments for five common hydroxylated bile acids and their parent bile acid, 5β‐cholanoic acid. J. Lipid Res. 26: 1068–1078, 1985.
 153. Whiting, M. J., and J. M. Watts. Cholesterol crystal formation and growth in model bile solutions. J. Lipid Res. 24: 861–868, 1983.
 154. Williamson, B. W. A., and I. W. Percy‐Robb. The interaction of calcium ions with glycocholate micelles in aqueous solution. Biochem. J. 181: 61–66, 1979.
 155. Williamson, B. W. A., and I. W. Percy‐Robb. Contribution of biliary lipids to calcium binding in bile. Gastroenterology 78: 696–702, 1980.
 156. Woodford, F. P. Enlargement of taurocholate micelles by added cholesterol and monoolein: self‐diffusion measurements. J. Lipid Res. 10: 539–545, 1969.

Contact Editor

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

* Required Field

How to Cite

Donna J. Cabral, Donald M. Small. Physical Chemistry of Bile. Compr Physiol 2011, Supplement 18: Handbook of Physiology, The Gastrointestinal System, Salivary, Gastric, Pancreatic, and Hepatobiliary Secretion: 621-662. First published in print 1989. doi: 10.1002/cphy.cp060331