You are on page 1of 14

Review

Received: 14 July 2008, Revised: 25 July 2008, Accepted: 28 July 2008 Published online 17 November 2008 in Wiley Interscience

(www.interscience.wiley.com) DOI 10.1002/bmc.1134

Microseparation techniques for the study of the enantioselectivity of drugplasma protein binding
John Wiley & Sons, Ltd.

Laura Escuder-Gilabert, Mara Amparo Martnez-Gmez, Rosa Mara Villanueva-Camaas, Salvador Sagrado and Mara Jos MedinaHernndez*
Determination of urinary androgen glucuronides

ABSTRACT: Stereoselectivity in protein binding can have a significant effect on the pharmacokinetic and pharmacodynamic properties of chiral drugs. The investigation of enantioselectivity of drugs in their binding with human plasma proteins and the identification of the molecular mechanisms involved in the stereodiscrimination by the proteins represent a great challenge for clinical pharmacology. In this review, the separation techniques used for enantioselective protein binding experiments are described and compared. An overview of studies on enantiomerprotein interactions, enantiomerenantiomer interactions as well as chiral drugdrug interactions, including allosteric effects, is presented. The contribution of individual plasma proteins to the overall enantioselective binding and the animal species variability in drugplasma protein binding stereoselectivity are reviewed. Copyright 2008 John Wiley & Sons, Ltd. Keywords: stereoselective drugplasma protein binding; separation techniques; enantiomerenantiomer interactions; allosteric effects; interspecies effect

Introduction
Drug action in living organisms is the result of a large number of pharmacological processes. In this sense, the interactions between drugs and biomembranes, plasma proteins, enzymes or receptors are decisive features of the final biological activity of drugs. Most of the pharmacological processes responsible for drug action, present a high degree of stereoselectivity resulting in a difference between the activities of drugs enantiomers. In fact, very often one of them is the most active while the other may produce side-effects and even toxicity in some cases. Pharmacological differentiation of enantiomers may occur at pharmacodynamic and pharmacokinetic levels. At pharmacokinetic level the differences between the enantiomers may be due mainly to differences in drug binding to different plasma proteins. Stereoselectivity in binding can have an important effect on the chiral drug disposition such as first-pass metabolism, clearance and protein and tissue binding (Chuang and Otagiri, 2006). Therefore, the investigation of enantioselectivity of drugs in their binding with human plasma proteins and the identification of the molecular mechanisms involved in the stereodiscrimination by the proteins represent a great challenge for clinical pharmacology. Methodologies able to perform such studies play an important role in the development of new drugs and new formulations as well as in investigations on bioavailability and metabolism. Furthermore, it has often become the basis of therapeutic drug monitoring and drug management in patients (Ding et al., 1999). In this review we focused our attention in the enantioselective binding of chiral drugs with plasma proteins.

* Correspondence to: M. J. Medina-Hernndez, Departamento de Qumica Analtica, Universitat de Valncia, C/Vicente Andrs Estells s/n, 46100Burjassot, Valencia, Spain. E-mail: maria.j.medina@uv.es Departamento de Qumica Analtica, Universitat de Valncia, C/ Vicente Andrs Estells s/n, 46100-Burjassot, Valencia, Spain Abbreviations used: ACE, affinity capillary electrophoresis; ACEC, affinity capillary electrochromatography; AGP, 1-acid glycoprotein; BOF-4272, sodium ( )-8-(3-methoxy-4-phenylsulfinylphenyl)-pyrazolo[1,5-a]-1,3,5triazine-4-olate monohydrate; BSA, bovine serum albumin; CE, capillary electrophoresis; CM, chylomicron; ED, equilibrium dialysis; EKC, electrokinetic chromatography; FA, frontal analysis; HDL, high-density lipoprotein; HPCE/FA, high-performance capillary electrophoresis/frontal analysis; HPFA, high-performance frontal analysis; HPLC, high-performance liquid chromatography; HSA, human serum albumin; K, affinity constant; K1, affinity constant for the primary (high affinity) binding site; K2, affinity constant for the secondary (low affinity) binding site; LDL, low-density lipoprotein; n, number of binding sites; n1, number of primary (high affinity) binding sites; n2, number of secondary (low-affinity) binding sites; nK, total bindig affinity; ORM, orosomucoid; ORM 1, F1/S genetic variant of 1-acid glycoprotein; ORM 2, A genetic variant of 1-acid glycoprotein; R, 1-acid glycoprotein without biantennary glycans; TM--CD, heptakis-(2,3,6-tri-o-methyl)-cyclodextrin; UF, ultrafiltration; UR, 1-acid glycoprotein containing biantennary glycans; VLDL, very-low-density lipoprotein. Contract/grant sponsor: Spanish Ministry of Science and Technology. Contract/grant sponsor: European Regional Development Fund; Contract/ grant number: SAF2005-01435. Contract/grant sponsor: Generalitat Valenciana; Contract/grant number: POSTD/2007/062.

225

Biomed. Chromatogr. 2009; 23: 225238

Copyright 2008 John Wiley & Sons, Ltd.

L. Escuder-Gilabert et al.

Separation Techniques for Enantioselective Protein Binding Experiments


Generally drugplasma protein binding is a reversible and kinetically rapid interaction; therefore it should be analyzed without disturbing binding equilibrium. In the literature, most of the studies devoted to the evaluation the stereoselective binding of chiral xenobiotics to plasma proteins are conducted in three steps: (i) equilibration of xenobiotic and protein (or plasma) mixtures; (ii) separation of the free xenobiotic (unbound fraction) and xenobioticprotein complex (bound fraction); and (iii) determination of the enantiomer concentration in one of the fractions (unbound or bound) by means of a achiral or chiral separation technique depending on the availability of individual enantiomers or racemic compounds, respectively. The first and second steps have been mainly performed by means of classical membrane-based separation methodologies, such as equilibrium dialysis and ultrafiltration, but also by means of chromatographic and electrophoretic techniques. The third step consists of the analysis of one of the fractions, usually the unbound fraction. If pure enantiomers have been used for the study, traditional achiral optical, chromatographic or electrophoretic techniques can be used. For racemic mixtures, the determination of enantiomers is required, generally in the unbound fraction, using different techniques such as liquid chromatography and capillary electrophoresis and different chiral selectors. Findings and applications of such methodologies reported in the literature are described in the following sections and are summarized in Table 1.

Membrane-based Separation Techniques


Among the different methodologies used for drugprotein binding studies, equilibrium dialysis (ED) and ultrafiltration (UF) are undoubtedly the most widely used because of their simplicity and general applicability to many different systems in vitro and ex vivo. Moreover, commercial high-throughput devices in the standard 96-well plate format have been developed, such as the Equilibrium Dialyzer-96TM from Harvard Bioscience (Holliston, MA, USA; Kariv et al., 2001) and The MultiScreenTM filter assembly with UltracelTM-PPB membrane from the Millipore Corporation (Bedford, MA, USA), specifically optimized for in vitro plasma protein binding assays. Equilibrium dialysis is based on the establishment of an equilibrium state between a protein compartment and a buffer compartment (usually containing the xenobiotic or ligand), which are separated by a membrane only permeable for a lowmolecular-weight ligand. Also the use of immobilized protein onto membranes instead of protein solutions has been proposed (Randon et al., 2000). Although there is no standard method for binding measurements, equilibrium dialysis has been often regarded as the reference method for the determination of the drugprotein binding profile. However, no available experimental data support this supposed superiority. Moreover, this method has many problems, including the time needed to reach equilibrium (324 h), volume shifts, Donnan effects, hindering of the passage of free ligand, non-specific adsorption to dialysis device and dialysis membrane, particularly for highly lipophilic drugs, among others (Oravcov et al., 1996). Using equilibrium dialysis and HPLC chiral stationary phases, the enantioselective binding of ibuprofen to rat plasma (Itoh et al., 1997), zopiclone (Fernandez et al., 1999) and chlorpheniramine (Hiep et al., 1999)

to human serum albumin (HSA), 1-acid glycoprotein (AGP) and human plasma, and cetrizine to human and guinea pig plasma (Gupta et al., 2006) have been evaluated (see Table 1). Many researchers have used ultrafiltration centrifugal devices for protein binding measurements. UF is a simple and rapid method in which centrifugation forces the buffer containing free drugs through the size-exclusion membrane and achieves a rapid separation of free from protein-bound drug. Ultrafiltration has been introduced widely for routine free drug monitoring in clinical laboratories, since it offers significant advantages represented by short analysis time (usually up to 30 min), simplicity, lack of dilution effects and volume shifts. The major controversy involves the stability of the binding equilibrium during the separation process, especially in the case of low-affinity interactions (Oravcov et al., 1996) and non-specific binding of drugs on filter membranes and plastic devices which can be avoided by means of a pre-treatment of the filter membranes (Lee et al., 2003). Ultrafiltration combined with chiral separation techniques (Table 1) has been used in the literature for evaluating the stereoselectivity in binding of bimoclomol to HSA, AGP and plasma (Visy et al., 2002); alprenolol to AGP (Imamura et al., 2002) or the tryptophan to bovine serum albumin (BSA) (Randon et al., 2000); warfarin, phenprocoumon and acenocoumarol have been tested for their stereoselective binding to AGP (Hazai et al., 2006); mefloquine to HSA and AGP (Zsila et al., 2008); individual sulbenicillin enantiomers to HSA (Tsuda et al., 2001); ketoprofen to HSA (Lagrange et al., 2000); aminohydantoins in rat, dog and human plasma (Peng et al., 1999); dihydrodiazepam and lorazepam acetate to AGP (Fitos et al., 1995); carbenicillin to HSA (Itoh et al., 1996); antihistamines to HSA; and basic drugs to plasma (Martnez-Gmez et al., 2007a; 2007b).

Chromatographic and Electrophoretic Techniques


Chromatographic methods have been used for a long time for the determination of drugprotein binding parameters. Two different strategies have been proposed for evaluating the stereoselective interaction between drugs and proteins. The first one consists of the use of protein stationary phases (affinity chromatography or biointeraction chromatography). In this strategy one particular protein is immobilized onto the stationary phase and the chiral drug is injected into the chromatographic system. Assuming that the immobilization procedure does not influence the binding properties of protein, the studies may provide information on relative affinity of ligand binding and on the binding site(s) where the interaction takes place. Moreover, this methodology allows the study of the enantioselective protein-binding phenomena as well as enantiomerenantiomer or chiral drug drug interactions which are difficult to evaluate by conventional methods since they require the determination of the isomeric composition of the equilibrium mixture, i.e. by the use of pseudoracemates or enantioselective chromatographic techniques (Oravcov et al., 1996; Shibukawa et al., 1995; Fitos et al., 1999; Chen and Hage, 2004; Chen et al., 2006). HSA-stationary phases have been shown to accurately reflect the binding behavior of non-immobilized (free) HSA including its native enantioselectivity. As a consequence, HSA-stationary phases have been successfully used as a quantitative probe of drug binding to albumin (when the binding is >60%) as well as a qualitative probe for drugdrug interactions (non-cooperative, cooperative interactions and independent binding). Contrary to

226

www.interscience.wiley.com/journal/bmc

Copyright 2008 John Wiley & Sons, Ltd.

Biomed. Chromatogr. 2009; 23: 225238

Table 1. Binding parameters (R)-enantiomer n=3 K = 597 M1 1.5 (S/R) UF/HPLC (Cosmosil, 5C18-AR) (S)-enantiomer nK (or K) ratio Itoh et al. (1996) Techniques (chiral selector or HPLC column) Reference

Binding parameters literature data for stereoselective protein binding studies

Chiral drug

Protein/s

(R)- or (S)-carbenicillin

HSA

227

Biomed. Chromatogr. 2009; 23: 225238

Determination of urinary androgen glucuronides

(RS)-carbenicillin

n=3 K = 627 M1

1.3 (S/R)

UF/HPLC (Cosmosil, 5C18-AR)

(R)- or (S)-sulbenicillin 1.16 (S/R) 2.1 (R/S) 1 (S/R) UF/ACE (HSA) 2.8 (S/R) 1.8 (S/R) 1.2

HSA

2.3 (R/S)

UF/HPLC (Mightysil, RP-18)

Tsuda et al. (2001)

(RS)-sulbenicillin

UF/HPLC (Mightysil, RP-18)

Copyright 2008 John Wiley & Sons, Ltd.

(RS)-brompheniramine (RS)-chlorpheniramine (RS)-hydroxyzine

HSA

Martnez-Gmez et al. (2007a)

(RS)-orphenadrine

(RS)-verapamil (RS)-propranolol

HSA

n1 = 0.91 K1 = 5.20 103 M1 n2 = 4.7 K2 = 1.6 102 M1 n1 = 1 K1 = 4.71 103 M1 n2 = 5 K2 = 1.4 102 M1 K = 9.39 102 M1 K = 9.20 102 M1 K = 5.3 103 M1 (R- or S-enantiomer) K = 1.26 103 M1 (R- or S-enantiomer) K = 2670 M1 K = not given 13.3 3.1 (R/S) 1 1.2 (R/S) 1.1 (S/R) 1 1.4 (R/S)

n1 = 0.70 K1 = 3.95 103 M1 n2 = 4.64 K2 = 248 M1 n1 = 1 K1 = 2.40 103 M1 n2 = 2 K2 = 423 M1 n3 = not given K3 = 244 M1 n1 = 0.91 K1 = 2.27 103 M1 n2 = 2.5 K2 = 3.5 102 M1 n1 = 1 K1 = 2.27 103 M1 n2 = 3 K2 = 2.3 102 M1 K = 2.60 103 M1 K = 1.69 103 M1 K = 6.3 103 M1 (R- or S-enantiomer) K = 1.67 104 M1 (R- or S-enantiomer) K = 850 M1 K = not given

Ding et al. (1999)

(RS)-ketoprofen

HSA

Lagrange et al. (2000)

(R)- or (S)-disopyramide

CE-FA/EKC (trimethyl-cyclodextrin) ED/HPLC (L-leucinamide pre-column chiral derivatization) UF/HPLC (L-leucinamide pre-column chiral derivatization) Hummel-Dreyer HPLC n1 = 1.06 K1 = 5.21 105 M1 n2 = 3.3 K2 = 5.0 103 M1 n1 = 1.00 K1 = 5.72 105 M1 n2 = 2.7 K2 = 9.0 103 M1 K = not given K = 2.56 106 M1 K = 3.57 106 M1 K = not given 1 1.36 (S/R) 1.71 (S/R) 1

Nakagawa et al. (2003)

www.interscience.wiley.com/journal/bmc

(R)- or (S)-warfarin

ORM1 ORM2 ORM1 ORM2

n1 = 1.10 K1 = 5.93 105 M1 n 2 = 2.4 K2 = 6.0 103 M1 n1 = 0.95 K1 = 5.99 105 M1 n2 = 2.4 K2 = 1.4 104 M1 K = not given K = 1.89 106 M1 K = 2.08 106 M1 K = not given

228
Binding parameters (R)-enantiomer Affinity constants not given UF/EKC (-ciclodextrin) UF/EKC (-ciclodextrin) (S)-enantiomer nK (or K) ratio Techniques (chiral selector or HPLC column) Reference Hazai et al. (2006) 2.61(S/R) 1.1 (R/S) 2.26(S/R) 1.12(R/S) 1.76 (R/S) 1.21(S/R) 1.85(S/R) 1.27 (S/R) 1.77(S/R) 1.5 (S/R) UF/EKC (-ciclodextrin) UF/HPLC (chiral OD column) UF/HPLC (AGP column) HPFA/HPLC (achiral) Imamura et al. (2002) Fitos et al. (1995) Kimura et al. (2006) CE/FA Mohamed et al. (1999) CE/FA Mohamed et al. (2000) HPFA/CE Ohnishi et al. 2002) Rodriguez Rosas et al. (1999a) n = 0.97 K = 6.59 105 M1 3.00 (S/R) HPFA/HPLC (ovomucoid immovilized column) HPFA/HPLC (Develosil 300 ODS UG5) HPFA/HPLC (Chiralcel OJ-R) nK = 2.19 104 M1 nK = 1.53 107 M1 nK = 8.52 106 M1 nK = 8.44 103 M1 nK = 1.31 107 M1 1.20 (R/S) 2.23 (S/R) 1.05 (R/S) 1.09 (S/R) 2.27 (R/S) HPFA/HPLC (YMC-Pack ODS-AM) Rodriguez Rosas et al. (1999b) Shibukawa et al. (2002a) Shibukawa et al. (2002b) (+)-(R)-enantiomer n = 1.0 K = 6.3 105 M1 K = 1.3 104 M1 K = 2.6 104 M1 nK = 9.49 106 M1 nK = 21.4 106 M1 nK = 1.06 106 M1 nK = 1.07 106 M1 nK = 1.02 106 M1 nK = 7.66 106 M1 nK = 3.66 107 M1 nK = 2.75 104 M1 nK = 1.99 105 M1 nK = 2.03 106 M1 nK = 2.12 104 M1 nK = 4.01 105 M1 n = 0.74 K = 3.17 107 M1 7.7 (S/R) 1.4 (S/R) 1.51(S/R) 2.58 (R/S) 1.17 (S/R) 1.03(S/R) 1.00 (R/S) 1.03 (R/S) 1.03(R/S) 1.02(S/R) 1.02(R/S) 1.01(S/R) 1.02(S/R) 1.00(S/R) 1.2 (R/S) n = 0.99 K = 2.15 105 M1 ()-(S)-enantiomer n = 1.0 K = 9.7 105 M1 K = 1.0 105 M1 K = 3.6 104 M1 nK = 14.4 106 M1 nK = 8.30 106 M1 nK = 1.25 106 M1 nK = 1.10 106 M1 nK = 1.02 106 M1 nK = 7.47 106 M1 nK = 3.57 107 M1 nK = 2.81 104 M1 nK = 1.96 105 M1 nK = 2.06 106 M1 nK = 2.43 104 M1 nK = 4.02 105 M1 n = 0.74 K = 2.59 107 M1 L. Escuder-Gilabert et al. nK = 2.64 104 M1 nK = 6.86 106 M1 nK = 8.09 106 M1 nK = 7.77 103 M1 nK = 2.97 107 M1

Table 1.

(Continued)

Chiral drug

Protein/s

(RS)-acenocoumarol

www.interscience.wiley.com/journal/bmc

(RS)-phenprocoumon

(RS)-warfarin

(RS)-alprenolol

ORM1 ORM2 AGP ORM1 ORM2 AGP ORM1 ORM2 AGP AGP

(RS)-dihydrodiazepam (RS)-lorazepam acetate (R)- or (S)-oxybutynin

AGP

(R)- or (S)-nilvadipine

Copyright 2008 John Wiley & Sons, Ltd.

(R)- or (S)-verapamil

(R)- or (S)-propranolol

ORM1 (UR) ORM1 (R) ORM2 (UR) ORM2 (R) HDL Normal LDL Oxidized LDL HDL Normal LDL Oxidized LDL HDL Normal LDL AGP

Semotiadil [(R)-isomer] Levosemotiadil [(S)-isomer] Semotiadil [(R)-isomer] Levosemotiadil [(S)-isomer] (R)- or (S)-oxybutynin

HSA

Biomed. Chromatogr. 2009; 23: 225238

(R)- or (S)-desethyloxybutynin

HSA AGP LDL HSA AGP

Table 1. Binding parameters (R)-enantiomer ()-(S)-enantiomer nK = 1.0 105 M1 nK = 5 105 M1 7.7(S/R) 1 1.0 UF/HPLC (AGP column) UF/HPLC (AGP column) (S)-enantiomer nK (or K) ratio Visy et al. (2002) Techniques (chiral selector or HPLC column) Reference

(Continued)

Chiral drug

Protein/s

()-bimoclomol

229

Biomed. Chromatogr. 2009; 23: 225238

(RS)-mefloquine

(+)-(R)-enantiomer nK = 1.3 104 M1 nK = 5 105 M1 Affinity constants not given

Zsila et al. (2008)

Determination of urinary androgen glucuronides

(+)- or ()-zopiclone

AGP HSA HSA AGP ORM 1 ORM 2 AGP HSA K = 2.23 104 M1 K = 524 M1 K = 8.5 103 M1 K = 293 105 M1 K = 3.04 103 M1 K = 297 M1 5.0 (+/) 1.5 (+/) 2.0 (R/S) K = 1.53 104 M1 K = 431 M1 ()-(R)-enantiomer n = 0.72 K = 11.95 105 M1 2.4 (/+) 1.7(+/) 2.6 (+/) 1.8 (+/) ED/HPLC (chiral carbamate cellulose column) ED/HPLC (amylose Chiral-pak AD column) UF/HPLC (L-leucinamide pre-column chiral derivatization)

Fernandez et al. (1999) Hiep et al. (1999) Glwka and Caldwell (2002)

()-chlorpheniramine

AGP HSA

(RS)-indobufen

HSA plasma

(S)-indobufen n = 1.0 K = 1.21 105 M1 2.1 (R/S)

(RS)-ibuprofen

Rat plasma

(+)-(S)-enantiomer n = 0.92 K = 4.65 105 M1 n = 1.87 K = 2.10 105 M1 n = 0.91 K = 5.76 104 M1

Itoh et al. (1997)

Copyright 2008 John Wiley & Sons, Ltd.

Aminohydantoins

ED/HPLC (SUMICHIRAL OA2500, 2500S) UF/HPLC (chiral polysacharide column)

Peng et al. (1999)

(RS)-compound I

(RS)-compound II

1.2(R/S) 1.3(R/S) 1.7(R/S) 1.4(R/S) 1.8(R/S) 3.0(R/S) 1.1 (/+)

()-BOF-4272

Rat plasma Dog plasma Human plasma Rat plasma Dog plasma Human plasma HSA

K = 1.21 104 M1 K = 1.68 104 M1 K = 1.08 106 M1 K = 3.89 104 M1 K = 5.07 104 M1 K = 1.40 106 M1 (+)-enantiomer K = 1.22 105 M1 n = 2.30 K = 1.01 104 M1 K = 1.34 104 M1 K = 6.42 107 M1 K = 2.74 104 M1 K = 2.83 105 M1 K = 4.61 105 M1 ()-enantiomer K = 2.32 105 M1 n = 1.30

HPFA/HPLC (-cyclodextrinimmobilized silica column)

Shibukawa et al. (1995)

Bovine serum albumin Rat serum albumin

www.interscience.wiley.com/journal/bmc

BOF-4272. sodium ()-8-(3-methoxy-4-phenylsulfinylphenyl)-pyrazolo[1,5-a]-1,3,5-triazine-4-olate monohydrate. Compound I, 1-[(phenylmethylene)-amino]-3-[(6)-1-carboxy-2-(4-hydroxyphenyl)-ethyl]-imidazoline-2,4-dione. Compound II, 1-[(phenylmethylene)-amino]-3-n-hexyl-(6)-5-isopropyl-imidazolidine-2,4-dione.

L. Escuder-Gilabert et al. the behavior of albumin, the immobilized 1-acid glycoprotein is not suitable as a screening tool for AGP binding affinity (Oravcov et al., 1996). The second chromatographic strategy for evaluating the enantioselective drugprotein binding consists of the use of high-performance size-exclusion chromatographic techniques in the modality of frontal analysis (HPFA). HPFA was first developed by using a restricted access type HPLC column which excludes large molecules of plasma protein but retains a drug of small molecular size (Shibukawa et al., 1999a, b). In HPFA, an excess volume of sample is injected directly to a restrictedaccess type HPLC column and mobile phase compositions that guarantee not to modify the binding equilibrium (usually physiological pH 7.4 of phosphate buffer without adding any organic modifier) are used. After sample injection, drug and protein are separated from each other, keeping the binding equilibrium. Finally, the unbound drug is eluted as a trapezoidal peak with a plateau region separated from protein. The plateau height and the peak area correspond to the unbound drug and total drug concentrations, respectively. Comparative studies against conventional ultrafiltration/HPLC method indicate the reliability of this method (Shibukawa et al., 1999b). One of the drawbacks of membrane-based methodologies is related to drugs with high affinity towards the protein since the free drug concentration is low and then difficult to detect. To overcome this problem in HPFA methodology the on-line preconcentration of unbound fraction has been proposed. By coupling HPFA with a chiral separation system such as a chiral HPLC column or an electrokinetic chromatography method, the preconcentrated unbound concentration of a racemic drug can be determined enantioselectively. These methodologies have been applied to determine the enantioselective protein binding of several racemic drugs, such as warfarin (He et al., 1997), fenoprofen (Shibukawa et al., 1993), ketoprofen (Shibukawa et al., 1992), nilvadipine (Shibukawa et al., 1994), BOF-4272 (Shibukawa et al., 1995), semotiadil and levosemotiadil (Rodriguez-Rosas et al., 1999a), oxybutynin (Shibukawa et al., 2002a) and N-desethyloxybytynin (Shibukawa et al., 2002b) Capillary electrophoresis (CE) offers some advantages in drugprotein binding studies such as low sample requirements, short analysis times and high separation efficiencies and sample throughput. Several CE approaches have been developed for the quantitative assessment of drugprotein interactions that include methods that use proteins as additive buffers in the modality called affinity capillary electrophoresis (ACE), or immobilized proteins into the capillary in the modality of affinity capillary electrochromatography (ACEC). ACE offers several possibilities for studying ligandprotein binding interaction depending on the disposition of protein and drug in the electrophoretic system such as: frontal analysis, HummelDreyer method, vacancy peak and vacancy affinity capillary electrophoresis (Busch et al., 1997). All these approaches are complementary rather than competitive since the information that can be obtained from each kind of experiment is different (Busch et al., 1997). Among these methodologies, for enantioselective drugprotein binding studies, frontal analysis is concluded to be superior to the others because of its greater simplicity, speed and versatility to study multiple equilibria (Busch et al., 1997). Frontal analysis (FA) is based on the injection of relatively large sample plugs (100200 nL), which consist of a pre-equilibrated mixture of drug and protein and therefore some complex may also be present. In ACE/FA, both hydrodynamic injection and electrokinetic injection are possible. In ACE/FA following hydrodynamic injection, a plug of drug protein mixed solution is introduced onto the capillary containing the running buffer (usually pH 7.4), and positive voltage is applied. In these conditions, basic drugs are positively charged, while plasma proteins such as HSA and AGP have a net negative charge. Therefore the unbound drug can be separated from the protein and the drugprotein complex. By the addition of a chiral selector in the running buffer, the enantiomers present in the unbound fraction are separated and their concentrations are calculated from their respective plateau heights. In ACE/FA with electrokinetic injection, only the unbound drug present in the pre-equilibrated mixture is introduced selectively onto the capillary by applying positive voltage (for positively charged drugs) at the sample injection side. Using this strategy the negatively charged protein and bound drug are not introduced. The unbound drug zone migrates through the capillary towards the cathodic end and can be separated into two zones of the enantiomers by the addition into the running buffer of an appropriate chiral selector in optimized experimental conditions. The unbound concentration of each enantiomer can be determined from their respective plateau heights (Ohara et al., 1995). Using these methodologies, Ohara et al. (1995) evaluated the enantioselective binding of verapamil and propranolol to HSA using trimethyl--cyclodextrin as chiral selector. The authors concluded that both injection methods gave the same results with good repeatability, and the results agreed well with those obtained by the conventional ultrafiltration method followed by chiral HPLC analysis. The sample injection volume in ACE methodologies (100200 nL) is smaller by more than two orders of magnitude than the sample volume required by the ultrafiltration method (Ding et al., 1999). The same methodologies can be used for the study of the individual enantiomer binding to proteins, avoiding in this case the addition of the chiral selector to the running buffer. The research group of A. Shibukawa and coworkers using highperformance capillary electrophoresis/frontal analysis (HPCE/FA) studied the interactions of individual enantiomers of nilvadipine (Mohamed et al., 1999), verapamil (Mohamed et al., 2000) and propranolol (Ohnishi et al., 2002) to plasma lipoproteins. The current major disadvantage of ACE/FA is the relatively high detection limit (micromolar range) and, thus, the sensitivity is insufficient for the analysis of clinical samples (nanomolar range). Limitations of this method also include the need to minimize protein-wall adsorption. As it occurs in affinity chromatography, the most important applications in the context of enantioselective drugprotein binding of ACEC are related to the study of drugdrug interaction and enantiomerenantiomer interactions (Ye et al., 2002).

Enantioselective DrugPlasma Protein Binding Studies


Human plasma contains over 60 proteins, HSA, AGP, lipoproteins and globulins being the most important drug binding proteins (Gonzlez-Alonso and Snchez-Navarro, 1998; Evans et al., 1992). All proteins can contribute to the plasma protein binding of the drug simultaneously, and the overall plasma protein binding is the sum of each binding. In fact, HSA shows high affinity towards neutral and acidic compounds; AGP is a major binding protein for many cationic drugs and to a lesser extent it binds

230

www.interscience.wiley.com/journal/bmc

Copyright 2008 John Wiley & Sons, Ltd.

Biomed. Chromatogr. 2009; 23: 225238

Determination of urinary androgen glucuronides some anionic and neutral drugs. Lipoproteins bind to non-ionic and lipophilic drugs and some anionic drugs while globulins interact insignificantly with the majority of drugs. In case of a hydrophobic basic drug, HSA, AGP and lipoproteins often contribute to the plasma protein binding simultaneously. Therefore, in vivo binding studies using plasma samples and in vitro binding studies using each plasma protein are helpful for understanding the plasma protein binding properties. Interactions of drugs with plasma proteins are selective and of diverse nature. Usually reversible interactions take place, while irreversible covalent binding of drugs to plasma proteins rarely occurs. This irreversible binding is responsible for certain types of toxicity such as the carcinogenic effect of chemical substances (GonzlezAlonso and Snchez-Navarro, 1998). Both reversible and irreversible interactions can be enantioselective for some chiral compounds. In this review only reversible interactions are considered. When protein and racemic chiral drugs interact, two diastereomeric adducts are formed with potential differences in the protein binding, which may result in different pharmacokinetic profiles for the individual enantiomers (Chuang et al., 2006). If both enantiomers bind to the same binding site with different affinities, competition between enantiomers occurs and the lesser-bound enantiomer is more displaced than the other. However, when enantiomers bind to different binding sites, the binding of one enantiomer does not produce an effect on the binding property of the other isomer. Therefore, since enantiomer enantiomer interactions may occur in various pharmacokinetic processes in the body, the pharmacokinetic properties after racemate administration may differ from those following administration of the individual isomers. In addition to enantiomerenantiomer interactions, other effects such as chiral drugdrug interactions and allosteric effects are of great importance in the evaluation of stereoselective binding of chiral drugs with plasmatic proteins. In Table 1 the enantioselectivity values of drugprotein binding reported in literature since 1995 are summarized. In this table enantioselectivity in drugprotein binding (nK or K ratio) is defined as the ratio between the affinity constant of enantiomer that presents the greatest affinity towards protein and the affinity constant of the other enantiomer. In the same column in brackets the relationship between enantiomers is indicated. Moreover, the methodology used for the evaluation ofthe enantioselective binding is indicated. bic pocket with a cationic region (Kwong, 1985; Ascoli et al., 2006). Two additional specific binding sites are also proposed for bilirubin and fatty acids (Chuang and Otagiri, 2006). In addition to the evaluation of affinity constants of enantiomers towards proteins, it is of special interest the study of the specific sites in the protein molecule where the enantiomers interact. To identify the binding sites of enantiomer drugs in the HSA molecule, and elucidate if the binding of enantiomers follows an independent or competitive model, various site marker ligands are used. Warfarin and phenylbutazone, diazepam and digitoxin were usually used as marker ligands representative of site I, II and III, respectively. In these studies, the displacement of equilibrium between a racemic drug or an individual enantiomer and HSA in the presence of a marker ligand is evaluated. Different studies can be found in the literature and are commented on below. Itoh et al. (1996) studied the stereoselective binding to HSA of carbenicillin, a semi-synthetic penicillin, as well as enantiomer enantiomer interactions. Binding of carbenicillin enantiomers was measured individually using (R)-carbenicillin and (S)-carbenicillin and racemic carbenicillin. It was found that the unbound fraction of (R)-carbenicillin was greater than that of (S)-carbenicillin, similar to that observed in human plasma. Data analysis showed a relatively high affinity binding site for (S)-carbenicillin, which was not observed for (R)-carbenicillin. Binding constants for the low-affinity sites were similar for both enantiomers. When the unbound fraction of each enantiomer was measured in the presence of different amounts of the opposite enantiomer, enantiomerenantiomer interaction was observed only at relatively high concentrations, which indicates that carbenicillin enantiomers mutually displace each other at the binding sites on HSA. The possibility of competitive interactions was supported by the results obtained with site marker ligands. The results suggest that both enantiomers are bound to site I, but not to site II, and that the enantiomers probably share common binding sites. Three binding sites on HSA were identified and one of the sites was stereoselective. The present study suggests the possibility of competitive interactions both at the stereoselective and non-stereoselective sites. This behavior is similar to reported data for other -lactam antibiotics such as cephalothin and phenoxymethylpenicillin (Itoh et al., 1996). Using a similar methodology the authors studied the stereoselective binding and degradation of sulbenicillin, a semisynthetic -lactam antibiotic, in the presence of HSA (Tsuda et al., 2001). The authors identified at least two types of binding sites on HSA for sulbenicillin enantiomers. The high-affinity site was stereoselective, the binding constant of (R)-sulbenicillin being approximately 2.3-fold greater than that of (S)-sulbenicillin. The use of site marker ligands revealed that the stereoselective site was site I. (R)-sulbenicillin and (S)-sulbenicillin appeared to displace each other competitively at both binding sites. On the other hand, it was observed that degradation of (R)-sulbenicillin in the presence of HSA was faster than that of (S)-sulbenicillin. Martnez-Gmez et al. (2007a) studied the enantioselective binding of the antihistamines brompheniramine, chlorpheniramine, hydroxyzine and orphenadrine to HSA. Owing to the high affinity of these drugs towards HSA, the authors proposed a procedure that includes (i) the ultrafiltration of preequilibrated samples containing HSA and the racemic drug, (ii) the treatment of the insoluble fraction obtained, which contains the bound fraction of drug, with an organic solvent to provoke the precipitation of proteins and the liberation of the

Human Serum Albumin Studies


HSA is the most abundant plasma protein in the circulatory system, with a typical concentration of 3545 g/L, responsible for the majority of drug binding in plasma. HSA possesses the highest enantioselectivity among plasma proteins. It shows high affinity towards neutral and acidic compounds (GonzlezAlonso and Snchez-Navarro, 1998; Evans et al., 1992; Kwong, 1985), but also for hydrophobic basic drugs. Most of these interactions occur at two well-defined regions of the HSA molecule, known as sites I and II, although a third site (site III) is also described. The warfarinazapropazone site (site I), which binds warfarin, coumarins, salicylates and many other drugs, is seen as a broad binding area that bears overlapping subsites. The indolebenzodiazepine site (site II), where other drugs such as diazepam, naproxen, ibuprofen and L-tryptophan bind, and site III, where digitoxin binds, are represented as a narrow hydropho-

231

Biomed. Chromatogr. 2009; 23: 225238

Copyright 2008 John Wiley & Sons, Ltd.

www.interscience.wiley.com/journal/bmc

L. Escuder-Gilabert et al. bound drug and (iii) the chiral separation and determination of enantiomers by ACE using HSA as chiral selector. The study performed to identify the binding sites of enantiomer drugs in the HSA molecule indicated that both enantiomers of brompheniramine and chlorpheniramine bind to site II and therefore they follow a competitive model. For orphenadrine, the least bound enantiomer binds to digitoxin site (site III) while for the most bound enantiomer no specific site can be defined. For hydroxyzine, the least bound enantiomer does not bind to any of these three sites; however, the most bound enantiomer binds preferentially to site I. So, for orphenadrine and hydroxyzine, an independent binding model of enantiomers to HSA could be considered. The affinity constants of drug enantiomer HSA revealed the existence of enantioselective binding of antihistamines to HSA (Table 1). The binding of orphenadrine to HSA presents the highest enantioselectivity among the studied antihistamines. When two or more drugs that bind at the same binding site are administered together, an increase in their unbound drug concentrations can be observed. Drugdrug inhibitory interaction is of great clinical significance mainly for drugs that are highly bound to plasma proteins (>90%), since a small decrease in bound drug fraction provokes an increase in the unbound drug fraction. These interactions can also occur between enantiomers of different drugs. Shibukawa et al. (1999b) developed an on-line HPFA system coupled with an ovomucoid immobilized chiral column system for study of drugdrug interactions between warfarin and phenylbutazone in plasma samples. Both warfarin and phenylbutazone bind strongly to site I on HSA. In human plasma spiked with racemic warfarine, (S)-warfarine was bound more strongly than (R)-warfarine. The effect of phenylbutazone on the plasma protein binding of warfarin was enantioselective. The addition of phenylbutazone increased the unbound concentrations of (R)-warfarin and (S)-warfarin by 1.63 and 2.12 times, respectively, resulting in the reversal of enantioselectivity. Similar results were observed for HSA, the protein responsible for the plasma protein bindings of both drugs. This is compatible with the known serious side effects of the coadministration of warfarin and phenylbutazone (Shibukawa et al., 1999b). Ohara et al. (1995) studied the enantioselective binding of racemic verapamil to HSA. The authors indicated that the unbound concentration of (S)-verapamil was 1.7 times higher than that of (R)isomer. Similar results were obtained by Ding et al. (1999) for verapamil. Moreover, when racemic ibuprofen was added into verapamil-HSA solution, (R)-verapamil was partially displaced while (S)-verapamil was not displaced at all. The authors indicate that the (R)-ibuprofen is the enantiomer responsible of the observed displacement. A binding synergism effect between bupivacaine and verapamil was observed and the binding site study suggests that verapamil and bupivacaine occupy different binding site of HSA (sites II and III, respectively). No obvious stereoselective binding of propranolol to HSA was observed (Ding et al., 1999). Allosteric interactions can be determinant in several biological processes. They take place when the interactions of a ligand with a binding agent change the interactions of another ligand with the same agent at a separate binding site. Such interactions can occur during the binding of drugs to plasma proteins such as HSA. Although in the molecule of HSA there are two main ligand binding regions, there are several indications that simultaneous binding of two ligands induces conformational changes of the protein. The most important allosteric interactions have been found for chiral ligands (Chen and Hage, 2004). Extensive investigations have been performed into the effect of chiral arylpropionic acid derivatives, so called profens, on the stereoselective binding of chiral drugs. Profens bound almost exclusively to HSA. Their primary high-affinity binding site on HSA, site II (benzodiazepine site), manifests more significant stereoselective behavior in comparison to site I (warfarin site). Some of them are not bound stereoselectively, such as (R)- and (S)-fenoprofen (Perrin, 1973). Published data conflict with respect to the enantioselective protein binding parameters of (R)- and (S)-ketoprofen and (R)- and (S)-flurbiprofen. Jin et al. (2008) concluded that flurbiprofen and ketoprofen display concentrationdependent and extensive binding to human plasma. The binding to high-affinity site was only slightly greater in the case of the (S)-flurbiprofen compared with the (R)-flurbiprofen. The binding of ketoprofen enantiomer was not stereoselective. Lagrange et al. (2000) indicated that only for high ketoprofen concentrations was weak enantioselectivity observed. The authors concluded that differences in HSA preparations used and/or the presence of interfering compounds may explain the variability in the reported protein binding of ketoprofen enantiomers. Other profens such as ibuprofen (Evans et al., 1989) and etodolac (Jin et al., 2008) show high enantioselectivity in HSA proteinbinding. In plasma, (S)-etodolac binds more strongly than (R)etodolac to HSA [the unbound fraction ratio of (S)-/(R)-etodolac was from 0.30 to 0.98; Jin et al., 2008]. In contrast, for ibuprofen, the binding of the R-enantiomer to HSA is stronger than Sibuprofen. Moreover, there is evidence for the existence of secondary sites for the binding of ibuprofen enantiomers (Fitos et al., 1999). Certain 3-substituted 1,4-benzodiazepines exhibit significantly enhanced binding of the (S)-enantiomers in the presence of some other ligands bound to site I. Fitos et al. (1999) studied the effect of ibuprofen enantiomers on the stereoselective binding of 3-acyloxy-1,4-benzodiazepines to HSA using both ultrafiltration and affinity chromatography with Sepharose-immobilized protein. The results indicate that, while the binding of (R)-benzodiazepines was unchanged, the free fractions of the (S)-enantiomers were subjected to various effects; instead of simple competition, ibuprofen enantiomers exert a complex effect on the HSA binding of the (S)-enantiomers of benzodiazepines and the overall effect depends on the structure of benzodiazepine. Systematic cobinding studies of benzodiazepines and ibuprofen enantiomers concluded that the selective displacement of the benzodiazepine (S)-enantiomers does not correspond only to simple competition; cooperative allosteric interaction between different binding sites also occurs. In fact for (S)-ibuprofen, in addition to the common binding site, another major binding region has been observed. The effects of (R)- and (S)-ibuprofen on the binding of benzodiazepines to HSA were also examined by biointeraction chromatography (Chen et al., 2006). The authors developed a general model to describe the binding of HSA to benzodiazepines and ibuprofen. This model considers an immobilized agent L which has at least two binding sites, one for injected solute A and the other for solute I that is used as a mobile-phase additive. In this model, the binding of I to L introduces a change in the conformation of L and causes the association equilibrium constant for A with L to change. An equation based on the retention factor k at any given concentration of I which is useful

232

www.interscience.wiley.com/journal/bmc

Copyright 2008 John Wiley & Sons, Ltd.

Biomed. Chromatogr. 2009; 23: 225238

Determination of urinary androgen glucuronides in quantitatively studying allosteric effects is developed. This model gave good agreement with previous reports examining the binding of benzodiazepines to HSA. greater enantioselectivity was observed in phenprocoumon binding. Acenocoumarol possessed the highest enantioselectivity in AGP binding due to the weak binding of its (R)-enantiomer. Kimura et al. (2006) investigated both the effect of genetic variants and biantennary glycans on the stereoselective binding of oxybutynin to AGP. For this purpose, the authors first isolated ORM 1 and ORM 2 variants from native AGP and then each variant were separated in two fractions with (R) and without (UR) biantennary glycans. After that, the affinity of individual (R)and (S)-oxibutynin towards each AGP fraction was studied by means of HPFA/HPLC. The results obtained showed that total binding affinities of oxybutynin enantiomers towards ORM 1 variants were higher than those for ORM 2. Moreover, stereoselective binding was only observed for ORM 1 variant, being opposite between ORM 1 fractions with or without biantennary glycans. On the other hand, besides changes in the drugprotein binding values related to structural alterations of protein molecules, the most important changes in the unbound drug fraction are related to diseases that produce variations in plasma proteins levels. In this sense, AGP is an acute-phase reactant protein and its serum levels are elevated in cases of certain chronic diseases such as rheumatoid arthritis, renal failure and cirrhosis. In these situations, the drug binding percentages to AGP are enhanced depending on the increased concentrations of the binding molecule; therefore a decrease in the pharmacological effect would be found. Imamura et al. (2002) investigated the stereoselective protein binding of alprenolol, a -adrenergic blocking agent, in renal disease patient sera compared with that in the sera of healthy volunteers. The in vitro stereoselective protein binding of alprenolol was determined in undiluted serum and in AGP solutions by ultrafiltration and chiral HPLC analysis using a Chiral OD column. ()-(S)-alprenolol showed slightly higher affinity towards AGP than (+)-(R)-alprenolol. The stereoselective serum protein binding of alprenolol was significantly altered in renal disease patients compared with healthy volunteers. A good correlation between the unbound (R)/(S) ratio and AGP concentration in serum was found. Other studies have also shown the stereoselectivity in the binding of chiral drugs to AGP. Fitos et al. (1995) studied the stereoselectivity in binding of dihydrodiazepam and lorazepam acetate to native AGP by ultrafiltration and HPLC using a chiralAGP stationary phase. The results obtained indicated in both cases that the (S)-enantiomers showed stronger binding, though to different extents. Others studies about the enantioselective binding of chiral drugs towards AGP and HSA and their relative impact on the binding of enantiomers towards whole plasma proteins is commented on later.

a1-Acid Glycoprotein Studies


As has been commented above, AGP (also called orosomucoid, ORM) is the most important plasma protein responsible for plasma protein binding of a basic drug. The AGP molecule consists of a single polypeptide chain of 183 amino acids and five N-glycan chains which have di-, tri- and tetra-antennary structures, with sialic acids as the terminal group. The glycan structures show microheterogeneity under physiological conditions, and the partially desialylated AGP is known to exist in plasma of patients with liver disease (Shibukawa et al., 1999a). Because sialic acid has a negative charge, it may contribute to the binding of basic drugs with AGP. The function of sialic acid groups at the terminal of AGP glycan chains with respect to chiral discrimination between enantiomers of two basic drugs (propranolol and verapamil), using the HPCE/FA method using heptakis(2,3,6-tri-o-methyl)--cyclodextrin as chiral selector, has been studied (Shiono et al., 1997). The authors found that the unbound concentration of (S)-verapamil was 1.3 times higher than that of (R)-verapamil in native AGP solution, and this selectivity was not affected by desialylation of AGP. Further, enzymatic elimination of end-terminal galactose residues of the desialylated AGP did not change the binding of either isomer of verapamil. On the other hand, the unbound concentration of (R)-propranolol was 1.27 times higher than that of (S)-propranolol in native AGP solution. Desialylation did not change the unbound concentration of (R)-propranolol, but caused the unbound concentration of (S)-propranolol to rise up to the same level of (R)propranolol, resulting in the loss of enantioselectivity. This result suggests that the sialic acid residues may be regarded as one origin of enantioselectivity in AGP-propranolol binding, while they are not responsible for the enantioselective AGPverapamil binding. Besides the high heterogeneity of glycans, the protein part of AGP has also been found to show polymorphism. The variants are encoded by two different genes: the F1 and S variants are encoded by the alleles of the same gene, while the A variant is encoded by a different gene (Dente et al., 1987). There is a difference of at least 22 amino acid residues between the F1/S (ORM 1) and A (ORM 2) variants, while F1 and S forms differ only in a few residues (Yuasa et al., 1997). ORM 1 and ORM 2 variants have been shown to possess different binding properties as well as stereoselectivity. Warfarin is a selective ligand of ORM 1 variant of human AGP and the possibility of high binding stereoselectivity was raised (Herv et al., 1998). Nakagawa et al. (2003) determined preference in binding of (S)-warfarin to ORM 1 variant in contrast with the higher affinity of (S)-disopyramide towards ORM 2 variant. In a paper by Hazai et al. (2006) coumarin-type anticoagulants, warfarin, phenprocoumon and acenocoumarol, were tested for their stereoselective binding to native AGP and ORM 1 and ORM 2 genetic variants. The results showed that all investigated compounds bind more strongly to ORM 1 variant than to ORM 2 and no significant enantioselectivity was observed in binding to ORM 2. Binding to native AGP (consisting of about 70% ORM 1 and 30% ORM 2) resembled binding to ORM 1 rather than to ORM 2. ORM 1 and human native AGP bind preferentially to (S)enantiomers of warfarin and acenocoumarol, while slightly

Lipoproteins Studies
Lipoproteins consist of a lipophilic core (cholesterol ester + triglycerides) surrounded by a surface layer comprising polar lipids (phospholipids + free cholesterol) and apolipoproteins. Plasma lipoproteins bind mainly the lipophilic neutral drugs and basic drugs (Urien, 1986), and act as the transport system in plasma circulation of these drugs. The binding study of plasma lipoproteins is important because considerable inter-individual differences as well as the variation depending on disease state are found in the plasma concentrations, which possibly affects the plasma distribution of the drug.

233

Biomed. Chromatogr. 2009; 23: 225238

Copyright 2008 John Wiley & Sons, Ltd.

www.interscience.wiley.com/journal/bmc

L. Escuder-Gilabert et al. Plasma lipoproteins are classified into several subclasses according to their density, such as high-density lipoprotein (HDL), lowdensity lipoprotein (LDL), very-low-density lipoprotein (VLDL) and chylomicron (CM). Among these, HDL and LDL are the most important drug-transporting proteins because of their higher plasma concentrations than others. Oxidized LDL resulting from LDL in vivo conversion has also been reported to be a highaffinity drug binding protein (Mohamed et al., 1999). In addition, since apolipoproteins and lipid constituents such as free cholesterol, cholesterol ester and some phospholipids are chiral compounds, the binding of a racemic drug to lipoproteins may be different between the enantiomers, which can be related to enantioselective pharmacokinetic properties. The enantioselective binding of individual nilvadipine, verapamil and propranolol enantiomers to plasma lipoproteins by capillary electrophoresis coupled with frontal analysis has been studied (Mohamed et al., 1999, 2000; Ohnishi et al., 2002). It was found that the binding of nilvadipine, verapamil and propranolol to HDL, LDL and oxidized LDL was non-specific and not enantioselective. Partition-like binding to the lipid part of these lipoproteins seemed to occur dominantly. The total binding affinities of nilvadipine (Mohamed et al., 1999) and verapamil (Mohamed et al., 2000) to LDL were about seven times stronger than those to HDL, and the oxidation of LDL enhanced the binding affinity significantly. For propranolol the total binding affinity to LDL was 17 times higher than that of propranololHDL binding (Ohnishi et al., 2002). et al., 1997). This means that, because of the stronger affinity of AGP, the protein binding in human plasma reflects the enantioselectivity of AGP. A similar enantioselective plasma protein binding behavior was observed for oxybutynin (Shibukawa et al., 2002a) and Ndesethyloxybytynin (Shibukawa et al., 2002b), a major active metabolite of oxybutynin. The results obtained showed that oxybutynin and N-desethyloxybutynin are bound in human plasma strongly and enantioselectively. The unbound fraction of (R)-oxybutynin and (S)-N-desethyloxybutynin was 1.56 and 1.96 times higher than their respective isomer, respectively. Again, AGP plays the dominant role in the enantioselective plasma protein binding. AGP has also been shown to be determinant on the biclomol total plasma protein binding stereoselectivity (Visy et al., 2002) showing the stronger binding of the ()-(S)-enantiomer. The binding of biclomol to HSA was found to be weak and not stereoselective. ()-(S)-Bimoclomol was displaced in the presence of specific marker ligands for the ORM 1 genetic variant of human AGP. Similary the protein binding of mefloquine, an antimalarian drug, has shown no significant stereoselectivity in binding to HSA, while the binding to AGP was stereoselective showing greater affinity towards the ()-enantiomer. Similar experiments performed with the genetic variants of AGP revealed that the favored binding of the ()-enantiomer occurs on the ORM 1 variant. On the ORM 2 variant, which takes up about 30% of native AGP, the stereoselectivity is reversed, and the binding of the (+)-enantiomer is stronger (Zsila et al., 2008). Fernandez et al. (1999) studied the binding of racemic zopiclone, a short-acting hypnotic agent, and of its two enantiomers to plasma proteins, HSA and AGP. The total plasma protein binding percentages were 79.3 5.5, 83.8 5.2 and 75.1 2.1%, for racemic zopiclone, ()-zopiclone and (+)-zopiclone, respectively. On the contrary, the binding of (+)-zopiclone to AGP or HSA was higher than that found for ()-zopiclone. This opposite stereoselectivity observed with total plasma proteins and isolated proteins could be explained by the preferential binding of ()zopiclone to other proteins (i.e. globulins, lipoproteins). In fact, binding percentages towards both HSA and AGP (<25% for individual enantiomers and racemic compound) cannot explain the binding to total plasma proteins (>75% in all cases). In the case of chlorpheniramine, HSA, AGP and total plasma proteins showed similar stereoselectivity (Hiep et al., 1999). (+)(S)-chlorpheniramine is more extensively bound than its antipode to total plasma proteins (38 vs 23%), to HSA (20 vs 15%) and to AGP (23 vs 5%).

Whole Plasma Proteins Studies and Contribution of Individual Proteins to Total Plasma Enantioselective Protein Binding
As has been stated in the previous sections, albumin and 1-acid glycoprotein are responsible for the differential protein binding properties between enantiomers. In some cases HSA and AGP have shown higher affinity towards the same enantiomer, but in others the opposite stereoselectivity has been observed. On the other hand, despite the higher concentration of HSA (475 600 M) with respect to AGP (20 M) in plasma, there are many examples in which the interaction of enantiomers with AGP determines the total protein binding stereoselectivity observed. Fitos et al. (1989) investigated the stereoselective binding of rac-acenocoumarol to HSA and AGP by affinity chromatography and by combined ultrafiltration and circular dichroism methods. For HSA, the enantiomeric constants ratio was K(R)/K(S) = 2, while for AGP, K(S)/K(R) = 3. In a later work the authors studied, besides HSA and AGP, the stereoselectivity in the binding of acenocoumarol to human plasma by chiral HPLC analysis of the ultrafiltrates on a Chiral-AGP column (Fitos et al., 1993). The results confirmed the previously detected inverse stereoselectivities. In addition, it was observed that in plasma the contribution of HSA dominated, although in pathological states, elevated AGP levels may compensate for stereoselective distribution. For semotiadil [(R)-isomer, Ca channel blocker] and its antipode levosemotiadil [(S)-isomer, Ca- and Na-channel blocker] AGP and HSA have also shown opposite enantioselectivities (Rodriguez Rosas et al., 1999a, b). The total binding affinities (nK) ratios were nK(S)/nK(R) = 3 and nK(R)/nK(S) = 1.2 for HSA and AGP, respectively. Moreover, the total binding affinity of AGP was 110 times [for (R)-isomer] and 30 times [for (S)-isomer] stronger than that of HSA. However, in human plasma, the unbound concentration of (R)-isomer is lower than that of (S)-isomer (Rodriguez Rosas

Other Whole Plasma Protein Studies


The binding of the enantiomers of indobufen, a non-steroidal anti-inflammatory drug, to human serum proteins has been investigated using the racemic mixture or the pure (+)-(S)enantiomer by means of ultrafiltration combined with HPLC chiral analysis (Glwka and Caldwell, 2002). Using racemic indobufen, the binding parameters of the two enantiomers were different, showing enantioselectivity in protein binding. The ()(R)-enantiomer was bound more strongly to HSA than the (+)(S)-enantiomer. When the binding of (+)-(S)-enantiomer was studied alone, the association constant was lower and the number of binding sites increased; therefore competition between the enantiomers occurred. In addition, the pharmacokinetics of free (unbound) and total indobufen enantiomers was studied

234

www.interscience.wiley.com/journal/bmc

Copyright 2008 John Wiley & Sons, Ltd.

Biomed. Chromatogr. 2009; 23: 225238

Determination of urinary androgen glucuronides following administration of a single oral dose of rac-indobufen to healthy volunteers and patients with obliterative atherosclerosis. In healthy volunteers the (+)-(S)-enantiomer, owing to its weaker binding to serum proteins, was eliminated faster than its ()-(R)-antipode. The mean unbound fraction of ()-(R)- and (+)(S)-indobufen was 0.45 and 0.43%, respectively. Levels of the free (+)-(S)-enantiomer were higher than its ()-(R)-antipode at steady state in patients with obliterative atherosclerosis (Glwka and Caldwell, 2002). Martnez-Gmez et al. (2007b) studied the enantioselective binding of five antihistamines (brompheniramine, chlorpheniramine, hydroxyzine, orphenadrine and phenindamine), two phenothiazines (promethazine and trimeprazine) and a local anaesthetic (bupivacaine) to whole plasma by determining the protein binding of drug enantiomers. Owing to the high protein binding values of these drugs the authors used the methodology previously described (Martnez-Gmez et al., 2007a). The enantioselectivity values, defined as the ratio between the bound drug concentrations of second and first eluted enantiomers were ranged from 1.01 to 2.51 indicating that a different degree of enantioselectivity in the binding of these compounds to plasma proteins exists. The observed decreased order of enantioselectivity was: phenindamine > trimeprazine > promethazine orphenadrine > bupivacaine >> chlorpheniramine hydroxyzine brompheniramine. Orphenadrine and trimeprazine showed concentration-dependent enantioselectivity and extensive binding to human plasma. 5.75 105 and 2.10 105 M1, respectively (Paliwal et al., 1993) and that the binding constants of (R)-ibuprofen and (S)-ibuprofen to HSA are 5.3 105 and 1.1 105 M1, respectively (Hage et al., 1995). The rat, dog and human protein binding of two aminohydantoin compounds, potential drug candidates for the treatment of cardiac antinfective and musculoskeletal diseases, has been shown to be enantioselective and species-dependent with higher protein binding in human plasma than in rat and dog plasma (Peng et al., 1999). However, for both compounds, the protein binding percentage and binding association constant were found to be greater for the (R)-enantiomer than for the (S)enantiomer for all species. The binding between BOF-4272, a new xanthine oxidase inhibitor, and several serum albumins has shown to be enantioselective and species-dependent, being stronger in the order rat < human < bovine (Shibukawa et al., 1995). The binding of BOF4272 with rat and bovine serum albumins exhibited slight but significant reversed enantioselectivity. Interestingly, the enantioselectivity of the binding between BOF-4272 and HSA changed depending on the total drug concentration. It was found that this change is due to the enantiomeric difference in the binding constant and the number of binding site per protein molecule. Species-dependency on the stereoselective binding of benzodiazepine and coumarin drugs was studied by chiral chromatographic techniques (Fitos et al., 2002). The applied methods were affinity chromatography on the albumins immobilized on Sepharose 4B, HSA and bovine serum albumin chiral HPLC columns, and chiral HPLC analysis of ultrafiltrates of solutions containing the racemic drug and the native protein. Significant differences between the species were found related to preferred configurations and conformations. The binding stereoselectivity of the 2,3-benzodiazepine drug, tofisopam, in human, was found to be opposite to that in all other species. In the binding of 1,4benzodiazepines, dog albumin was found to be very similar to HSA. Highly preferred binding of (S)-phenprocoumon was found with dog albumin. The binding of bimoclomol enantiomers to human, monkey, dog, rat and mouse plasma was investigated by equilibrium dialysis using individual radioactive biclomol enantiomers or ultrafiltration (Visy et al., 2002). Comparative binding studies indicated considerable species dependency. It was observed that the ()enantiomer was preferably bound in each case, but both the degree of binding and the stereoselectivity were different in the plasma of the species investigated. Species dependency has also been reported for other drugs which are preferably bound to AGP (Lima and Haughey, 1981). Gupta et al. (2006) characterized the pharmacokinetics of cetirizine enantiomers. Moreover, protein binding in both the guinea pig and human plasma were performed by means of equilibrium dialysis combined with HPLC analysis using a chiral-AGP column. The cetirizine pharmacokinetics was found to be stereoselective in the guinea pig, the clearance and steady-state volume of distribution for (S)-cetirizine being half those for (R)-cetirizine, approximately. In the protein binding experiment the unbound fraction for (S)-cetirizine in guinea pig plasma was found to be 2.12.3 times higher than for (R)-cetirizine, indicating that protein binding is the primary factor affecting pharmacokinetics of cetirizine enantiomers. Protein binding of cetirizine in human plasma was stereoselective, with the unbound fraction of (S)enantiomer being 1.5 times higher than that for (R)-enantiomer. The effect of protein binding on the pharmacokinetics of the cetirizine enantiomers could be extrapolated to humans.

Species Dependency on Stereoselective DrugProtein Binding Studies


Proteins of different animal species sometimes exhibit different binding abilities, because the amino acid sequence differs somewhat among animals. The recognition of eventual stereoselective plasma binding differences between humans and other species commonly used in pharmacokinetic studies is of utmost importance, since experimental data in animals has to be extrapolated to humans. Stereoselective differences in protein binding between different species, even opposite stereoselectivity, have been described for the enantiomers of propranolol, verapamil, warfarin, ofloxacin, leukotriene D4-antagonist, MK-571, carbonic anhydrase inhibitor, MK-927 and disopyramide, which have been reviewed in Oravcov et al. (1996). Further studies with other compounds had also shown interspecies discrepancies (Pistolozzi and Bertucci, 2008), but also a good agreement between human and other mammalian results has been reported. Ibuprofen is a non-steroidal antiinflammatory drug for which it is reported that unidirectional conversion from (R)-ibuprofen to (S)-ibuprofen takes place in vivo in humans (Baillie et al., 1989; Rudy et al., 1991) rats (Chen et al., 1990; Itoh et al., 1997) and several other animal species (Chen et al., 1991). Itoh et al. (1997) measured protein binding of individual ibuprofen enantiomers to rat plasma. The number of binding sites and the binding constants for each enantiomer to the albumin of rat plasma samples were: n = 1.00 and K = 1.21 105 M1 for (R)-ibuprofen, and n = 0.91 and K = 5.76 104 M1 for (S)-ibuprofen. Stereoselectivity in plasma protein binding was shown to be similar in rats and humans [the affinity of (R)-enantiomer was higher than that of (S)-enantiomer in both plasma cases], but the binding constants of ibuprofen enantiomers obtained in rat plasma were slightly smaller than those reported in humans. It is reported that the binding constants of (R)-ibuprofen and (S)-ibuprofen in human plasma are

235

Biomed. Chromatogr. 2009; 23: 225238

Copyright 2008 John Wiley & Sons, Ltd.

www.interscience.wiley.com/journal/bmc

L. Escuder-Gilabert et al.

Conclusions
There is a growing interest in the evaluation of the differential pharmacokinetic and pharmacodynamic profiles between enantiomers due to the undoubted great impact of stereochemistry on the pharmacology of chiral drugs. Among the different pharmacological parameters to be assessed, protein binding of enantiomers is a crucial factor affecting the pharmacokinetics, pharmacodinamics and toxicity of chiral drugs since it determines the free drug concentration. Several methodologies have been reported in the literature to evaluate the stereoselective binding of chiral xenobiotics to plasma proteins. Most of them are based on the integration of separation methodologies commonly used to estimate protein binding of xenobiotics, such as equilibrium dialysis or ultrafiltration, with methodologies used for the separation of enantiomers, such as electrophoretic or chromatographic techniques using several chiral selectors. Among them, ultrafiltration (combined with chiral chromatographic or electrophoretic methodologies) seems to be the preferred choice due to its simplicity and speed. Moreover, the commersialitzation of 96-well plate format devices specifically optimized for in vitro plasma protein binding assays allows high-throughput estimations as demanded today by the pharmaceutical industry. Other methodologies, such as affinity chromatography (or biointeraction chromatography) and affinity capillary electrophoresis/frontal analysis, allow the simultaneous assessment of proteinchiral drug interaction and enantiomer resolution in a single step. However, a major concern of these methodologies is the alteration of the native protein binding properties due to the immobilization of the protein or the presence of the chiral selector in affinity chromatography and affinity capillary electrophoresis, respectively. Regarding the stereoselectivity of plasma proteins, albumin and 1-acid glycoprotein have been shown to be responsible for the differential protein binding properties between enantiomers. On the contrary, although lipoproteins are important binding proteins to neutral and lipophilic drugs, they have not been shown to exhibit stereoselectivity in binding, at least for the compounds studied. On the other hand, the relative contribution of albumin and 1-acid glycoprotein to the stereoselective binding of the whole plasma proteins its highly determined by the higher concentration of albumin in plasma but also by the affinity degree of each enantiomer towards the individual proteins; in fact there are many examples in which the interaction of enantiomers with 1-acid glycoprotein determines the total protein binding stereoselectivity observed. Regarding the interspecies effect on protein binding stereoselectivity, there is no general rule or behavior. In some cases a good agreement has been observed between humans and other mammalian results, but in many cases different affinity degree or even the opposite stereoselectivity has been reported. Therefore, since it is not always possible to extrapolate to humans results obtained from other mammalian species, animals chosen for pharmacokinetic, pharmacodynamic and toxicological assessment in drug development must be carefully selected. Acknowledgements The authors acknowledge the Spanish Ministry of Science and Technology (MCYT) and the European Regional Development Fund (ERDF) (project SAF2005-01435) for financial support. L.

Escuder-Gilabert is grateful to the Generalitat Valenciana for the grant (APOSTD/2007/062).

References
Ascoli GA, Domenici E and Bertucci C. Drug binding to human serum albumin: abridged review of results obtained with high-performance liquid chromatography and circular dichroism. Chirality 2006; 18: 667 679. Baillie TA, Adams WJ, Kaiser DG, Olanoff LS, Halstead GW, Harpootlian H and Van Giessen GJ. Mechanistic studies of the metabolic chiral inversion of (R)-ibuprofen in humans. Journal of Pharmacology and Experimental Therapeutics 1989; 249: 517523. Busch MHA, Carells LB, Boelens HFM, Kraak JC and Poe H. Comparison of five methods for the study of drug protein binding in affinity capillary electrophoresis. Journal of Chromatography A 1997; 777: 311328. Chen CS, Chen TL and Shieh W-R. Metabolic stereoisomeric inversion of 2-arylpropionic acids. On the mechanism of ibuprofen epimerization in rats. Biochimica et Biophysica Acta 1990; 1033: 16. Chen CS, Shieh WR, Lu P-H, Harriman S and Chen C-Y. Metabolic stereoisomeric inversion of ibuprofen in manmmals. Biochimica et Biophysica Acta 1991; 1078: 411 417. Chen J and Hage DS. Quantitative analysis of allosteric drugprotein binding by biointeraction chromatography. Nature Biotechnology 2004; 22: 14451448. Chen J, Fitos I and Hage DS. Chromatographic analysis of allosteric effects between ibuprofen and benzodiazepines on human serum albumin. Chirality 2006; 18: 24 36. Chuang VTG and Otagiri M. Stereoselective binding of human serum albumin. Chirality 2006; 18: 159166. Dente L, Pizza MG, Metspalu A and Cortese R. Structure and expression of the genes-coding for human 1-acid glycoprotein. EMBO Journal 1987; 6: 22892296. Ding Y, Zhu X and Lin B. Study of interaction between drug enantiomers and serum albumin by capillary electrophoresis. Electrophoresis 1999; 20: 18901894. Evans AM, Nation Rl, Sansom LN, Bochner F and Somogyi AA. Stereoselective plasmaprotein binding of ibuprofen enantiomers. European Journal of Clinical Pharmacology 1989; 36: 283290. Evans WE, Schentag JJ, Jusko WJ and Relling MW. Applied Pharmacokinetics. Principles of Therapeutic Drug Monitoring. Edwards Brothers: Vancouver, 1992. Fernandez C, Gimenez F, Thuillier A and Farinotti R. Stereoselective binding of zopiclone to human plasma proteins. Chirality 1999; 11: 129132. Fitos I, Visy J, Magyara A, Kajtr J and Simonyi M. Inverse stereoselectivity in the binding of acenocoumarol to human serum albumin and to 1-acid glycoprotein. Biochemical Pharmacology 1989, 38: 22592262. Fitos I, Visy J, Simonyi M and Hermansson J. Stereoselective distribution of acenocoumarol enantiomers in human plasmachiral chromatographic analysis of the ultrafiltrates. Chirality 1993; 5: 346349. Fitos I, Visy J, Simonyi M and Hermansson J. Separation of enantiomers of benzodiazepines on the Chiral-AGP column. Journal of Chromatography A 1995; 709: 265273. Fitos I, Visy J, Simonyi M and Hermansson J. Stereoselective allosteric binding interaction on human serum albumin between ibuprofen and lorazepam acetate. Chirality 1999; 11: 115120. Fitos I, Visy J and Simonyi M. Species-dependency in chiraldrug recognition of serum albumin studied by chromatographic methods. Journal Biochemical Biophysical Methods 2002; 54: 7184. Glwka FK and Caldwell J. Protein binding of indobufen enantiomers: pharmacokinetics of free fraction-studies after single or multiple doses of rac-indobufen. Chirality 2002; 14: 736741. Gonzlez-Alonso I and Snchez-Navarro A. Biofarmacia y Farmacocintica, Vol. II. Sntesis: Madrid, 1998; 467. Gupta A, Hammarlund-Udenaesa M, Chatelain P, Massingham R and Jonsson EN. Stereoselective pharmacokinetics of cetirizine in the guinea pig: role of protein binding. Biopharmaceutics and Drug Disposition 2006; 27: 291297. Hage DS, Noctor TAG and Wainer IW. Characterization of the protein binding of chiral drugs by high-performance affinity chromatography: interactions of R- and S-ibuprofen with human serum albumin. Journal of Chromatography A 1995; 693: 2332. Hazai E, Visy J, Fitos I, Bikdi Z and Simonyi M. Selective binding of

236

www.interscience.wiley.com/journal/bmc

Copyright 2008 John Wiley & Sons, Ltd.

Biomed. Chromatogr. 2009; 23: 225238

Determination of urinary androgen glucuronides


coumarin enantiomers to human 1-acid glycoprotein genetic variants. Bioorganic and Medicinal Chemistry 2006; 14: 19591965. He JY, Shibukawa A, Tokunaga S and Nakagawa T. Protein-binding highperformance frontal analysis of (R)- and (S)-warfarin on HSA with and without phenylbutazone. Journal of Pharmaceutical Sciences 1997; 86: 120125. Herv F, Caron G, Duche JC, Gaillard P, Rahman NA, Tsantili-Kakoulidou A, Carrupt PA, dAthis P, Tillement JP and Testa B. Ligand specificity of the genetic variants of human 1-acid glycoprotein: generation of a three-dimensional quantitative structureactivity relationship model for drug binding to the a variant Molecular Pharmacology 1998; 54: 129138. Hiep BT, Gimenez F, Khanh V, Hung NK, Thuillier A, Farinotti R and Fernandez C. Binding of chlorpheniramine enantiomers to human plasma proteins. Chirality 1999; 11: 501504. Imamura H, Komori T, Ismail A, Suenaga A and Otagiri M. Stereoselective protein binding of alprenolol in the renal diseased state. Chirality 2002; 14: 599603. Itoh T, Nakashima K, Tsuda Y and Yamada H. Stereoselective binding of carbenicillin epimers to human serum albumin. Chirality 1996; 8: 201 206. Itoh T, Maruyama J, Tsuda Y and Yamada H. Stereoselective pharmacokinetics of ibuprofen in rats: effect of enantiomerenantiomer interaction in plasma protein binding. Chirality 1997; 9: 354361. Jin Y-X, Tang Y-H and Zeng S. Analysis of flurbiprofen, ketoprofen and etodolac enantiomers by pre-column derivatization RP-HPLC and application to drugprotein binding in human plasma. Journal of Pharmaceutical and Biomedical Analysis 2008; 46: 953958. Kariv I, Cao H and Oldenburg KR. Development of a high throughput equilibrium dialysis method. Journal of Pharmaceutical Sciences 2001; 90: 580587. Kimura T, Shibukawa A and Matsuzaki K. Biantennary glycans as well as genetic variants of 1-acid glycoprotein control the enantioselectivity and binding affinity of oxybutynin. Pharmaceutical Research 2006; 23: 10381042. Kwong TC. Free drug measurements: methodology and clinical significance. Clinica Chimica Acta 1985; 151: 193216. Lagrange F, Phourcq F, Matoga M and Bannwarth B. Binding of ketoprofen enantiomers in various human albumin preparations. Journal of Pharmaceutical and Biomedical Analysis 2000; 23: 793802. Lee K-J, Mower R, Hollenbeck T, Castelo J, Johnson N, Gordon P, Sinko PJ, Holme K and Lee Y-H. Modulation of nonspecific binding in ultrafiltration protein binding studies. Pharmaceutical Research 2003; 20: 10151021. Lima JJ and Haughey DB. Disopyramide binding to serum protein in man and animals. Drug Metabolism and Disposition 1981; 9: 582583. Martnez-Gmez MA, Sagrado S, Villanueva-Camaas RM and MedinaHernndez MJ. Evaluation of enantioselective binding of antihistamines to human serum albumin by ACE. Electrophoresis 2007a; 28: 2635 2643. Martnez-Gmez MA, Sagrado S, Villanueva-Camaas RM and MedinaHernndez MJ. Evaluation of enantioselective binding of basic drugs to plasma by ACE., Electrophoresis 2007b; 28: 30563063. Mohamed NAL, Kuroda Y, Shibukawa A, Nakagawa T, El Gizawy S, Askal HF and El Kommos ME. Binding analysis of nilvadipine to plasma lipoproteins by capillary electrophoresisfrontal analysis. Journal of Pharmaceutical and Biomedical Analysis 1999; 21: 10371043. Mohamed NAL, Kuroda Y, Shibukawa A, Nakagawa T, El Gizawy S, Askal HF and El Kommos ME. Enantioselective binding analysis of verapamil to plasma lipoproteins by capillary electrophoresisfrontal analysis. Journal of Chromatography A 2000; 875: 447453. Nakagawa T, Kishino S, Itoh S, Sugawara M and Miyazaki K. Differential binding of disopyramide and warfarin enantiomers to human 1-acid glycoprotein variants. British Journal of Clinical Pharmacology 2003; 56: 664669. Ohara T, Shibukawa A and Nakagawa T. Capillary electrophoresis/frontal analysis for microanalysis of enantioselective protein binding of a basic drug Analytical Chemistry 1995; 67: 35203525. Ohnishi T, Mohamed NAL, Shibukawa A, Kuroda Y, Nakagawa T, El Gizawy S, Askal H and El Kommos ME. Frontal analysis of drugplasma lipoprotein binding using capillary electrophoresis. Journal of Pharmaceutical and Biomedical Analysis 2002; 27: 607614. Oravcov J, Bhs B and Lindner W. Drugprotein binding studies. New trends in analytical and experimental methodology. Journal of Chromatography B 1996; 677: 128. Paliwal JK, Smith DE, Cox SR, Berardi RR, Dunn-Kucharski VA and Elta GH. Stereoselective, competitive, and nonlinear plasma protein binding of ibuprofen enantiomers as determined in vivo in healthy subjects. Journal of Pharmacokinetics and Biopharmaceutics 1993; 21: 145161. Peng SX, Henson C and Wilson LJ. Simultaneous determination of enantioselective plasma protein binding of aminohydantoins by ultrafiltration and chiral high-performance liquid chromatography. Journal of Chromatography B 1999; 732: 3137. Perrin JH. Circular dichroic investigation of binding of fenoprofen, 2(3phenoxyphenyl)-propionic acid, to human-serum albumin. Journal of Pharmacy and Pharmacology 1973; 25: 208212. Pistolozzi M and Bertucci C. Species-dependent stereoselective drug binding to albumin: a circular dichroism study. Chirality 2008; 20: 552 558. Randon J, Garnier F, Rocca JL and Masterrena B. Optimization of the enantiomeric separation of tryptophan analogs by membrane processes. Journal of Membrane Science 2000; 175: 111117. Rodriguez Rosas ME, Shibukawa A, Ueda K and Nakagawa T. Enantioselective protein binding of semotiadil and levosemotiadil determined by high-performance frontal analysis. Journal of Pharmaceutical and Biomedical Analysis 1997; 15: 15951601. Rodriguez Rosas ME, Shibukawa A, Yoshikawa Y, Kuroda Y and Nakagawa T. Binding study of semotiadil and levosemotiadil with 1-acid glycoprotein using high-performance frontal analysis. Analytical Biochemistry 1999a; 274: 2733. Rodriguez Rosas ME, Shibukawa A, Yoshikawa Y, Kuroda Y and Nakagawa T. Binding study of semotiadil and levosemotiadil with human serum albumin using high-performance frontal analysis. Analytical Sciences 1999b; 15: 217222. Rudy AC, Knight PM, Brater DC and Hall SD. Stereoselective metabolism of ibuprofen in humans: Administration of R-, S- and racemic ibuprofen. Journal of Pharmacology and Experimental Therapeutics 1991; 259: 11331139. Shibukawa A, Terakita A, He JY and Nakagawa T. High-performance frontal analysis high-performance liquid-chromatographic system for stereoselective determination of unbound ketoprofen enantiomers in plasma after direct sample injection. Journal of Pharmaceutical Sciences 1992; 81: 710715. Shibukawa A, Nagao M, Terakita A, He JY and Nakagawa T. Highperformance frontal analysis high-performance liquidchromatographic system for the enantioselective determination of unbound fenoprofen concentration in protein-binding equilibrium. Journal of Liquid Chromatography 1993; 16: 903914. Shibukawa A, Nakao C, Sawada T, Terakita A, Morokoshi N and Nakagawa T. Determination of the unbound concentration of hydrophobic drugs in albumin solutions by high-performance frontal analysis using a diol-silica column. Journal of Pharmaceutical Sciences 1994; 83: 868873. Shibukawa A, Kadohara M, He J-y, Nishimura M, Naito S and Nakagawa T. Study of the enantioselective binding between BOF-4272 and serum albumins by means of high-performance frontal analysis. Journal of Chromatography A 1995; 694: 8189. Shibukawa A, Kuroda Y and Nakagawa T. High-performance frontal analysis for drugprotein binding study. Journal of Pharmaceutical and Biomedical Analysis 1999a; 18: 10471055. Shibukawa A, Kuroda Y and Nakagawa T. Development of highperformance frontal analysis and the application to the study of drugplasma protein binding. TrAC-Trends in Analytical Chemistry 1999b; 18: 549556. Shibukawa A, Ishizawa N, Kimura T, Sakamoto Y, Ogita K, Matsuo Y, Kuroda Y, Matayatsuk C, Nakagawa T and Wainer IW. Plasma protein binding study of oxybutynin by high-performance frontal analysis. Journal of Chromatography B 2002a; 768: 177188. Shibukawa A, Yoshikawa Y, Kimura T, Kuroda Y, Nakagawa T and Wainer IW. Binding study of desethyloxybutynin using high-performance frontal analysis method. Journal of Chromatography B 2002b; 768: 189197. Shiono H, Shibukawa A, Kuroda Y and Nakagawa T. Effect of sialic acid residues of human 1-acid glycoprotein on stereoselectivity in basic drugprotein binding. Chirality 1997; 9: 291296. Tsuda Y, Tsunoi T, Watanabe N, Ishida M, Yamada H and Itoh T. Stereoselective binding and degradation of sulbenicillin in the presence of human serum albumin. Chirality 2001; 13: 236243. Urien S. In: ProteinBinding and Drug Transport, Tillement JP, Lindenlaub E (eds). Schattauer: Stuttgart, 1986; 6375.

237

Biomed. Chromatogr. 2009; 23: 225238

Copyright 2008 John Wiley & Sons, Ltd.

www.interscience.wiley.com/journal/bmc

L. Escuder-Gilabert et al.
Visy J, Fitos I, Mdy G, rge L, Krajcsi P and Simonyi M. Enantioselective plasma protein binding of bimoclomol. Chirality 2002; 14: 638642. Ye M, Zou H, Liu Z, Wu R, Lei Z and Ni J. Study of competitive binding of entantiomers to protein by affinity capillary electrochromatograpy. Journal of Pharmaceutical and Biomedical Analysis 2002; 27: 651660. Yuasa I, Umetsu K, Vogt U, Nakamura H, Nanba E, Tamaki N and Irizawa Y. Human orosomucoid polymorphism: molecular basis of the three common ORM1 alleles, ORM1*F1, ORM1*F2, and ORM1*S. Human Genetics 1997; 99: 393398. Zsila F, Visy J, Mdy G and Fitos I. Selective plasma protein binding of antimalarial drugs to 1-acid glycoprotein. Bioorganic and Medicinal Chemistry 2008; 16: 37593772.

238
www.interscience.wiley.com/journal/bmc Copyright 2008 John Wiley & Sons, Ltd. Biomed. Chromatogr. 2009; 23: 225238

You might also like