09.06.2013 Views

Miniaturization of Drug Solubility and Dissolution Testings - Doria

Miniaturization of Drug Solubility and Dissolution Testings - Doria

Miniaturization of Drug Solubility and Dissolution Testings - Doria

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Division <strong>of</strong> Pharmaceutical Technology<br />

Faculty <strong>of</strong> Pharmacy<br />

University <strong>of</strong> Helsinki<br />

<strong>Miniaturization</strong> <strong>of</strong> <strong>Drug</strong> <strong>Solubility</strong> <strong>and</strong> <strong>Dissolution</strong><br />

<strong>Testings</strong><br />

Tiina Heikkilä<br />

ACADEMIC DISSERTATION<br />

To be presented, with the permission <strong>of</strong> the Faculty <strong>of</strong> Pharmacy <strong>of</strong> the University <strong>of</strong><br />

Helsinki, for public examination in auditorium 1041, Viikki Biocenter 2 (Viikinkaari 5 E),<br />

on 11 th June 2010, at 12 noon.<br />

Helsinki 2010


Supervisors: Docent Leena Peltonen<br />

Division <strong>of</strong> Pharmaceutical Technology<br />

Faculty <strong>of</strong> Pharmacy<br />

University <strong>of</strong> Helsinki<br />

Helsinki, Finl<strong>and</strong><br />

Pr<strong>of</strong>essor Marjo Yliperttula<br />

Division <strong>of</strong> Biopharmacy <strong>and</strong> Pharmacokinetics<br />

Faculty <strong>of</strong> Pharmacy<br />

University <strong>of</strong> Helsinki<br />

Helsinki, Finl<strong>and</strong><br />

Pr<strong>of</strong>essor Jouni Hirvonen<br />

Division <strong>of</strong> Pharmaceutical Technology<br />

Faculty <strong>of</strong> Pharmacy<br />

University <strong>of</strong> Helsinki<br />

Helsinki, Finl<strong>and</strong><br />

Reviewers: Pr<strong>of</strong>essor Jukka Rantanen<br />

Department <strong>of</strong> Pharmaceutics <strong>and</strong> Analytical Chemistry<br />

Faculty <strong>of</strong> Pharmaceutical Sciences<br />

University <strong>of</strong> Copenhagen<br />

Copenhagen, Denmark<br />

Associate Pr<strong>of</strong>essor Christos Reppas<br />

Department <strong>of</strong> Pharmaceutical Technology<br />

Faculty <strong>of</strong> Pharmacy<br />

University <strong>of</strong> Athens<br />

Athens, Greece<br />

Opponent: Doctor Juha Kiesvaara<br />

Head <strong>of</strong> Formulation Research<br />

Orion Pharma R&D<br />

Turku, Finl<strong>and</strong><br />

© Tiina Heikkilä<br />

ISBN 978-952-10-6190-5 (paperback)<br />

ISBN 978-952-10-6191-2 (PDF, http://e-thesis.helsinki.fi)<br />

ISSN 1795-7079<br />

Helsinki University Print<br />

Helsinki, Finl<strong>and</strong>, 2010


Abstract<br />

<strong>Solubility</strong> <strong>and</strong> drug dissolution are <strong>of</strong> crucial importance for drug formulations. Poor<br />

water-solubility <strong>of</strong> drug c<strong>and</strong>idates is a major obstacle in drug development, since the oral<br />

route is the most patient convenient <strong>and</strong> cost effective way to deliver drugs. In some cases<br />

the low aqueous solubility may limit the bioavailability when the absorption <strong>of</strong> the drug is<br />

dissolution limited. About 40% <strong>of</strong> the current lead optimization compounds suffer from<br />

poor solubility.<br />

Improvements in drug solubility/dissolution testing technologies (e.g. high throughput<br />

screening, HTS) can enhance the possibilities <strong>of</strong> the lead compounds to success in the later<br />

stages <strong>of</strong> drug development process. Typically HTS protocols measure the kinetic<br />

solubility involving co-solvents (e.g. dimethyl sulfoxide), which might enhance the in<br />

vitro solubility or give erroneous results if the potential drug c<strong>and</strong>idates are eliminated.<br />

Also, measurement <strong>of</strong> dissolution pr<strong>of</strong>iles is not available by the present HTS methods.<br />

For these reasons, there is a need for improvements in HTS solubility formats.<br />

Traditionally the in vitro dissolution tests are studied by pharmacopoeial methods,<br />

which are not utilizable in the drug discovery stage because <strong>of</strong> the large amount <strong>of</strong><br />

compounds needed. However, dissolution studies at the drug discovery stage could be<br />

useful e.g. to classify compounds based on their dissolution rates. In addition, the initial<br />

steps <strong>of</strong> dissolution process might be lost by the regulatory dissolution methods.<br />

The aim <strong>of</strong> this thesis was to miniaturize traditional drug dissolution <strong>and</strong> solubility<br />

testing methods. Systematic down scaling <strong>of</strong> methods was done towards the development<br />

<strong>of</strong> both equilibrium <strong>and</strong> kinetic 96-well plate solubility/dissolution methods.<br />

<strong>Miniaturization</strong> <strong>of</strong> the regulatory dissolution methods <strong>and</strong> shake-flask solubility<br />

measurements on the 96-well plates was successful. 96-well plate methods for equilibrium<br />

drug solubilities, as well as for drug dissolution pr<strong>of</strong>iles as a function <strong>of</strong> time, were<br />

developed. The former method is the first true equilibrium solubility method, which can<br />

be used in the screening <strong>of</strong> drug solubility <strong>and</strong> dissolution phenomena at the early stages<br />

<strong>of</strong> drug development process. This method was also tested using fasted state human<br />

intestinal fluid as a medium for the first time. Human intestinal fluid <strong>and</strong> data obtained<br />

might turn out to be important for very low water-solubility compounds. Surface tension<br />

based microtensiometry was also presented as an alternative method for kinetic HTS <strong>of</strong><br />

drug solubility properties, e.g. for classifying solubility <strong>of</strong> compounds not suitable for UVanalysis.<br />

Channel flow methodology was introduced enabling especially the kinetic<br />

follow-up at the very beginning <strong>of</strong> the dissolution process.<br />

This thesis provides directly applicable new miniaturized methods for the drug<br />

solubility <strong>and</strong> dissolution experiments. These methods enhance the throughput <strong>and</strong><br />

underst<strong>and</strong>ing <strong>of</strong> drug solubility/dissolution phenomena <strong>and</strong> pr<strong>of</strong>iles in drug discovery <strong>and</strong><br />

improve success in the later stages <strong>of</strong> the drug development process.


Acknowledgements<br />

This work was carried out at the Division <strong>of</strong> Pharmaceutical Technology, Faculty <strong>of</strong><br />

Pharmacy, University <strong>of</strong> Helsinki, Finl<strong>and</strong>.<br />

I wish to thank all the people who have been involved with this study.<br />

First, I would like to express my warmest gratitude to my supervisors Pr<strong>of</strong>essor Jouni<br />

Hirvonen, Docent Leena Peltonen <strong>and</strong> Pr<strong>of</strong>essor Marjo Yliperttula. In particular, I want to<br />

thank Leena for her constant patience <strong>and</strong> practical support. You were precent when<br />

needed <strong>and</strong> I felt like no stupid questions were. I wish everybody had a supervisor like<br />

you! Jouni <strong>and</strong> Marjo, first <strong>of</strong> all, thank you for this opportunity. Thank you also for your<br />

comments to all the versions <strong>of</strong> manuscripts <strong>and</strong> reports I have written, your scientifical<br />

guidance <strong>and</strong> providing excellent working facilities.<br />

The reviewers <strong>of</strong> the thesis, Pr<strong>of</strong>essor Jukka Rantanen <strong>and</strong> Associate Pr<strong>of</strong>essor<br />

Christos Reppas, are acknowledged for their careful work <strong>and</strong> valuable comments on<br />

impoving the manuscrpit.<br />

I am also grateful to all my co-authors, Pr<strong>of</strong>essor Patrick Augustijns, Doctor Milja<br />

Karjalainen, Pr<strong>of</strong>essor Kyösti Kontturi, Doctor Timo Laaksonen, Pr<strong>of</strong>essor Frank<br />

Lammert, Doctor Peter Liljeroth, M.Sc. Krista Ojala, M.Sc. Kirsi Partola, M.Sc. Saila<br />

Taskinen <strong>and</strong> Pr<strong>of</strong>essor Arto Urtti for their co-operation <strong>and</strong> scientific contributions.<br />

It has been very pleasent to work at Viikki because <strong>of</strong> all you co-workers. I felt warmly<br />

wellcome from the very beginning <strong>and</strong> I have got all the practical help I ever needed to<br />

perform this work.<br />

I would also like to thank my co-workers at the Lauttasaari Centralpharmacy for<br />

keeping me updated in practical pharmacy work. Especially, I want to thank pharmacy<br />

owner Liisa Heikkinen.<br />

In vitro in vivo relevance (IVIVRe) project in co-operation with Orion Pharma is<br />

acknowledged for the financial support.<br />

Finally, I would like to thank my most lowed ones. Especially, I would like to thank<br />

my Mum for introducing me the pharmacy <strong>and</strong> encouraging me to study this far, <strong>and</strong><br />

Sanna, thank you for endless support. Emmi, Elsa, Petter <strong>and</strong> Alpo, thank you for just<br />

being there, <strong>and</strong> Timo, thank you for your love all the way. I made it!<br />

Espoo, May 2010<br />

Tiina Heikkilä


Contents<br />

Abstract 3<br />

Acknowledgements 4<br />

List <strong>of</strong> original publications 7<br />

Abbreviations 8<br />

1 Introduction 9<br />

2 Review <strong>of</strong> the literature 11<br />

2.1 <strong>Solubility</strong> in drug discovery <strong>and</strong> development 11<br />

2.1.1 Kinetic solubility 13<br />

2.1.1.1 Dimethyl sulfoxide (DMSO) as a co-solvent 13<br />

2.1.1.2 Kinetic solubility assays 14<br />

2.1.2 Equilibrium solubility 15<br />

2.2 In vitro dissolution methods for solid dosage forms 16<br />

2.2.1 Regulatory dissolution methods 18<br />

2.2.2 Initial steps <strong>of</strong> the dissolution process 19<br />

2.2.3 In vivo mimicking <strong>of</strong> dissolution methods 22<br />

2.3 In vivo dissolution 23<br />

2.3.1 Effect <strong>of</strong> pH 23<br />

2.3.2 Role <strong>of</strong> bile acids <strong>and</strong> phospholipids 24<br />

2.3.3 Hydrodynamics in gastrointestinal (GI) tract 25<br />

2.4 In vitro in vivo (IVIV) correlation 26<br />

2.4.1 Biopharmaceutics Classification System (BCS) 26<br />

2.4.2 Biopharmaceutics <strong>Drug</strong> Disposition Classification System (BDDCS) 27<br />

2.4.3 BCS Biowaivers 28<br />

3 Aims <strong>of</strong> the study 30<br />

4 Experimental 31<br />

4.1 Materials 31<br />

4.1.1 Chemicals (I-IV) 31<br />

4.1.2 Human intestinal fluid (HIF) (IV) 31<br />

4.2 Methods 31<br />

4.2.1 Preparation <strong>of</strong> tablets (I, II) 31<br />

4.2.2 <strong>Dissolution</strong> methods 32<br />

4.2.2.1 Basket method (I, II) 32


4.2.2.2 Intrinsic dissolution rate method (I, II) 33<br />

4.2.2.3 Channel flow method (I, II) 33<br />

4.2.2.4 Micro dissolution method (previously unpublished data) 33<br />

4.2.2.5 96-well plate method (IV) 34<br />

4.2.3 <strong>Solubility</strong> methods 34<br />

4.2.3.1. Shake-flask method (III, IV) 34<br />

4.2.3.2 96-well plate method (IV) 35<br />

4.2.3.3 Surface tension based kinetic HTS solubility method (III) 35<br />

4.2.4 Analyses 35<br />

4.2.4.1 Analyses <strong>of</strong> human intestinal fluid (HIF) (IV) 35<br />

4.2.4.2 Solid-state characterization (IV) 35<br />

5 Results <strong>and</strong> discussion 37<br />

5.1 <strong>Dissolution</strong> testing 37<br />

5.1.1 Channel flow method vs. the regulatory dissolution methods (I, II) 37<br />

5.1.2 Micro dissolution studies in different media (previously unpublished data) 40<br />

5.1.3 <strong>Dissolution</strong> pr<strong>of</strong>iles in 96-well plate (IV) 41<br />

5.1.4 Scaling down the dissolution testings (I, II, III, IV) 41<br />

5.2 <strong>Solubility</strong> screening 42<br />

5.2.1 HTS 96-well plate methods (III, IV) 42<br />

5.2.2 Polymorphic changes (IV) 47<br />

5.3 Human intestinal fluid (HIF) 49<br />

5.3.1 Characterization <strong>of</strong> fasted state HIF (IV) 49<br />

5.3.2 HIF in solubility testing (IV) 50<br />

6 Conclusions 52<br />

7 References 53


List <strong>of</strong> original publications<br />

This thesis is based on the following publications, which are referred to in the text by their<br />

respective roman numerals (I-IV):<br />

I Peltonen, L., Liljeroth, P., Heikkilä, T., Kontturi, K., Hirvonen, J., 2003.<br />

<strong>Dissolution</strong> testing <strong>of</strong> acetylsalicylic acid by a channel flow method –<br />

correlation to USP basket <strong>and</strong> intrinsic dissolution methods. European<br />

Journal <strong>of</strong> Pharmaceutical Sciences, 19, 395-401.<br />

II Peltonen, L., Liljeroth, P., Heikkilä, T., Kontturi, K., Hirvonen, J., 2004. A<br />

novel channel flow method in determination <strong>of</strong> solubility properties <strong>and</strong><br />

dissolution pr<strong>of</strong>iles <strong>of</strong> theophylline tablets. STP Pharma Sciences, 14, 389-<br />

394.<br />

III Heikkilä, T., Peltonen, L., Taskinen, S., Laaksonen, T., Hirvonen, J., 2008.<br />

96-well plate surface tension measurements for fast determination <strong>of</strong> drug<br />

solubility. Letters in <strong>Drug</strong> Design <strong>and</strong> Discovery, 5, 471-475.<br />

IV Heikkilä, T., Karjalainen, M., Ojala, K., Partola, K., Lammert, F.,<br />

Augustijns, P., Urtti, A., Yliperttula, M., Peltonen, L., Hirvonen, J., 2010.<br />

Equilibrium drug solubility measurements in 96-well plates. International<br />

Journal <strong>of</strong> Pharmaceutics, submitted.<br />

Reprinted with the permission from the Elsevier Ltd (I, IV), Edition de Santé (II), <strong>and</strong><br />

Bentham Science Publishers Ltd (III).<br />

7


Abbreviations<br />

BCS Biopharmaceutics Classification System<br />

BDDCS Biopharmaceutics <strong>Drug</strong> Disposition Classification System<br />

CFC channel flow cell<br />

CMC critical micelle concentration<br />

CSD Cambridge Structural Database<br />

DMSO dimethyl sulfoxide<br />

DSC differential scanning calorimetry<br />

EMEA European Medicines Agency<br />

FaSSIF Fasted state simulated intestinal fluid<br />

FDA US Food <strong>and</strong> <strong>Drug</strong> Administration<br />

GI gastrointestinal<br />

HIF human intestinal fluid<br />

HPCD hydroxypropyl-β-cyclodextrin<br />

HPLC high pressure liquid chromatography<br />

HTS high throughput screening<br />

IDR intrinsic dissolution rate<br />

Int.Ph. The International Pharmacopoeia<br />

IR immediate release<br />

IUPAC The International Union <strong>of</strong> Pure <strong>and</strong> Applied Chemistry<br />

IVIV in vitro in vivo<br />

JP Japanese Pharmacopoeia<br />

MAD maximum absorbable dose<br />

MCC microcrystalline cellulose<br />

Ph. Eur. The European Pharmacopoeia<br />

rpm rotations per minute<br />

SDS sodium dodecyl sulphate<br />

USP The United States Pharmacopeia<br />

WHO World Health Organization<br />

XRPD X-ray powder diffraction<br />

8


1 Introduction<br />

Oral route is the most patient convenient <strong>and</strong> cost effective way to deliver drugs.<br />

However, inadequate aqueous solubility <strong>and</strong> dissolution in intestines <strong>and</strong> colon limits the<br />

bioavailability <strong>of</strong> drugs. Nowadays even 40% <strong>of</strong> the lead optimization compounds are<br />

poorly water-soluble (Lipinski et al., 1997). Although the poor solubility can be<br />

occasionally overcome by novel formulations, e.g. oral lipid-based (Porter <strong>and</strong> Charman,<br />

2001; Hauss, 2007) or amorphous formulations (Kawakami 2009), or delivery systems,<br />

e.g. cyclodextrins (Szejtli, 1988; Brewster <strong>and</strong> L<strong>of</strong>tsson, 2007), it still increases the cost<br />

<strong>and</strong> time <strong>of</strong> the new drug product to enter the market. Instead, improvement <strong>of</strong> drug<br />

solubility/dissolution testing methods (e.g. by high throughput screening, HTS) in the drug<br />

discovery stage can enhance the possibilities <strong>of</strong> the lead compound to success in the later<br />

stages <strong>of</strong> the drug development process.<br />

HTS solubility methods measure the solubility <strong>of</strong> numerous compounds at small<br />

quantities (Pan et al., 2001; Chen et al., 2002b; Takano et al., 2006; Bard et al., 2008; Dai<br />

et al., 2008; Alelyunas et al., 2009). Discovery protocols typically measure the kinetic<br />

solubility involving small amounts <strong>of</strong> co-solvents like dimethyl sulfoxide (DMSO). The<br />

co-solvents might enhance the in vitro solubility due to the co-solvent effect (Chen et al.,<br />

2002b) <strong>and</strong> the solubility in vivo might be totally different. Especially, for the poorly<br />

soluble drugs the solubility can increase manifold with co-solvents. Also, erroneous<br />

decisions can be made if wrong drug c<strong>and</strong>idates are eliminated because <strong>of</strong> rapid uptake <strong>of</strong><br />

water by DMSO. When the water concentration in the DMSO solution gets too high, some<br />

lead compounds may fall out <strong>of</strong> solution <strong>and</strong> give a false negative result (Homon <strong>and</strong><br />

Nelson, 2006). For these reasons, there is a need for improvements in HTS solubility<br />

formats.<br />

<strong>Drug</strong> dissolution testing <strong>of</strong> solid dosage forms is a routine process in pharmaceutical<br />

industry to evaluate drug release characteristics <strong>and</strong> to predict in vivo behavior.<br />

Traditionally the in vitro dissolution tests are studied by pharmacopoeial methods (e.g.<br />

paddle <strong>and</strong> basket methods) in the product development process for quality control, e.g. to<br />

assess the lot-to-lot quality <strong>and</strong> to detect changes in the drug formulation or manufacturing<br />

process. In the development <strong>of</strong> the drug product dissolution tests have been used to study<br />

the excipients (Omelczuk <strong>and</strong> McGinity, 1993; Uesu et al., 2000), <strong>and</strong> also to determine<br />

bioequivalence. In the drug discovery process pharmacopoeial dissolution methods are not<br />

utilizable because <strong>of</strong> the large amount <strong>of</strong> lead compounds <strong>and</strong>/or drug c<strong>and</strong>idates needed,<br />

however, dissolution studies at the drug discovery stage could be used to e.g. to classify<br />

compounds based on their dissolution rates or to show whether the dissolution has taken<br />

place.<br />

Biopharmaceutics Classification System (BCS), which divides drugs into four classes<br />

according to their solubility <strong>and</strong> permeability properties, correlates in vitro drug product<br />

dissolution <strong>and</strong> in vivo bioavailability (Amidon et al., 1995). After BCS the meaning <strong>and</strong><br />

application <strong>of</strong> relevant solubility/dissolution media has become more important <strong>and</strong> a<br />

broad range <strong>of</strong> dissolution media have been developed, e.g. FaSSIF (fasted state simulated<br />

intestinal fluid) <strong>and</strong> FeSSIF (fed state simulated intestinal fluid) (Galia et al., 1998). Still,<br />

the most physiological medium for solubility <strong>and</strong> dissolution studies would be human<br />

9


intestinal fluid (HIF), because drugs are usually dissolved <strong>and</strong> absorbed in the small<br />

intestine. The miniaturized drug solubility <strong>and</strong> dissolution methods requested above<br />

should enable the use <strong>of</strong> HIF when appropiate.<br />

In this thesis, different approaches were studied to miniaturize solubility <strong>and</strong><br />

dissolution testing methods. The in vitro in vivo (IVIV) correlation <strong>of</strong> the solubility testing<br />

was studied by using different buffer solutions <strong>and</strong> fasted state human intestinal fluid as<br />

media. The literature review focuses on the different available solubility <strong>and</strong> dissolution<br />

methods <strong>and</strong> on the IVIV correlation.<br />

10


2 Review <strong>of</strong> the literature<br />

2.1 <strong>Solubility</strong> in drug discovery <strong>and</strong> development<br />

IUPAC (The International Union <strong>of</strong> Pure <strong>and</strong> Applied Chemistry) defines solubility as<br />

‘the analytical composition <strong>of</strong> a mixture or solution which is saturated with one <strong>of</strong> the<br />

components <strong>of</strong> the mixture or solution, expressed in terms <strong>of</strong> the proportion <strong>of</strong> the<br />

designated component in the designated mixture or solution’ (Lorimer <strong>and</strong> Cohen-Adad,<br />

2003). <strong>Solubility</strong> can further be defined in a more specific manner, like unbuffered,<br />

buffered <strong>and</strong> intrinsic solubility. Unbuffered solubility means the solubility <strong>of</strong> a saturated<br />

solution <strong>of</strong> the compound at the final pH <strong>of</strong> the solution. Buffered (apparent) solubility<br />

refers to solubility at a given pH, measured in a defined pH-buffered system, <strong>and</strong> usually<br />

does not perceive the influence <strong>of</strong> salt formation with counterions <strong>of</strong> the buffering system<br />

on the measured solubility value. Intrinsic solubility (So) is the solubility <strong>of</strong> the neutral<br />

form <strong>of</strong> an ionizable compound. For neutral compounds all three definitions coincidence.<br />

<strong>Solubility</strong> measurements determine either kinetic or equilibrium (thermodynamic)<br />

solubility depending on the experimental set-up (Alsenz <strong>and</strong> Kansy, 2007).<br />

Compounds can be divided into highly or poorly soluble ones. Traditionally<br />

compounds were considered to be poorly soluble if the solubility was less than 100 μg/ml.<br />

In 2006 Fligge <strong>and</strong> Schuler redefined the concept <strong>of</strong> poor solubility by HTS method; a<br />

compound is poorly soluble when the solubility is less than 20 μg/ml, partly soluble when<br />

the solubility is 20-80 μg/ml <strong>and</strong> soluble when the solubility is over 80 μg/ml. According<br />

to Lipinski (2004), 15% <strong>of</strong> the drug-like compounds <strong>and</strong> 40% <strong>of</strong> lead optimization<br />

compounds were insoluble at a concentration <strong>of</strong> less than or equal to 20 μg/ml. The<br />

definitions <strong>of</strong> drug-likeness <strong>and</strong> lead-likeness have been discussed e.g. by Oprea et al.<br />

(2001), but in general it can be said that in the lead optimization process a lead compound<br />

is moved towards a drug c<strong>and</strong>idate having drug-like properties. In the pharmacopoeias<br />

solubility is indicated by the descriptive terms (Table 1).<br />

Table 1. The descriptive terms for the solubility used in the pharmacopoeias (USP,<br />

2009;Ph.Eur., 2009).<br />

Term Parts <strong>of</strong> solvent required for 1 part <strong>of</strong> solute<br />

Very soluble<br />

Less than 1<br />

Freely soluble<br />

From 1 to 10<br />

Soluble<br />

From 10 to 30<br />

Sparingly soluble<br />

From 30 to 100<br />

Slightly soluble<br />

From 100 to 1000<br />

Very slightly soluble<br />

From 1000 to 10 000<br />

Practically insoluble, or insoluble<br />

Greater than or equal to 10 000<br />

11


The requirement to the drug c<strong>and</strong>idate’s solubility has been studied by the concept <strong>of</strong><br />

maximum absorbable dose (MAD) (Johnson <strong>and</strong> Swindell, 1996):<br />

MAD = S 6.5 VL k a t (1)<br />

The unit <strong>of</strong> MAD is milligram, the transit time, t, is 270 minutes <strong>and</strong> the estimated luminal<br />

volume, VL, is 250 ml. S 6.5 marks the solubility at pH 6.5 (mg/ml) <strong>and</strong> ka the<br />

transintestinal absorption rate constant (min -1 ). In the MAD equation the small intestine<br />

can be considered as a tube, which has the volume <strong>of</strong> 250 ml. The saturated drug<br />

suspension is set into that tube, <strong>and</strong> the timer is started. The absorption rate constant<br />

describes how fast the dissolved quantity <strong>of</strong> the drug ‘leaks’ out <strong>of</strong> the tube each minute.<br />

After 270 minutes the total amount <strong>of</strong> ‘leakage’ is the estimate <strong>of</strong> the amount <strong>of</strong> the drug<br />

that should be intestinally absorbed. If that estimated quantity is close to the entire drug<br />

dosage, the intestinal absorption should be 100%.<br />

Lipinski (2000) has developed further the MAD concept based on the most common<br />

clinical potency <strong>of</strong> drug dose, 1 mg/kg. According to Lipinski, in the case <strong>of</strong> permeability<br />

being the average, the equilibrium (thermodynamic) solubility should be 52 μg/ml at pH<br />

6.5-7. For the low permeability drugs the equilibrium solubility should be 207 μg/ml <strong>and</strong><br />

for the high permeability drugs 10 μg/ml, in order to be therapeutically effective. Low<br />

dose drugs (0.1 mg/kg) with high permeability can be effective even at the equilibrium<br />

solubility <strong>of</strong> 1 μg/ml <strong>and</strong>, on the opposite cases, 10 mg/kg dose drugs having low<br />

permeability properties might need equilibrium solubilities <strong>of</strong> as high as 2100 μg/ml to be<br />

effective. The advantage <strong>of</strong> the MAD model is that it easily shows the impact that the<br />

required dose might have to the permeability, solubility <strong>and</strong> fraction absorbed <strong>of</strong> the dosed<br />

compound.<br />

The solubility determinations are performed several times along the drug discovery<br />

<strong>and</strong> development processes, only varying the focus, the amount <strong>of</strong> the drug used <strong>and</strong> also<br />

the level <strong>of</strong> the purity <strong>and</strong> crystallinity <strong>of</strong> the drug c<strong>and</strong>idate. In the early drug discovery<br />

stage the goal is to rank compounds according to the solubility <strong>and</strong> to aid in the structure<br />

activity chemistry relationships; in the drug development stage the goal is a more accurate<br />

numerical solubility number (Lipinski, 2004). More precisely, during the lead<br />

identification the solubility assays are, e.g., used to recognize compounds with potential<br />

liabilities, while in the lead optimization solubility information is used to make informed<br />

decisions on potential development c<strong>and</strong>idates. Also the equilibrium solubility is used to<br />

confirm earlier kinetic solubility results in the lead optimization. In the drug development<br />

stage the dissolution testing is an important area for solubility assays (Dokoumetzidis <strong>and</strong><br />

Macheras, 2006). In the earlier drug development stages the goal is, e.g., to identify ratelimiting<br />

steps, while in the later stages the dissolution tests are used, e.g., in the quality<br />

control <strong>and</strong> bioequivalence studies. Figure 1 depicts the solubility <strong>and</strong> dissolution assays<br />

in the drug discovery <strong>and</strong> development processes.<br />

12


New<br />

medicines<br />

proposal<br />

DISCOVERY DEVELOPMENT<br />

Target<br />

assessment<br />

Lead<br />

identi-<br />

fication<br />

Figure 1 <strong>Solubility</strong> <strong>and</strong> dissolution determinations in the drug discovery <strong>and</strong> development<br />

stages.<br />

2.1.1 Kinetic solubility<br />

Lead<br />

optimi-<br />

zation<br />

Kinetic Equilibrium<br />

solubility solubility<br />

(HTS) (shake-flask)<br />

Entry into<br />

human<br />

enable<br />

The measurement <strong>of</strong> the kinetic (non-equilibrium) solubility usually starts from the<br />

dissolved compound <strong>and</strong> represents the maximum solubility <strong>of</strong> the fastest precipitating<br />

material, which is typically not determined. The precipitating material can be any <strong>of</strong> the<br />

existing different solid forms <strong>of</strong> the compound, for example, polymorphic changes can<br />

happen even within minutes (Aaltonen et al., 2006). Also the formation <strong>of</strong> different salts<br />

may be possible depending on the presence <strong>of</strong> ionic species in the medium. In practice, the<br />

small volume (μl) <strong>of</strong> the stock solution (most <strong>of</strong>ten DMSO), in which the compound is<br />

dissolved, is added to the aqueous solution until the solubility limit is reached.<br />

Experiments can be performed by adding equal volumes <strong>of</strong> DMSO stock solution to the<br />

aqueous medium spaced a minute apart (Lipinski et al., 1997). In an alternative<br />

experiment, a dilution series <strong>of</strong> DMSO stock solution is prepared into separate vials<br />

(Kibbey et al., 2001). The kinetic solubility is defined to be the concentration at which the<br />

precipitation starts.<br />

2.1.1.1 Dimethyl sulfoxide (DMSO) as a co-solvent<br />

The amount <strong>of</strong> DMSO in the aqueous solution is increased in the kinetic solubility<br />

experiments by every addition <strong>of</strong> the DMSO stock solution. The final DMSO<br />

concentration varies <strong>and</strong> concentrations 0.5% (Bard et al., 2008), 1% (Chen et al., 2002b;<br />

Taub et al., 2002; Bard et al., 2008), 2% (Kibbey et al., 2001; Dehring et al., 2004), 5%<br />

(Pan et al., 2001; Bard et al., 2008) <strong>and</strong> even up to 20% (Kramer et al., 2009) have been<br />

reported. DMSO might induce a higher water-solubility than the compound would have<br />

without the DMSO due to the supersaturation or co-solvent effects (Chen et al., 2002b;<br />

L<strong>of</strong>tsson <strong>and</strong> Hreinsdottir, 2006; Bard et al., 2008). The co-solvents increase the solubility<br />

Phase<br />

I<br />

13<br />

Phase<br />

II<br />

Phase<br />

III<br />

<strong>Dissolution</strong><br />

(regulatory methods, intrinsic dissolution rate)<br />

Registration<br />

On<br />

market


y being miscible with water <strong>and</strong> acting as better solvents for the solute in question.<br />

DMSO is an efficient solvent for poorly water-soluble compounds because <strong>of</strong> its<br />

amphipathic structure with highly polar domain <strong>and</strong> two apolar groups enabling it to mix<br />

both hydrophilic (with water) <strong>and</strong> hydrophobic (organic compounds) environments<br />

(Santos et al., 2003). Supersaturated solutions are solutions, which have higher solute<br />

concentration than in the equilibrium state. The equilibrium is established between the<br />

solute in the solution <strong>and</strong> the excess <strong>of</strong> (undissolved) compound. DMSO enhances the<br />

solubility in a compound specific way (Chen et al., 2002b; Taub et al., 2002). Taub et al.<br />

(2002) investigated the possibilities <strong>of</strong> replacing DMSO by another co-solvent, for<br />

example ethanol or HPCD (hydroxypropyl-β-cyclodextrin), but the effect <strong>of</strong> enhanced<br />

solubility remained. According to Alsenz <strong>and</strong> Kansy (2007), the DMSO concentration<br />

should be kept as low as possible (1%) to reduce the potential co-solvent effect <strong>and</strong> to<br />

achieve better correlation with the equilibrium solubility methods. The final co-solvent<br />

<strong>and</strong> the amount <strong>of</strong> it should always be reported, although the solubility results obtained<br />

from different kinetic methods are typically not comparable to each other <strong>and</strong> the kinetic<br />

solubility values are not even expected to be reproducible between different kinetic<br />

methods (Stuart <strong>and</strong> Box, 2005). Comparison <strong>of</strong> the results is complicated because <strong>of</strong> the<br />

different DMSO concentrations at the end-point (Chen et al., 2002b). Di <strong>and</strong> Kerns (2006)<br />

have reviewed the strategies for optimizing bioassays with DMSO, including<br />

improvements in storage <strong>and</strong> h<strong>and</strong>ling. Serial dilutions in DMSO are recommended, <strong>and</strong><br />

these dilutions should be added directly to the assay media. Assays should also run at the<br />

lowest useful concentrations.<br />

DMSO, the current industry st<strong>and</strong>ard solvent for screening purposes, has some very<br />

problematic features, like it suffers from the rapid uptake <strong>of</strong> water; if the water content<br />

gets too high in the DMSO solutions, some compounds may fall out <strong>of</strong> solution <strong>and</strong> give a<br />

false negative result (Homon <strong>and</strong> Nelson, 2006). DMSO is hygroscopic <strong>and</strong> can absorb<br />

water up to 10% <strong>of</strong> its original volume (Di <strong>and</strong> Kerns, 2006), which also causes reduced<br />

solubility with the crystallization in the freeze-thaw cycles. DMSO has also several<br />

systemic adverse effects, e.g. nausea, vomiting, diarrhea, anaphylactic reactions,<br />

bradycardia <strong>and</strong> heart block are reported (O’Donnell et al., 1981; Davis et al., 1990;<br />

Rapoport et al., 1991; Stroncek et al., 1991; Styler et al., 1992). Besides, even 20% <strong>of</strong> the<br />

compounds in the commercial libraries are poorly soluble in DMSO according to Balakin<br />

et al., 2004.<br />

2.1.1.2 Kinetic solubility assays<br />

In general, kinetic solubility values are measured in the drug discovery process to screen<br />

the lead compounds. A high-throughput turbidimetric method for kinetic solubility assay,<br />

in which the detection was based on nephelometry (Davis <strong>and</strong> Parke, 1942), was<br />

introduced by Lipinski et al. (1997). Detection is based on the intensity <strong>of</strong> the scattered<br />

light as a function <strong>of</strong> concentration. In general, kinetic solubility measurements are fast<br />

<strong>and</strong> suitable for the HTS processes h<strong>and</strong>ling more than 100.000 samples per day <strong>and</strong> cover<br />

14


the required concentration range in assays used in early discovery (Homon <strong>and</strong> Nelson,<br />

2006). More details about the methods are presented in the Table 2.<br />

Table 2. Kinetic solubility in the drug discovery stage. Precipitation quantified by<br />

nephelometry or UV-absorbance.<br />

Sensitiveness<br />

Sample volume<br />

DMSO conc.<br />

Advantages<br />

Nephelometry a-d Diode array UV a UV plate reader d-g<br />

20-54 μM<br />

200-300 μl<br />

0.67-5%<br />

Colored compounds do<br />

not interfere,<br />

Shaking or<br />

ultrasonication is not<br />

required.<br />

5-65 μg/ml<br />

2.5 ml<br />

0.67%<br />

Accurate determination based<br />

on a calibration curve.<br />

0.5-1 μM<br />

200 μl<br />

0.5–5%<br />

Able to detect impurities<br />

or low UV absorption.<br />

Disadvantages Small undissolved Colored compounds miscalled Compounds should<br />

particles cause as insoluble.<br />

have sufficient<br />

misinterpretations;<br />

absorbance in the<br />

High/low<br />

region > 230 nm,<br />

concentrations<br />

Absorption <strong>of</strong> residues<br />

can be problematic.<br />

or other materials at the<br />

same wavelength as<br />

drug compounds.<br />

a b c d<br />

Lipinski et al., 1997; Bevan <strong>and</strong> Lloyd, 2000; Dehring et al., 2004; Hoelke et al., 2009;<br />

e f g<br />

Pan et al., 2001; Chen et al., 2002b; Bard et al., 2008<br />

2.1.2 Equilibrium solubility<br />

The equilibrium (thermodynamic) solubility test is performed by dispensing a surplus <strong>of</strong><br />

solid compound into a liquid medium. After reaching the equilibrium state between the<br />

excess <strong>of</strong> undissolved compound <strong>and</strong> saturated solution, the saturation solubility can be<br />

defined. To confirm that equilibrium state has been reached, the solubility has to be<br />

constant for several consecutive time-points, which takes several hours or days, even<br />

months to reach. Usually equilibrium solubility is studied in the later stages <strong>of</strong> drug<br />

15


discovery process <strong>and</strong> it is regarded as the golden st<strong>and</strong>ard for development needs (Alsenz<br />

<strong>and</strong> Kansy, 2007).<br />

A classical equilibrium solubility assay is the shake-flask method. In this method an<br />

excess <strong>of</strong> solid drug is dispensed into aqueous solution, after which the container is shaken<br />

at certain speed <strong>and</strong> temperature. Samples are collected regularly, usually during 24-48<br />

hours, to make certain that the equilibrium plateau has been reached. The shake-flask<br />

method has not been <strong>of</strong>ficially defined in pharmacopoeias, so several protocols exist<br />

(Avdeef et al., 2000; Blasko et al., 2001; Kerns, 2001; Bergström et al., 2002; Lipinski et<br />

al., 2003; Chen <strong>and</strong> Venkatesh, 2004; Glomme et al., 2005; Avdeef, 2007; Zhou et al.,<br />

2007) (Table 3).<br />

Table 3. Different protocols to perform shake-flask solubility measurements (N.A. not<br />

available in the article).<br />

Blasko et al., 2001<br />

Kerns, 2001<br />

Bergström et al., 2002<br />

Chen <strong>and</strong> Venkatesh, 2004<br />

Glomme et al., 2005<br />

Avdeef, 2007<br />

Zhou et al., 2007<br />

Medium<br />

volume (ml)<br />

5.0<br />

0.45<br />

0.05-1.0<br />

1.0<br />

2.0<br />

2.0-20.0<br />

1.0-5.0<br />

Shaking<br />

(rpm)<br />

N.A.<br />

N.A.<br />

300<br />

N.A.<br />

450<br />

N.A.<br />

600<br />

16<br />

Residue removal Quantification<br />

filtration<br />

filtration<br />

centrifugation<br />

filtration<br />

filtration<br />

filtration or<br />

centrifugation<br />

filtration or<br />

centrifugation<br />

2.2 In vitro dissolution methods for solid dosage forms<br />

<strong>Dissolution</strong> (dissolution rate) can be expressed by the Noyes-Whitney equation:<br />

HPLC<br />

HPLC<br />

HPLC-UV/fluorescence<br />

HPLC-UV<br />

HPLC-UV<br />

HPLC-UV<br />

HPLC-UV<br />

(LC-MS)<br />

dm DA<br />

= (Cs – Cb). (2)<br />

dt h<br />

The rate <strong>of</strong> mass transfer <strong>of</strong> solute molecules or ions through a static diffusion layer,<br />

dm/dt, is directly proportional to the area available for molecular or ionic migration, A, the<br />

concentration difference, Cs-Cb, <strong>and</strong> D, the diffusion coefficient <strong>of</strong> the solute (usually<br />

m 2 /s) <strong>and</strong> is inversely proportional to the thickness <strong>of</strong> the boundary layer, h. Cs is the


saturation solubility <strong>and</strong> Cb is the concentration <strong>of</strong> drug at a particular time in the bulk<br />

dissolution medium (Noyes <strong>and</strong> Whitney, 1897). If the Cb does not exceed 10% <strong>of</strong> the Cs,<br />

the dissolution is said to occur under sink conditions. In the table 4 are the factors<br />

affecting the dissolution rate.<br />

Table 4. Factors affecting the dissolution rate.<br />

Term Factor Examples<br />

A Particle size <strong>and</strong> shape<br />

- Aggregation<br />

Dokoumetzidis <strong>and</strong> Macheras,<br />

2006<br />

- Crystal shape affected by crystallization process<br />

Fini et al., 1995<br />

Cs<br />

Cb<br />

h<br />

D<br />

Solid-state form<br />

- Change in the solid form properties<br />

Complex formation<br />

Salts<br />

- common ion-effect<br />

Solubilizing agents<br />

Temperature<br />

<strong>Dissolution</strong> medium<br />

- e.g. pH (ionizable compounds), co-solvents<br />

Volume <strong>of</strong> medium<br />

Processes that remove dissolved solute from the medium<br />

- e.g. adsorption<br />

Thickness <strong>of</strong> boundary layer<br />

- affected by agitation<br />

Diffusion coefficient <strong>of</strong> solute in the dissolution medium<br />

- affected by the viscosity <strong>of</strong> medium<br />

17<br />

Pudipeddi <strong>and</strong> Serajuddin, 2005<br />

Brewster <strong>and</strong> L<strong>of</strong>tsson, 2007<br />

Serajuddin, 2007<br />

Bakatselou et al., 1991<br />

Hill <strong>and</strong> Petrucci, 1999<br />

Taub et al., 2002<br />

Dressman et al., 1998<br />

Lindenberg et al., 2005<br />

Alsenz et al., 2007<br />

Dressman et al., 1998


The importance <strong>of</strong> dissolution to the drug absorption <strong>and</strong> bioavailability is emphasized in<br />

the Biopharmaceutics Classification System (see also later chapters) by Dn, dissolution<br />

number (Amidon et al., 1995):<br />

Dn =<br />

t<br />

t<br />

res<br />

Diss<br />

where tres is the mean intestinal residence time (ca. 180 min, which differs from MAD<br />

model time) <strong>and</strong> tDiss is the time required for a drug particle to dissolve. From the above<br />

equation it is obvious that the higher the Dn, the better the drug absorption, <strong>and</strong> the<br />

maximal drug absorption is expected when Dn>1. In the case <strong>of</strong> Dn


Table 5. The <strong>of</strong>ficial dissolution testing conditions.<br />

Medium<br />

Volume (ml)<br />

pH <strong>of</strong> buffers<br />

Temperature ( o C)<br />

Enzymes<br />

Agitation (rpm)<br />

Basket method<br />

Paddle method<br />

a<br />

FDA, 1997a<br />

Ph. Eur. USP a JP<br />

900<br />

1.0-7.5<br />

37±0.5<br />

case-by-case<br />

100<br />

50<br />

500, 900, 1000<br />

1.2-6.8<br />

37±0.5<br />

case-by-case<br />

50-100<br />

50-75<br />

2.2.2 Initial steps <strong>of</strong> the dissolution process<br />

19<br />

900<br />

1.2-6.8<br />

37±0.5<br />

case-by-case<br />

The main difference between the stirring methods described above <strong>and</strong> intrinsic<br />

dissolution method (IDR) (Ph. Eur., 2009; USP, 2009) or channel flow cell technique<br />

(Compton et al., 1988; Brown et al., 1993; Compton et al., 1993) is that in the latter<br />

methods only one tablet side is exposed to the dissolution medium (Figure 2).<br />

100<br />

50


Figure 2 Schematic illustrations <strong>and</strong> liquid flows <strong>of</strong> a) Intrinsic dissolution apparatus <strong>and</strong> b)<br />

Channel flow cell method.<br />

Intrinsic dissolution rate (IDR)<br />

The IDR method is based on the idea that the dissolution environment (surface area,<br />

temperature, pH, stirring speed <strong>and</strong> ionic strength) <strong>of</strong> pure drug substances is kept constant<br />

during the whole measurement (Nicklasson et al., 1981; Nicklasson et al., 1985). The fluid<br />

flow has been studied (Mauger, 1996; Mauger et al., 2003) <strong>and</strong> the IDR method has also<br />

been successfully miniaturized requiring only 5 mg <strong>of</strong> pure drug (Berger et al., 2007;<br />

Persson et al., 2008). As an advantage the compression pressure, dissolution medium<br />

volume or die position have no significant effect on IDR (Yu et al., 2004). The IDR <strong>of</strong> the<br />

pure drug is an important physicochemical property, which is useful, e.g., when predicting<br />

20


problems related to the absorption <strong>of</strong> the drugs. The evaluation <strong>of</strong> the IDR <strong>of</strong> pure drug<br />

substances demonstrates chemical purity <strong>and</strong> equivalence.<br />

The intrinsic rate (j) is expressed in terms <strong>of</strong> dissolved mass <strong>of</strong> compound per time per<br />

exposed area (mg min -1 cm -2 ):<br />

V dc<br />

j = , (4)<br />

A dt<br />

where V is volume, A is area <strong>of</strong> the drug disk, c is concentration <strong>and</strong> t is time. IDR can<br />

also be calculated by Levich equation (Yu et al., 2004):<br />

2 / 3<br />

D <br />

j = 0.62 ( 1/<br />

6<br />

v<br />

1/<br />

2<br />

) Cs, (5)<br />

where Cs is saturated solubility or concentration, D is the diffusion coefficient, ω is the<br />

angular velocity <strong>of</strong> the rotating disk <strong>and</strong> v is the kinematic viscosity <strong>of</strong> the dissolution<br />

medium. It is assumed in this model that equilibrium is predominant at the interface <strong>of</strong><br />

solid <strong>and</strong> liquid, <strong>and</strong> the liquid at the interface is the same as the liquid in bulk dissolution.<br />

Since the intrinsic dissolution is primarily a rate phenomenon, it is expected to correlate<br />

more closely with in vivo drug dissolution dynamics than solubility. Still, the correlation<br />

between the solubility <strong>and</strong> intrinsic dissolution rate <strong>of</strong> a drug compound has been found<br />

both in water or aqueous buffer solutions (Nicklasson et al., 1981; Gu et al., 1987;<br />

Nicklasson et al., 1988) <strong>and</strong> water-ethanol binary solvent systems (Nicklasson <strong>and</strong> Brodin,<br />

1984). Also the correlation between the IDR <strong>and</strong> BCS solubility classification has been<br />

found by Yu et al. (2002).<br />

Channel flow cell (CFC) method<br />

The hydrodynamics (laminar flow) in the channel flow cell (CFC) is well defined<br />

(Compton et al., 1988; Brown et al., 1993; Compton et al., 1993), which enables<br />

acquisition <strong>of</strong> quantitative information <strong>of</strong> the dissolution kinetics <strong>and</strong> mechanism, even at<br />

early stages <strong>of</strong> the process, <strong>and</strong> modelling. This information is <strong>of</strong>ten lost in the traditional<br />

dissolution methods. In addition, accurate modelling <strong>of</strong> the mass transfer in the CFC<br />

enables comparisons <strong>of</strong> different dissolution mechanisms <strong>and</strong> experimental results. The<br />

flow rate in the channel presents an additional experimental parameter that can be used to<br />

tune the rate <strong>of</strong> mass transfer in the flowing aqueous phase.<br />

When the dissolution experiments involve pure compounds, the model consists <strong>of</strong> a<br />

convective mass transfer in the medium coupled to a dissolution reaction at the tablet<br />

surface with an equilibrium constant, K (Singh <strong>and</strong> Dutt, 1985). Theoretical expression to<br />

the concentration, c b , as a function <strong>of</strong> time:<br />

21


c b mA<br />

(t) = K (1 – exp (- t)) (6)<br />

V<br />

where m is the mass transfer coefficient, A is the exposed area <strong>of</strong> the tablet <strong>and</strong> V is the<br />

volume <strong>of</strong> the dissolution medium <strong>and</strong> t is time.<br />

In the CFC the tablet is located in the cavity so that the disintegration <strong>of</strong> the tablet is<br />

hindered when, e.g., the properties <strong>of</strong> disintegration excipients do not affect the<br />

dissolution as in the traditional methods. In the case <strong>of</strong> tablet formulations, it is not<br />

expected that the mass transfer coefficient would be equal for all tablets <strong>and</strong> only a<br />

function <strong>of</strong> the flow rate. With tablet formulations including excipients, diffusion <strong>and</strong><br />

dissolution inside the tablet <strong>and</strong> the sink conditions were included in the model (Higuchi,<br />

1961):<br />

2 b<br />

b<br />

V c ( t)<br />

V V 2K<br />

K − c ( t)<br />

t + + ( + ) ln ( ) = 0 (7)<br />

2 o<br />

2 o<br />

D A c mA D A c K<br />

T<br />

T<br />

where DT is the diffusion coefficient <strong>of</strong> the drug inside the tablet <strong>and</strong> c 0 the concentration<br />

<strong>of</strong> the drug in the dry tablet.<br />

The combination <strong>of</strong> dissolution testing <strong>and</strong> solid-state analysis has been proposed to be<br />

called physico-relevant dissolution testing (Aaltonen <strong>and</strong> Rades, 2009). Recently, the<br />

channel flow method has been utilized, e.g., to study the solvent-mediated solid-state<br />

transformation during dissolution (Aaltonen et al., 2006; Lehto et al., 2008), the solid-state<br />

behavior during dissolution <strong>of</strong> amorphous drugs (Savolainen et al., 2009) <strong>and</strong> the effect <strong>of</strong><br />

texture on the intrinsic dissolution behavior (Tenho et al., 2007).<br />

2.2.3 In vivo mimicking <strong>of</strong> dissolution methods<br />

Alternative methods have been developed to complete the lacks <strong>of</strong> the regulatory methods.<br />

In general, the <strong>of</strong>ficial basket <strong>and</strong> paddle dissolution techniques suffer from high<br />

variability in results (Achanta et al., 1995; Qureshi <strong>and</strong> McGilvery, 1999). Further, the<br />

basket method suffers, e.g., from the hydrodynamic ‘dead zone’ under the basket,<br />

blockings in the mesh by drug or excipient <strong>and</strong> hindered visual examination <strong>of</strong> the<br />

disintegration process (Banaker, 1992). The paddle method, for its part, suffers above all<br />

from the coning effect (Beckett et al., 1996).<br />

Several un<strong>of</strong>ficial in vitro dissolution methods exist (Pernarowski et al., 1968;<br />

Shepherd et al., 1972; Shah et al., 1973; Cakiryildis et al., 1975), some <strong>of</strong> them modified<br />

from the regulatory methods (Qureshi <strong>and</strong> Shabnam, 2003; Röst <strong>and</strong> Quist, 2003). The<br />

new methods, e.g. the dissolution stress-test device, are more complicated (Garbacz et al.,<br />

2008; Garbacz et al., 2009). This method mimics the discontinuous movements <strong>of</strong> the<br />

dosage form <strong>and</strong> breaks <strong>of</strong> it from the intestinal fluid in the GI tract. Even more<br />

complicated release <strong>and</strong> solubility testing methods are the physiologically based TIM-1<br />

22


(stomach <strong>and</strong> small intestine) <strong>and</strong> TIM-2 systems (large-intestine) (Havenaar <strong>and</strong><br />

Menikus, 1994). For example the TIM-1 has four compartments simulating stomach,<br />

duodenum, jejunum <strong>and</strong> ileum. The TIM devices mimic, e.g., the peristaltic movements,<br />

GI transition times, pH <strong>of</strong> GI tract <strong>and</strong> secretion <strong>of</strong> bile, gastric acids <strong>and</strong> enzymes. Results<br />

with the TIM-1 system compared to the in vivo data have been similar (Blanquet et al.,<br />

2004; Souliman et al., 2006). Still, the new dissolution devices need a large volume <strong>of</strong><br />

medium, which enables their use only in the later stages <strong>of</strong> drug development process.<br />

2.3 In vivo dissolution<br />

The most common <strong>and</strong> patient convenient way to deliver drugs is the oral route. After oral<br />

administration, the physiology <strong>of</strong> the GI tract affects the dissolution <strong>and</strong> absorption <strong>of</strong> the<br />

drug. The GI tract is divided into three major anatomical regions (the stomach, the small<br />

intestine <strong>and</strong> the colon), from which the small intestine is the most important site for drug<br />

absorption. The length <strong>of</strong> the small intestine is about 3.3 meters (Ho et al., 1983) <strong>and</strong> the<br />

area available for absorption is about 200 m 2 (Ashford, 2007). The large area <strong>of</strong> small<br />

intestine is possible because <strong>of</strong> the folds <strong>of</strong> Keckring, the villi <strong>and</strong> the microvilli. The<br />

absorption sites <strong>of</strong> drugs in the small intestine differ from each other. Highly permeable<br />

drugs absorb mainly at the tips <strong>of</strong> the villi <strong>and</strong> the low permeability drugs may diffuse<br />

down the length <strong>of</strong> the villi to be absorbed over a larger surface area (Artursson et al.,<br />

2001). The main anatomical barrier for the drug absorption is the enterocytes, which<br />

apical side constitutes <strong>of</strong> densely packed microvilli <strong>and</strong> the lateral membranes connected<br />

through tight junctions. The absorption routes from the intestinal epithelium are based on<br />

paracellular passive diffusion, transcellular pathway (passive diffusion, carrier-mediated,<br />

receptor-mediated) or carrier-mediated efflux pathways (Back <strong>and</strong> Rogers, 1987; Tsuji<br />

<strong>and</strong> Tamai, 1996; Stenberg et al., 2000; Artursson et al., 2001; Martinez <strong>and</strong> Amidon,<br />

2002).<br />

Before possible absorption the drug needs to be released <strong>and</strong> dissolved from the solid<br />

dosage form to the GI fluids. The pH, composition, volume <strong>and</strong> hydrodynamics <strong>of</strong> the<br />

contents in the lumen are the main factors in the dissolution <strong>of</strong> drugs in the GI tract<br />

(Dressman et al., 1998).<br />

2.3.1 Effect <strong>of</strong> pH<br />

The pH in the GI tract varies depending on the anatomical region <strong>and</strong> the nutrition state.<br />

Stomach has an acidic pH in the fasted state, median values are found between 1.7 <strong>and</strong> 2.4<br />

with high interindividual variations. After meal the gastric median pH increases up to 6.4-<br />

6.7 depending on the type <strong>and</strong> volume <strong>of</strong> the meal. The pH <strong>of</strong> stomach is returned acidic<br />

2-4 hours after the meal (Dressman et al., 1990; Kalantzi et al., 2006a). Intestinal pH is<br />

higher than the gastric pH. In the fasting state the median pH is about 6.1-7.2 (Lindahl et<br />

al., 1997; Kalantzi et al., 2006a; Moreno et al., 2006; Brouwers et al., 2006; Clarysse et<br />

al., 2009b) <strong>and</strong> in the fed state 6.1-6.6 (Dressman et al., 1990; Kalantzi et al., 2006a;<br />

23


Clarysse et al., 2009b), with less interindividual variation compared to the gastric pH<br />

(Table 6). The pH has been reported to be similar in the jejunum (median 7.0) <strong>and</strong><br />

duodenum (median 6.5) in the fasted state (Moreno et al., 2006). The pH in the colon<br />

depends on the region. In the caecum the pH is around 5.6-6.6, increasing up to 7-7.5 in<br />

the distal parts <strong>of</strong> the colon (Fallingborg et al., 1989). Especially, for ionizable compounds<br />

the pH <strong>of</strong> the GI tract is important; Ratio <strong>of</strong> unionized <strong>and</strong> ionized forms depends on pH.<br />

Usually the ionized form has higher solubility than unionized form. For example weak<br />

acids has slow solubility <strong>and</strong> dissolution in stomach, whereas they can achieve high<br />

solubility <strong>and</strong> dissolution in intestine. On the contrary, for poorly soluble compounds the<br />

pH <strong>of</strong> GI is not the rate-limiting step for dissolution, but in the GI tract is not enough fluid<br />

available or the entire dose <strong>of</strong> the drug dissolves too slowly (Dressman et al., 1998).<br />

2.3.2 Role <strong>of</strong> bile acids <strong>and</strong> phospholipids<br />

Bile is secreted to the small intestine. The secretion <strong>of</strong> bile is at maximum about 30<br />

minutes after ingestion <strong>of</strong> a meal <strong>and</strong> the gallbladder is emptied up to 73% (Mathus-<br />

Vliegen et al., 2004). The major components <strong>of</strong> bile are the bile acids, phospholipids,<br />

cholesterol <strong>and</strong> free fatty acids. The bile salt concentration compared to the phospholipid<br />

concentration has been reported to be 2-3:1 in the fed state <strong>and</strong> 9:1 in the fasted state<br />

(Persson et al., 2005; Kalantzi et al., 2006b). Bile has an important role in improving<br />

solubility <strong>and</strong> dissolution <strong>of</strong> poorly water-soluble drugs. Especially, bioavailability <strong>of</strong><br />

poorly soluble drugs increases manifold when administered in fed state or with lipid-based<br />

formulations (Persson et al., 2005; Sunesen et al., 2005; Kalantzi et al., 2006b). The effect<br />

could result from decreased first pass metabolism, lymphatic uptake (Porter et al., 1996),<br />

inhibition <strong>of</strong> efflux transporters (Wu et al., 2005) or increased solubility (Persson et al.,<br />

2005; Kalantzi et al., 2006b) <strong>and</strong> dissolution. Bile acids enhance solubility due to the<br />

wetting (Bakatselou et al., 1991) <strong>and</strong>/or solubilization effect (Bakatselou et al., 1991;<br />

Charman et al., 1993) which has been demonstrated e.g. with steroidal compounds<br />

(Mithani et al., 1995).<br />

For the prediction <strong>of</strong> in vivo dissolution <strong>and</strong> solubility, simulated intestinal fluids, like<br />

FaSSIF <strong>and</strong> FeSSIF, have been developed (Galia et al., 1998). Jantratid et al. (2008)<br />

proposed further a set <strong>of</strong> media to represent various regions <strong>of</strong> the gastrointestinal tract:<br />

fasted state simulated gastric fluid (FaSSGF) (Vertzoni et al., 2005), fed state simulated<br />

gastric fluid (FeSSGF), fasted <strong>and</strong> fed state simulated upper small intestine fluids<br />

(FeSSIF-V2, FaSSIF-V2) <strong>and</strong> various media for conditions that represent digestion.<br />

Simulated media contain e.g. lecithin <strong>and</strong> sodium taurocholate to represent the<br />

phospholipids <strong>and</strong> the bile salts.<br />

24


Table 6. The compositions <strong>of</strong> fasted <strong>and</strong> fed state HIF.<br />

Bile Acids/salts (mM)<br />

Taurocholic acid (%)<br />

Glycocholate (%)<br />

Glycochenodeoxycholate (%)<br />

Glycodeoxycholate (%)<br />

Taurochenodeoxycholate (%)<br />

Taurodeoxycholate (%)<br />

Phospholipids (mM)<br />

Lyso-phosphatidylcholine (%)<br />

Phosphatidylcholine (%)<br />

Parameters<br />

Osmolality (mOsm/kg)<br />

pH<br />

Buffer capacity (mmol/l pH -1 )<br />

Fasted HIF Fed HIF<br />

2.0-3.5 a-e<br />

22.0 a , 44.95±18.17 d<br />

25.0 a , 17.21±6.37 d<br />

18.0 a , 13.77±9.59 d<br />

1.0 a , 10.40±7.32 d<br />

20.0 a , 7.89±2.16 d<br />

1.0 a , 4.71±2.47 d<br />

0.2–0.6 a,i<br />

99.0 a<br />

1.0 a<br />

200–271 b-d<br />

6.6-7.5 a-f<br />

2.4-2.8 a-e<br />

25<br />

8.0-15.0 a,g,i<br />

18.0 a<br />

25.0 a<br />

20.0 a<br />

13.0 a<br />

15.0 a<br />

9.0 a<br />

0.07-4.8 a,g,i<br />

87.0 a<br />

13.0 a<br />

287 k<br />

5.1-6.1 f,j,k<br />

13.2-14.6 a<br />

a Persson et al., 2005; b Clarysse et al., 2009b; c Lindahl et al., 1997; d Moreno et al., 2006;<br />

e Brouwers et al., 2006; f Kalantzi et al., 2009a; g Mansbach et al., 1975; h Sheng et al., 2009;<br />

i Arm<strong>and</strong> et al., 1996; j Clarysse et al., 2009a; k Dressman et al., 1998<br />

2.3.3 Hydrodynamics in gastrointestinal (GI) tract<br />

In the fasted state GI tract has only little or no motoric activity. Still, the ‘’house-keeping’’<br />

waves empty the stomach regularly about every two hours from the stomach to the small<br />

intestine. In the fed state mixing is more active <strong>and</strong> dissolution is expected to be more<br />

efficient. The gastric emptying time is higher after meal ingestion or after large liquid<br />

volume compared to the fasted state <strong>and</strong> the lipid contents <strong>of</strong> the meal increases the transit<br />

time further (Sunesen et al., 2005). For small intestine the transit time varies from three to<br />

four hours <strong>and</strong> stays the same whether the drug is taken with meal or not (Sunesen et al.,<br />

2005). Instead, the flow velocity varies in different parts <strong>of</strong> the small intestine e.g. being<br />

higher in the ileum as in the jejunum (Ho et al., 1983). The volume <strong>of</strong> stomach <strong>and</strong><br />

intestine in fasted state is 20-30 ml <strong>and</strong> 120-350 ml (Dressman et al., 1998). Under fed<br />

conditions the volume might be as high as 1500 ml in the upper intestine.


2.4 In vitro in vivo (IVIV) correlation<br />

IVIV correlation has been defined as ‘a predictive mathematical model describing the<br />

relationship between in vitro property <strong>of</strong> a dosage form <strong>and</strong> relevant in vivo response, for<br />

example plasma drug concentration or amount <strong>of</strong> drug absorbed’ (FDA, 1997b). IVIV<br />

correlation serves as a surrogate for in vivo data <strong>and</strong> is also used to validate the dissolution<br />

methods <strong>and</strong> in quality control. Three main levels <strong>of</strong> the IVIV correlation have been<br />

determined: level A, level B <strong>and</strong> level C (FDA, 1997b). The level A correlation represents<br />

the linear relationship between in vitro dissolution rate <strong>and</strong> in vivo input curves <strong>of</strong> the drug<br />

from the dosage form. Level A correlation is the most informative <strong>and</strong> recommended, if<br />

possible. Table 7 compares the release testing in vitro <strong>and</strong> in vivo. Medium, volume,<br />

hydrodynamics, temperature, location <strong>of</strong> the formulation <strong>and</strong> amount <strong>of</strong> the drug are<br />

considered.<br />

Table 7. Differences between the in vitro <strong>and</strong> in vivo release testing <strong>of</strong> immediate release<br />

drug products.<br />

Parameter IN VIVO IN VITRO<br />

Medium GI fluids<br />

Water,<br />

Buffer solutions,<br />

Simulated media<br />

Volume 250 ml (fasted state)<br />

Depending on the device<br />

(interindividual variation) (from μl:s to1000 ml)<br />

Hydrodynamics GI motility Mechanical agitation,<br />

Depending on the device<br />

Temperature Body temperature Depending on the test<br />

circumstances<br />

Location <strong>of</strong> the formulation Variation with time Depending on the device<br />

Amount <strong>of</strong> the drug Decreases as it is absorbed Constant in closed systems,<br />

(e.g. open system)<br />

Decreases in open systems<br />

2.4.1 Biopharmaceutics Classification System (BCS)<br />

The meaning <strong>of</strong> BCS is to correlate in vitro drug product dissolution <strong>and</strong> in vivo<br />

bioavailability (Amidon et al., 1995). In practice, BCS divides drugs into four classes<br />

according to their solubility <strong>and</strong> permeability properties. The drugs are classified as highly<br />

soluble if the highest orally administered single dose can be completely dissolved in 250<br />

ml <strong>of</strong> water, independently <strong>of</strong> the pH value at the range <strong>of</strong> 1.0–7.5 at 37 o C. Further, the<br />

drug is considered to be highly permeable when 90% or more <strong>of</strong> the orally administered<br />

dose is absorbed in the small intestine. The ideal drug would be class I drug having high<br />

26


solubility <strong>and</strong> permeability. The dissolution test for these drugs is only to verify that the<br />

drug is rapidly released from the dosage form. For the BCS class I drugs the IVIV<br />

correlation exist if dissolution rate is slower than gastric emptying rate. For BCS class II<br />

drugs the drug absorption is expected to be dissolution limited <strong>and</strong> IVIV correlation<br />

should exist if the dose is not very high. For BCS class III drugs limited or no IVIV<br />

correlation is expected because permeability is the rate limiting step. These drugs dissolve<br />

rapidly, while the slow absorption restricts bioavailability during the passage in the<br />

absorption site in the GI tract. The BCS class IV drugs are <strong>of</strong>ten not very suitable for oral<br />

administration at all, so limited or no IVIV correlation is expected.<br />

Fleischer et al. (1999) noticed that food (high-fat meals) effects to the extent <strong>of</strong><br />

bioavailability could generally be predicted based on the BCS. For example, high-fat<br />

meals have no significant effect on the extent <strong>of</strong> availability for class I drugs, although<br />

food may delay the stomach emptying time. Class II drugs would benefit from high-fat<br />

meals, e.g., because <strong>of</strong> additional solubilization <strong>and</strong> inhibition <strong>of</strong> efflux transporters in the<br />

lumen. High-fat meal effects would be decreased if the class II drugs were formulated<br />

towards having better solubility.<br />

2.4.2 Biopharmaceutics <strong>Drug</strong> Disposition Classification System (BDDCS)<br />

In 2005 Wu <strong>and</strong> Benet represented a modified version <strong>of</strong> BCS called Biopharmaceutics<br />

<strong>Drug</strong> Disposition Classification System (BDDCS) (Table 8), which is used in the<br />

prediction <strong>of</strong> overall drug disposition, including the rates <strong>of</strong> elimination routes <strong>and</strong> the<br />

effects <strong>of</strong> efflux <strong>and</strong> absorptive transporters on oral drug absorption. Accordingly, the<br />

extent <strong>of</strong> drug metabolism would be a better predictor than the 90% absorption<br />

characteristic. It was noticed that BCS class I <strong>and</strong> II compounds are eliminated primarily<br />

via metabolism, while class III <strong>and</strong> IV compounds end up unchanged into the urine <strong>and</strong><br />

bile; if the major route <strong>of</strong> elimination for a drug is metabolism, then the drug exhibits the<br />

properties <strong>of</strong> high permeability compounds, <strong>and</strong> if the major route <strong>of</strong> elimination is renal<br />

<strong>and</strong> biliary excretion <strong>of</strong> unchanged drug, the drug is classified as low permeability<br />

compound. Using the extent <strong>of</strong> metabolism instead <strong>of</strong> permeability (BCS) changed the<br />

classification <strong>of</strong> 10 drugs, changed 11 multiple classified drugs to one class, <strong>and</strong> added 38<br />

drugs to the totally 168 drugs’ BDDCS list. The BDDCS predicted 93% <strong>of</strong> the drugs’<br />

permeabilities correctly related to major route <strong>of</strong> elimination (Benet et al., 2008).<br />

27


Table 8. The Biopharmaceutics <strong>Drug</strong> Disposition Classification System by Wu <strong>and</strong> Benet<br />

(2005). The major route <strong>of</strong> elimination (metabolized vs. unchanged) serves as the<br />

permeability criteria.<br />

Class I<br />

High <strong>Solubility</strong><br />

Extensive Metabolism<br />

Class III<br />

High <strong>Solubility</strong><br />

Poor Metabolism<br />

28<br />

Class II<br />

Poor <strong>Solubility</strong><br />

Extensive Metabolism<br />

Class IV<br />

Poor <strong>Solubility</strong><br />

Poor Metabolism<br />

The BDDCS can also predict the effects <strong>of</strong> transporters. According to the Wu <strong>and</strong> Benet<br />

(2005), the transporter effects are minimal for class I drugs; efflux transporter effects<br />

predominate for the class II drugs <strong>and</strong> transporter-enzyme interplay in the intestines is<br />

important for the CYP3A substrates <strong>and</strong> phase 2 conjugation enzymes; absorptive<br />

transporter effects predominate for the class III drugs <strong>and</strong>; absorptive or efflux transporter<br />

effects could be important for the class IV drugs.<br />

2.4.3 BCS Biowaivers<br />

BCS is utilized for the so-called biowaivers, which means that in vivo bioavailability<br />

<strong>and</strong>/or bioequivalence studies may be waived. In that case a dissolution test could be<br />

adopted as the surrogate basis for the decision as to whether two pharmaceutical products<br />

are equivalent. According to the regulatory authorities (e.g. FDA, 2000; EMEA, 2002),<br />

only highly soluble <strong>and</strong> highly permeable BCS class I drugs can have a biowaiver status in<br />

IR solid oral-dosage forms. Furthermore, biowaiver products may not contain excipients<br />

that could influence the absorption <strong>of</strong> a drug compound, drug compound can not have a<br />

narrow therapeutic index, <strong>and</strong> the product can not be designed to be absorbed from the<br />

oral cavity.<br />

Yu et al. (2002) have proposed to extend the biowaiver status to BCS class II <strong>and</strong> III<br />

drugs. They criticized the regulatory limits <strong>of</strong> solubility <strong>and</strong> permeability being too strict.<br />

The changes they proposed were to narrow the required solubility pH range to the 1.0-6.8<br />

<strong>and</strong> to reduce the high permeability requirement from 90% to 85%. Also the boundaries<br />

for solubility <strong>of</strong> acidic BCS class II drugs have been shown to be too restrictive by<br />

Yazdanian et al. (2004). Other studies (Cheng et al., 2004; Kortejärvi et al., 2005;<br />

Kortejärvi et al., 2007; Kortejärvi et al., 2010) <strong>and</strong> reports (WHO, 2006) have been<br />

published recommending a biowaiver status for BCS class III drugs. Benet et al. (2008)<br />

recommended the regulatory authorities to add the extent <strong>of</strong> drug metabolism (i.e. 90%


metabolized) as an alternative method for the extent <strong>of</strong> drug absorption (i.e. 90%<br />

absorbed) in defining class I drugs suitable for biowaivers. This would increase the<br />

number <strong>of</strong> class I drugs eligible for biowaivers.<br />

29


3 Aims <strong>of</strong> the study<br />

The overall aim <strong>of</strong> the study was to develop predictive, reliable <strong>and</strong> rapid kinetic <strong>and</strong><br />

equilibrium solubility <strong>and</strong> dissolution screening methods, which do not require large<br />

amounts <strong>of</strong> compounds <strong>and</strong> would (ultimately) correlate well with in vivo results. Several<br />

solubility <strong>and</strong> dissolution testing methods were developed <strong>and</strong> characterized by using<br />

different volumes <strong>of</strong> aqueous buffer solutions <strong>and</strong> fasted state HIF as media. The specific<br />

aims <strong>of</strong> the study were:<br />

1. To construct <strong>and</strong> characterize the channel flow method <strong>and</strong> to compare it with<br />

the regulatory USP basket <strong>and</strong> intrinsic dissolution methods (I, II).<br />

2. To miniaturize dissolution testing stepwise from the regulatory dissolution<br />

methods <strong>and</strong> shake-flask solubility measurements via micro dissolution method<br />

to the 96-well plates (micro dissolution testing is previously unpublished data).<br />

3. To develop 96-well plate HTS methodology to determine both the equilibrium<br />

solubility <strong>and</strong> dissolution rate while avoiding the co-solvent(s) related<br />

problems. The model compounds were initially in solid form as in the<br />

traditional equilibrium solubility method in shake-flasks (IV).<br />

4. To develop kinetic HTS drug solubility testing method suitable for compounds<br />

having problems with UV absorption. The method is based on compounds’<br />

surface tension measurements by a novel microtensiometer (III).<br />

5. To compare the aqueous equilibrium solubility/dissolution results in the 96well<br />

plate method with corresponding testings in fasted state HIF medium (IV).<br />

30


4 Experimental<br />

4.1 Materials<br />

4.1.1 Chemicals (I-IV)<br />

The model compounds <strong>and</strong> references to the corresponding publications are summarized<br />

in Table 9. The model compounds were from Orion Pharma (Finl<strong>and</strong>) except for<br />

antipyrine (Aldrich Chemical Company Inc., USA), dicl<strong>of</strong>enac sodium (Sigma-Aldrich<br />

Chemie GmbH, Germany), indomethacin (Fluka Biochemika, Italy), piroxicam (Hawkins<br />

Inc., USA) <strong>and</strong> theophylline anhydrous (Helm AG, Germany). As tablet excipients were<br />

used microcrystalline cellulose (MCC) (Avicel PH 102, FMC, Irel<strong>and</strong>) (I, II), lactose<br />

(monohydr., 80 M, DMV International, The Netherl<strong>and</strong>s) (I, II), magnesium stearate<br />

(University Pharmacy, Finl<strong>and</strong>) (II) <strong>and</strong> talc (Pharmia Oy, Finl<strong>and</strong>) (II). All the other<br />

laboratory chemicals were <strong>of</strong> analytical grade.<br />

4.1.2 Human intestinal fluid (HIF) (IV)<br />

The HIF was collected from five volunteers after an overnight fast by introducing one<br />

double-lumen catheter (Salem Sump Tube 14 Ch, external diameter 4.7 mm, Sherwood<br />

Medical, Petit Rechain, Belgium) in the duodenum (D2/D3) via the mouth. Sampling <strong>of</strong><br />

HIF was continued for 3 hours every 15 minutes. HIF was stored at -30 o C or colder <strong>and</strong><br />

before solubility studies the individual HIF samples were pooled together.<br />

4.2 Methods<br />

4.2.1 Preparation <strong>of</strong> tablets (I, II)<br />

Tabletting masses <strong>of</strong> 200 g (I) <strong>and</strong> 150 g (II) were prepared. The powders were mixed for<br />

15 min in turbula mixer (Turbula T2C, Willy A. Bach<strong>of</strong>en AG Machinenfabrik,<br />

Switzerl<strong>and</strong>). Tablets were directly compressed with a single punch tabletting machine<br />

(Korsch EK-0, Erweka Apparatebau, Germany) using a flat punch with a diameter <strong>of</strong> 9<br />

mm. The target weights <strong>of</strong> tablets were 250 mg.<br />

31


Table 9. The model compounds.<br />

Model compound<br />

Molecular<br />

structure<br />

BCS<br />

class<br />

32<br />

pKa<br />

Acidic<br />

/basic<br />

Paper<br />

Acetylsalicylic acid C16H12O6 I 3.48 A I<br />

Antipyrine C11H12N2O I 0.7 B IV<br />

Atenolol C14H22N2O3 III 9.16 B IV<br />

Caffeine C8H10N4O2 I 0.73 B IV<br />

Carbamazepine C15H12N2O II 13.94 A IV<br />

Dicl<strong>of</strong>enac sodium C14H10Cl2NNaO2 II 4.18 A IV<br />

Furosemide C12H11ClN2O5S IV 3.04 A IV<br />

Hydrochlorothiazide C7H8ClN3O4S2 III 8.95 A IV<br />

Ibupr<strong>of</strong>en C13H18O2 II 4.41 A III, IV<br />

Indomethacin C19H16ClNO4 II 3.93 A III, IV<br />

Ketopr<strong>of</strong>en C16H14O3 II 4.23 A IV<br />

Metformin HCl C4H12ClN5 III 13.86 B IV<br />

Metoprolol tartrate C15H25NO3 I 9.17 B IV<br />

Naproxen C14H14O3 II 4.84 A IV<br />

Phenytoin sodium C15H11N2NaO2 II 8.33 A IV<br />

Piroxicam C15H13N3O4S II 3.80,4.50,<br />

13.33<br />

A/B IV<br />

Propranolol HCl C16H22ClNO2 I 9.14 B IV<br />

Ranitidine HCl C13H23ClN4O3S III 8.40 B IV<br />

Rifampicin C43H58N4O12 II 9.92 A IV<br />

Spironolactone C24H32O4S IV N.A. N.A. IV<br />

Theophylline<br />

anhydrous<br />

C7H8N4O2 II 8.60,1.70 A/B II<br />

Trimethoprim C14H18N4O3 II 7.2 B IV<br />

4.2.2 <strong>Dissolution</strong> methods<br />

4.2.2.1 Basket method (I, II)<br />

USP basket method (Sotax AT 7, Sotax, Switzerl<strong>and</strong>) was used as a reference dissolution<br />

method (USP XXVI, 2003). The rotational speed was 50 rpm, the volume <strong>of</strong> medium was


900 ml <strong>and</strong> the temperature was 23 o C. The samples were analyzed by a spectrophotometer<br />

(Perkin-Elmer Lambda 2, Germany).<br />

4.2.2.2 Intrinsic dissolution rate method (I, II)<br />

The intrinsic dissolution test was done according to the USP XXVI (2003), which is based<br />

on Wood’s rotating disk method (Wood, 1965). The model compounds (100 mg) were<br />

pressed with a hydraulic pressure <strong>of</strong> 1500 kg (Pye Unicam, UK) <strong>and</strong> rotated at 50 rpm in<br />

900 ml <strong>of</strong> the medium at the temperature <strong>of</strong> 23 o C. (Sotax AT 7, Sotax, Germany). The<br />

samples were conducted into a spectrophotometer (Perkin-Elmer Lambda 2, Germany) for<br />

analyses.<br />

4.2.2.3 Channel flow method (I, II)<br />

In the channel flow system only one tablet surface was in contact with the media (Figure<br />

3). The tablet was located at the centre <strong>of</strong> the channel flow cell in order to avoid the edgeeffects<br />

(Compton et al., 1988). The flow rate <strong>of</strong> the aqueous phase was controlled by a<br />

peristaltic pump (Ismatec IPN-12, Switzerl<strong>and</strong>) <strong>and</strong> temperature (23 o C) by a thermostatted<br />

bath. The channel flow cell was coupled to a UV-VIS spectrophotometer (HP8541A,<br />

Hewlett-Packard, USA).<br />

Figure 3 a) The channel flow cell <strong>and</strong> b) the channel flow test apparatus. Modified from<br />

publications I <strong>and</strong> II<br />

.<br />

4.2.2.4 Micro dissolution method (previously unpublished data)<br />

By this method the solubility <strong>of</strong> selected model compounds (ibupr<strong>of</strong>en, ketopr<strong>of</strong>en,<br />

indomethacin <strong>and</strong> piroxicam) was determined in 0.1 M hydrochloric acid solution in pH<br />

1.2 <strong>and</strong> 6.8 buffer solutions <strong>and</strong> in biorelevant medium, FaSSIF. Excess <strong>of</strong> drug<br />

33


compound was added into 2 ml flat-bottomed test tubes <strong>of</strong> the micro dissolution apparatus<br />

μDISS ® (pION, USA). <strong>Dissolution</strong> medium was added into the test tubes <strong>and</strong> the solution<br />

was stirred at 180 rpm using magnet stirrers. The temperature <strong>of</strong> the test solutions was<br />

maintained at 37 o C with the aid <strong>of</strong> a water bath surrounding the test tube block. The test<br />

solutions were kept under these circumstances for 3 hours after which the solutions were<br />

filtered using preheated Millipore PVDF 0.45 μm filters. The solutions were then left at<br />

37 o C for 24 hours <strong>and</strong> the concentration <strong>of</strong> the solutions were measured by detecting the<br />

UV absorbance <strong>of</strong> the solutions in situ.<br />

Quantification <strong>of</strong> the sample solutions was determined against a calibration curve with<br />

four to five dilutions <strong>of</strong> st<strong>and</strong>ard stock solution. The concentrations <strong>of</strong> the solutions were<br />

determined at the absorption maximum wavelength. Further, to assure the buffer capacity<br />

<strong>of</strong> the solutions, pH <strong>of</strong> the sample solutions was measured after the solubility runs.<br />

4.2.2.5 96-well plate method (IV)<br />

Model drugs were first dissolved into a presolvent (methanol or water), <strong>and</strong> the presolvent<br />

solutions were then pipetted into 96-well plates (96F untreated 260895, Nunc, Denmark),<br />

after which the presolvent was evaporated in order to avoid co-solvent effects. After<br />

evaporation, 250 μl <strong>of</strong> the medium was added to the wells <strong>and</strong> the 96-well plates were<br />

shaken orbitally at 600 rpm (Heidolph titramax 100, Germany). Samples were withdrawn<br />

after 30 min, 1, 2, 5, 24 <strong>and</strong> 30 hours <strong>and</strong> pipetted into UV 96-well plates (UV FB<br />

microtiter plates 8404, Thermo Fisher Oy, USA) for analyses with UV plate reader<br />

λ=230-400 nm (Varioskan, Thermo Electron corporation, Finl<strong>and</strong>). The tests were<br />

performed at 23 o C.<br />

4.2.3 <strong>Solubility</strong> methods<br />

4.2.3.1. Shake-flask method (III, IV)<br />

Shake-flask method was used as a reference method in the solubility tests (Lipinski,<br />

2003). The model compounds were weighted into 25 ml (III) or 40 ml (IV) glass vials,<br />

which were mixed by magnetic stirrer at 100 rpm at 23 o C during the tests. Samples were<br />

taken after 1, 2, 3, 4, 5 <strong>and</strong> 24, even after 30 hours for some compounds, until the<br />

equilibrium plateau was reached. Samples were filtered (0.45 μm, Minisart, Sartorius,<br />

UK) before analyses by UV-spectrophotometry (Pharmacia LKB, Ultrospec, Sweden).<br />

34


4.2.3.2 96-well plate method (IV)<br />

The method was described in chapter 4.2.2.5.<br />

4.2.3.3 Surface tension based kinetic HTS solubility method (III)<br />

The model drugs were first dissolved into a presolvent (methanol, ethanol or DMSO). 12-<br />

22 samples were prepared so that the first one had a drug concentration <strong>of</strong> ca. 100 mg/ml,<br />

<strong>and</strong> each subsequent sample was diluted to half the concentration <strong>of</strong> the previous one.<br />

Final sample was always a blank sample with no drug. Then, the presolvent solutions were<br />

added to the aqueous medium in the 96-well plate by a volume <strong>of</strong> 50 μl. Surface tension<br />

based measurements were performed with recently introduced Delta-8 multichannel<br />

microtensiometer (Kibron Inc., Finl<strong>and</strong>).<br />

4.2.4 Analyses<br />

4.2.4.1 Analyses <strong>of</strong> human intestinal fluid (HIF) (IV)<br />

Bile salts were measured by GC-MS-selected ion monitoring analysis developed by<br />

Lütjohann et al. (2004). Free bile acids were extracted with diethylether <strong>and</strong><br />

trimethylsilylated after alkaline hydrolysis <strong>and</strong> acidification. The chromatographic<br />

separation was performed with a DB-XLB capillary column in a constant flow mode <strong>of</strong><br />

0.8 ml/min with a final temperature <strong>of</strong> 290 o C having helium as a carrier gas.<br />

Total phospholipid levels were determinated enzymatically according to Gurantz et al.<br />

(1981), using phospholipase D <strong>and</strong> choline oxidase. The reagent kit was from Wako<br />

Chemicals (Neuss, Germany).<br />

4.2.4.2 Solid-state characterization (IV)<br />

X-ray powder diffraction (XRPD) was used to study the solid state forms <strong>of</strong> the model<br />

compounds <strong>and</strong> to confirm possible (pseudo)polymorphic changes. Measurements were<br />

done by an X-ray powder diffraction theta-theta diffractometer (Bruker AXS D8 advance,<br />

Bruker AXS GmbH, Germany). The XRPD experiments were performed in symmetrical<br />

reflection mode with CuKα radiation (1.54 Å) using Göbel Mirror bent gradient multilayer<br />

optics. The scattered intensities were measured with a scintillation counter. The angular<br />

range was from 5 o to 30 o with the steps <strong>of</strong> 0.1 o <strong>and</strong> the measuring time was 20 s/step.<br />

The reference codes to identify experimental polymorphs were from CSD database<br />

(Cambridge Structural Database, The Cambridge Crystallographic Data Centre, UK)<br />

35


(Allen, 2002). The reference crystal structure codes were for dicl<strong>of</strong>enac form I (P21/c)<br />

SIKLIH02 (Castellari <strong>and</strong> Ottani, 1997), dicl<strong>of</strong>enac form II (C2/c) SIKLIH05 (Muangsin<br />

et al., 2004), dicl<strong>of</strong>enac form III (Pcan) SIKLIH04 (Jaiboon et al., 2001), indomethacin αform<br />

INDMET2 (Chen et al., 2002a), indomethacin γ-form INDMET (Kistenmacher <strong>and</strong><br />

Marsh, 1972), piroxicam monohydrate CIDYAP0 (Bordner, 1984), piroxicam α-form<br />

BIYSEH06 (Vrecer et al., 2003), piroxicam β-form BIYSEH (Kojic-Prodic et al., 1982)<br />

<strong>and</strong> spironolactone polymorph 1 from acetone ATPRCL10 (Dideborg <strong>and</strong> Dupont, 1972).<br />

Differential scanning calorimetry (DSC) was used to further confirm the XRPD<br />

results. The thermal analyses were performed with a Mettler DSC 823e (Mettler-Toledo<br />

AG, Switzerl<strong>and</strong>) <strong>and</strong> analyzed by STAR e s<strong>of</strong>tware (Mettler-Toledo AG, Switzerl<strong>and</strong>).<br />

Samples were weighted (4-5 mg) to the aluminium pans <strong>and</strong> crimped by aluminium caps<br />

with a pinhole. The samples were heated at the rate <strong>of</strong> 10 o C/min from 25 o C to the 10 o C<br />

above the melting point <strong>of</strong> each compound. The measurements were performed under N2<br />

atmosphere <strong>and</strong> the calibration was performed with indium.<br />

36


5 Results <strong>and</strong> discussion<br />

5.1 <strong>Dissolution</strong> testing<br />

5.1.1 Channel flow method vs. the regulatory dissolution methods (I, II)<br />

<strong>Dissolution</strong> pr<strong>of</strong>iles <strong>of</strong> pure acetylsalicylic acid <strong>and</strong> theophylline from the channel flow<br />

method correlated well with the intrinsic dissolution pr<strong>of</strong>iles. In both methods the<br />

dissolution rate <strong>of</strong> acetylsalicylic acid at the time-point 25-min was around 5-10 mg/cm 2 at<br />

pH 1.2 <strong>and</strong> around 20-25 mg/cm 2 at pH 6.8 (I). The dissolution rates <strong>of</strong> theophylline at the<br />

time-point 10-min were around 11-12 mg/cm 2 at pH 1.2 <strong>and</strong> 10 mg/cm 2 at pH 6.8 by both<br />

the methods (II).<br />

The solubility <strong>of</strong> acetylsalicylic acid tablets (I) containing microcrystalline cellulose<br />

(MCC) (compositions A <strong>and</strong> C) was slightly increased by the lowering pH in the channel<br />

flow method opposite to the basket method. This could be due to the MCC swelling<br />

properties being in more important role in the channel flow method compared to the<br />

solubility <strong>of</strong> acetylsalicylic acid. Swelling <strong>of</strong> MCC affects the flow <strong>and</strong> the mass transfer<br />

characteristics hindering MCC’s disintegrating effect <strong>of</strong> the tablet in the CFC. The<br />

dissolution results with tablets containing only lactose as an excipient correlated well<br />

between the methods <strong>and</strong> the rank order <strong>of</strong> the dissolution pr<strong>of</strong>iles with all tablet batches<br />

(A-C) was similar by both methods (Figure 4).<br />

37


(a)<br />

(c)<br />

pH 1.2 pH 1.2<br />

pH 6.8<br />

Figure 4 Channel flow dissolution <strong>of</strong> acetylsalicylic acid tablet formulations A-C at pH 1.2 (a)<br />

<strong>and</strong> at pH 6.8 (c) (Flow rate was ca. 190 ml/h). <strong>Dissolution</strong> <strong>of</strong> acetylsalicylic acid<br />

with the USP basket method at pH 1.2 (b) <strong>and</strong> at pH 6.8 (d). Modified from paper I.<br />

The tablet compositions were more complicated with theophylline (II) containing four<br />

excipients (MCC, lactose, talc, magnesium stearate). The amounts <strong>of</strong> MCC <strong>and</strong> lactose<br />

varied between the tablet batches (A-D). The drug release rate measured by the channel<br />

flow method was decreased as the amount <strong>of</strong> MCC was increased. The USP basket<br />

method gave different result; the slowest drug release was from the tablets having the<br />

highest amount <strong>of</strong> lactose (A) or MCC (D). The difference can be explained by the<br />

different dissolution set-ups. In the channel flow method the position <strong>of</strong> the tablet hinders<br />

the disintegrating effect <strong>of</strong> MCC <strong>and</strong> even the formulation C (MCC 56%) remained as a<br />

coherent matrix. From the basket method the tablet formulation C disintegrated<br />

immediately. Also with theophylline the two methods lead to similar conclusions (Figure<br />

5).<br />

(b)<br />

(d)<br />

38<br />

B<br />

B A<br />

A<br />

C<br />

A C<br />

pH 6.8


(a)<br />

(c)<br />

pH 1.2<br />

pH 6.8<br />

(d)<br />

(b)<br />

Figure 5 Channel flow dissolution <strong>of</strong> theophylline from tablet formulations A-D at pH 1.2 (a)<br />

<strong>and</strong> at pH 6.8 (c) (Flow rate was ca. 190 ml/h). <strong>Dissolution</strong> <strong>of</strong> theophylline with the<br />

USP basket method at pH 1.2 (b) <strong>and</strong> at pH 6.8 (d) (A , B , C, D). Modified<br />

from paper II.<br />

The main difference between the channel flow <strong>and</strong> USP basket methods is that in the<br />

channel flow method only one tablet surface is exposed to the dissolution medium <strong>and</strong>,<br />

thus, the disintegration <strong>of</strong> the tablet is hindered. This enables collection <strong>of</strong> accurate<br />

information from the initial steps <strong>of</strong> the dissolution process. The intrinsic dissolution<br />

method is based on the idea <strong>of</strong> keeping the dissolution area <strong>of</strong> pure compound constant<br />

(Nicklasson et al., 1981; Nicklasson <strong>and</strong> Brodin, 1984) <strong>and</strong> achieving the dissolution by<br />

rotating or shear-like motion in the medium, which differs from the laminar flow in the<br />

channel flow method (Compton et al., 1988). Also the mass transfer can be followed with<br />

the channel flow method, which allows comparison between the different dissolution<br />

mechanisms <strong>and</strong> the experimental results. Channel flow cell method is also used in the<br />

modelling because <strong>of</strong> the well-defined laminar flow in the CFC (Aaltonen et al., 2006;<br />

Lehto et al., 2008).<br />

39<br />

pH 1.2<br />

pH 6.8


5.1.2 Micro dissolution studies in different media (previously unpublished<br />

data)<br />

Micro dissolution studies were performed in Orion Pharma with four model drugs<br />

(ibupr<strong>of</strong>en, indomethacin, ketopr<strong>of</strong>en <strong>and</strong> piroxicam) in three different media (aqueous<br />

solution pH 1.2, buffer solution pH 6.8 <strong>and</strong> FaSSIF) at +37 o C. Table 10 presents results<br />

comparing equilibrium solubilities with micro dissolution, shake-flask <strong>and</strong> equilibrium 96well<br />

plate methods. The latter were both performed at room temperature (23±1 o C).<br />

Table 10. Equilibrium solubilities (mg/ml) <strong>of</strong> four model compounds in different media<br />

(SF=Shake-flask, MD=Micro dissolution <strong>and</strong> HTS=equilibrium 96-well plate<br />

method, n=2).<br />

Compound<br />

Ibupr<strong>of</strong>en<br />

Indomethacin<br />

Ketopr<strong>of</strong>en<br />

Piroxicam<br />

SF<br />

pH 1.2<br />

0.07<br />

0.004<br />

0.08<br />

0.06<br />

MD<br />

pH 1.2<br />

0.05<br />

0.001<br />

0.13<br />

0.19<br />

SF<br />

pH 6.8<br />

2.9<br />

0.4<br />

5.5<br />

0.2<br />

40<br />

MD<br />

pH 6.8<br />

3.5<br />

0.6<br />

5.4<br />

0.6<br />

MD<br />

FaSSIF<br />

pH 6.5<br />

3.3<br />

0.6<br />

5.6<br />

0.4<br />

HTS<br />

HIF<br />

pH 6.2<br />

2.0<br />

2.4<br />

3.2<br />

0.8<br />

As expected, the equilibrium drug solubilities by the micro dissolution <strong>and</strong> shake-flask<br />

methods were higher at pH 6.8 than at pH 1.2 (pKa values <strong>of</strong> the model compounds are<br />

mostly around 4). The micro dissolution solubilities in the pH 6.8 buffer solution <strong>and</strong><br />

FaSSIF (pH 6.5) correlated well suggesting that the buffer might be an adequate medium<br />

at this solubility scale for the model compounds, although FaSSIF contains taurocholic<br />

acid <strong>and</strong> lecithin. Instead, the solubility results in HIF (pH 6.2) were somewhat lower for<br />

ibupr<strong>of</strong>en <strong>and</strong> ketopr<strong>of</strong>en <strong>and</strong> somewhat higher for piroxicam <strong>and</strong> indomethacin compared<br />

to the aqueous buffer pH 6.8 or FaSSIF pH 6.5. The compositions <strong>of</strong> fasted state HIF <strong>and</strong><br />

FaSSIF are more complicated than the composition <strong>of</strong> aqueous buffer at pH 6.8 containing<br />

e.g. bile acids, which enhance solubility due to the solubilization effect (Bakatselou et al.,<br />

1991; Charman et al., 1993) <strong>and</strong>/or wetting (Bakatselou et al., 1991). Here, the dissolution<br />

method (micro dissolution vs. 96-well plate) has important role to give similar results.<br />

According to our study, the medium (aqueous buffer, FaSSIF, HIF) has minor role. The<br />

use <strong>of</strong> HIF (<strong>and</strong> FaSSIF) in the solubility studies may have more pronounced importance<br />

for very low solubility compounds at the discovery stage, although the more simple, fast<br />

<strong>and</strong> economic methods/techniques are preferred also there.


5.1.3 <strong>Dissolution</strong> pr<strong>of</strong>iles in 96-well plate (IV)<br />

Besides the equilibrium solubility determinations, the 96-well plate method can be used to<br />

study the dissolution rates <strong>of</strong> drugs as a function <strong>of</strong> time (IV). Figure 6 distinguishes<br />

nicely the dissolution pr<strong>of</strong>iles <strong>of</strong> ibupr<strong>of</strong>en <strong>and</strong> indomethacin regardless the shaking<br />

speeds in question. At the moment the method is useful for the classification <strong>of</strong> dissolution<br />

rates because <strong>of</strong> rather long centrifugation times before the sampling. By shortening the<br />

centrifugation time, more detailed dissolution studies are possible for more rapidly<br />

dissolving materials.<br />

Concentration (mg/ml)<br />

Figure 6 Aqueous dissolution pr<strong>of</strong>iles <strong>of</strong> ibupr<strong>of</strong>en <strong>and</strong> indomethacin at different shaking<br />

speeds (300, 600 <strong>and</strong> 1200 rpm) (n=3). Modified from paper IV.<br />

5.1.4 Scaling down the dissolution testings (I, II, III, IV)<br />

The scaling down <strong>of</strong> the dissolution testing from the traditional basket method (e.g. USP,<br />

2009; Ph. Eur., 2009) via micro dissolution method to the 96-well plates was successful.<br />

Main focus in the dissolution testing has been in the quality control <strong>and</strong> the several stages<br />

<strong>of</strong> drug development (FDA, 1997a). The large amount <strong>of</strong> drug compound needed has <strong>of</strong>ten<br />

limited the use <strong>of</strong> dissolution testing at the drug discovery stage, which the new methods<br />

now enable.<br />

Alongside with the regulatory dissolution methods, the small scale dissolution methods<br />

can be used to study the results, e.g., at the initial states (0-5 min) <strong>of</strong> the dissolution<br />

process. <strong>Dissolution</strong> pr<strong>of</strong>iles from the early stage <strong>of</strong> drug discovery process can be used to<br />

improve the knowledge on lead c<strong>and</strong>idates. Further, while more complicated, hard to get<br />

<strong>and</strong> expensive media, e.g. HIF, can be utilized with these small scale methods, if needed.<br />

41


5.2 <strong>Solubility</strong> screening<br />

5.2.1 HTS 96-well plate methods (III, IV)<br />

Both kinetic (III) <strong>and</strong> equilibrium (IV) 96-well plate methods were developed in this thesis<br />

work. The fact that amphiphilic compounds decrease the surface tension (γ) <strong>of</strong> water<br />

linearly within a concentration range limited by air/water partitioning at the lower<br />

concentration end, <strong>and</strong> solubility/critical micelle concentration (CMC) at the higher end,<br />

was utilized in the development <strong>of</strong> the kinetic method. The solubility value can be<br />

determined from these isotherms, if the compound is more likely to precipitate than to<br />

form micelles. The phenomenon is described by the classical Gibbs adsorption isotherm:<br />

Γ = −<br />

1 ∂γ<br />

RT ∂ ln c<br />

where Γ is surface excess, R is gas constant, T is temperature <strong>and</strong> c is concentration.<br />

The measured Gibbs isotherm for ibupr<strong>of</strong>en is shown in Figure 7. The solubility <strong>of</strong><br />

ibupr<strong>of</strong>en can be obtained from the focal point between the sloping part <strong>and</strong> the level part<br />

<strong>of</strong> the isotherm. After this point, solid <strong>and</strong> solvated phases are in equilibrium <strong>and</strong> the<br />

amount <strong>of</strong> the free compound is constant in the solution. At higher concentrations, all<br />

material exceeding the solubility limit will be in the solid phase, which has no effect on<br />

the surface tension. Figure 7 also shows the effect <strong>of</strong> pH on the isotherm <strong>of</strong> ibupr<strong>of</strong>en. The<br />

solubility <strong>of</strong> ibupr<strong>of</strong>en is lower at pH 1.2 (0.19 mg/ml) than in pure water (0.23 mg/ml).<br />

This is seen from the shift <strong>of</strong> the adsorption curve.<br />

42<br />

(8)


Figure 7 Gibbs isotherm for ibupr<strong>of</strong>en in aqueous solution. Surface tension is reduced with<br />

increasing ibupr<strong>of</strong>en concentration until a solubility limit is reached. By lowering the<br />

pH decreases the solubility <strong>of</strong> ibupr<strong>of</strong>en, which is seen from the transition point at<br />

lower concentration (n=12, methanol was used as a presolvent). Modified from paper<br />

III.<br />

The slope <strong>of</strong> the linear part just before the transition point is inversely proportional to the<br />

cross-sectional area <strong>of</strong> the compound at the interface (Suomalainen et al., 2004) (Figure<br />

7). With ibupr<strong>of</strong>en the molecular area differed between the two pHs; the absolute value <strong>of</strong><br />

the slope for ibupr<strong>of</strong>en at pH 6.8 was 43.0 corresponding the cross-sectional area per<br />

molecule 0.0096 Å 2 , compared to the absolute value <strong>of</strong> the slope (17.9) <strong>and</strong> cross-sectional<br />

area (0.023 Å 2 ) at pH 1.2. With indomethacin the slopes <strong>and</strong> cross-sectional areas were<br />

similar in both media. These differencies may be due to the different interactions between<br />

solvent <strong>and</strong> solute molecule.<br />

A linear dependency was found between the solubilities in the microtensiometer <strong>and</strong><br />

shake-flask measurements for ibupr<strong>of</strong>en (Figure 8). Best fit <strong>of</strong> the linear curve gave a<br />

slope <strong>of</strong> 1.16 with an intercept at around 180 μg/ml. Differences in the solubility between<br />

the samples were clearly detected, <strong>and</strong> the correlation was at a reasonably good level for<br />

fast screening purposes <strong>and</strong>, thus, clear distinction between low <strong>and</strong> high solubility<br />

environments is possible. The indomethacin results did not fall on the same fitted line as<br />

the ibupr<strong>of</strong>en results, which is likely due to the differences in the precipitation rates <strong>of</strong> the<br />

two compounds. Still, also with indomethacin it was possible to distinguish between the<br />

high <strong>and</strong> low solubility conditions. Not at all unexpectedly, for both the model compounds<br />

the solubilities measured by the surface tension method were consistently higher than<br />

those from the (equilibrium) shake-flask method − due to the co-solvent effect (Fligge <strong>and</strong><br />

Schuler, 2006). This is common feature to all HTS kinetic methods (Pan et al., 2001; Chen<br />

et al., 2002b; Bard et al., 2008) (Figures 8 <strong>and</strong> 9).<br />

43


Figure 8 Measured ibupr<strong>of</strong>en solubilities with the novel surface tension method (y-axis) as a<br />

function <strong>of</strong> the solubilities from the shake-flask measurements (x-axis). Modified from<br />

paper III.<br />

Several variables were tested in order to study the effects <strong>of</strong> different solvent systems on<br />

the reliability <strong>and</strong> applicability <strong>of</strong> the microtensiometry measurements. Studied media<br />

environments included, e.g., electrolytes, <strong>and</strong> surfactants. NaCl <strong>and</strong> CaCl2 (0.9% <strong>and</strong> 2%)<br />

had only limited effects on the solubility <strong>of</strong> ibupr<strong>of</strong>en. Surfactants, sodium dodecyl<br />

sulphate (SDS) <strong>and</strong> Tween 80, increased the solubility slightly below their CMC values,<br />

but the increase was much more evident above the CMC values. The pH effect on the<br />

acidic model compounds was as expected; the higher the pH, the higher the solubility.<br />

Most importantly, the presence <strong>of</strong> other materials in the medium did not disturb the<br />

determinations.<br />

44


Figure 9 <strong>Drug</strong> solubilities by the surface tension (black, n=8-12) <strong>and</strong> shake-flask (grey, n=2)<br />

measurements; effect <strong>of</strong> the dissolution media. Modified from paper III.<br />

Most <strong>of</strong> the HTS methods measure the kinetic solubility when the lead compound is at<br />

dissolved state in the beginning <strong>of</strong> the measurement <strong>and</strong> the co-solvent(s) may<br />

overestimate the solubility (Chen et al., 2002b). The HTS equilibrium solubility method<br />

was developed to avoid that, so in the starting point the compound was in the solid form.<br />

The results from the equilibrium 96-well plate correlated well (R 2 =0.931) with the values<br />

from the shake-flask method over five orders <strong>of</strong> magnitude in aqueous solubility with<br />

twenty model drugs (Figure 10) (IV). The relative deviation <strong>of</strong> the 96-well plate results<br />

was always less than 7-fold. Relative overestimation by the 96-well plate method was the<br />

most significant for compounds with solubility values lower than 10 μg/ml (circled area in<br />

Figure 10). Still, the 96-well plate method gave reliable solubility pr<strong>of</strong>iling: 98% <strong>of</strong> the<br />

tests were within the predictive confidence intervals <strong>of</strong> 95%.<br />

45


Figure 10 The solubilities from shake-flask (y-axis) <strong>and</strong> equilibrium HTS methods (x-axis).<br />

Modified from paper IV.<br />

Despite the medium, the methods gave similar values. Even 10.000-fold changes in<br />

solubility were detected depending on the medium, but the correlation between the<br />

equilibrium HTS <strong>and</strong> shake-flask methods remained good. For example, the solubility <strong>of</strong><br />

dicl<strong>of</strong>enac sodium at pH 1.2 was 0.002 mg/ml <strong>and</strong> at pH 6.8 15.3 mg/ml using the HTS<br />

method; the corresponding values obtained by the shake-flask method were 0.003 mg/ml<br />

<strong>and</strong> 24.0 mg/ml, respectively. The solubilities were determined accurately over a wide<br />

solubility scale (0.002–169.2 mg/ml), which is a true advantage <strong>of</strong> the 96-well plate<br />

method. Compounds with vastly different solubilities can be determined with the method<br />

at small drug quantities <strong>and</strong> utilize method in the early drug discovery stage when only<br />

small quantities <strong>of</strong> compounds are available. Equilibrium 96-well plate method from<br />

Alelyunas et al. (2009) included HPLC, which might slow down the process <strong>and</strong>,<br />

traditionally, the equilibrium solubility has not been measured until the early phase <strong>of</strong><br />

drug development stage (Lipinski, 2003).<br />

Also the kinetic HTS method (microtensiometry) is suitable for drug solubility<br />

screening purposes at the drug discovery stage. Although the method has few limitations,<br />

e.g. it might be difficult to distinguish between the precipitation <strong>and</strong> micelle formation<br />

behavior or some very soluble compounds might not sufficiently partition to the water/air<br />

interface, it has several benefits compared to other kinetic HTS methods. The method does<br />

not require st<strong>and</strong>ard curves or filtrations, solid residues do not disturb the determinations,<br />

large concentration scales are measurable <strong>and</strong> low UV absorption <strong>of</strong> poorly soluble drugs<br />

does not matter (Bevan <strong>and</strong> Lloyd, 2000; Pan et al., 2001). It can be stated that the method<br />

is very well utilizable for poorly soluble drugs, which partition sufficiently to the water/air<br />

interface. Kinetic HTS method was also used to measure the solubility <strong>of</strong> ibupr<strong>of</strong>en as a<br />

46


function <strong>of</strong> time. Samples were remeasured after 5 hours, in which time the solubility<br />

values did not change too much indicating that overestimation was more due to the cosolvent<br />

effect than supersaturation. The dissolution <strong>of</strong> ibupr<strong>of</strong>en was fast <strong>and</strong> the<br />

dissolution pr<strong>of</strong>ile as a function <strong>of</strong> time just ascertained that dissolution had taken place. It<br />

is compound specific, how fast the dissolution takes place <strong>and</strong> the equilibrium solubility is<br />

reached.<br />

Further, ibupr<strong>of</strong>en <strong>and</strong> indomethacin results from both the 96-well plate methods were<br />

compared to each other at aqueous media pH 1.2 <strong>and</strong> 6.8. In both cases the 96-well plate<br />

methods gave maximum 7-fold higher solubility values compared to the shake-flask<br />

method. Only when using ethanol or methanol as presolvents the kinetic solubility values<br />

were more than 7-fold higher than the shake-flask solubility values (Figure 8). The<br />

overestimation by the HTS 96-well plate methods compared to the shake-flask method<br />

could be due to the different scales more than due to the partitioning during the kinetic <strong>and</strong><br />

equilibrium solubility testings.<br />

5.2.2 Polymorphic changes (IV)<br />

During the shake-flask solubility tests, changes in the solid-state form <strong>of</strong> piroxicam were<br />

recognized; the solubility <strong>of</strong> piroxicam was first increased in the aqueous solutions, after<br />

which it was decreased to the equilibrium level. Also the color <strong>of</strong> the solid residue was<br />

changed during the test from white (anhydrate) to yellow (monohydrate) indicating the<br />

change in the physical state <strong>of</strong> the material. The changes were confirmed by XRPD <strong>and</strong><br />

DSC techniques. Piroxicam raw material contains α- <strong>and</strong> β-forms (both anhydrates), but<br />

crystalline solid residues <strong>of</strong> piroxicam monohydrate <strong>and</strong> β-form (anhydrate) were formed<br />

in the solubility test at 5h time-point (Kojic-Prodic et al., 1982; Bordner, 1984; Vrecer et<br />

al., 2003) (Figure 11). For the other model compounds, no polymorphic changes were<br />

noticed during the shake-flask determinations.<br />

47


Figure 11 XRPD diffraction patterns <strong>of</strong> piroxicam in the shake-flask measurements. The<br />

diffractograms represent the solid residue recrystallized from methanol (starting<br />

material in HTS), raw material (starting material in shake-flask), pure piroxicam<br />

monohydrate <strong>and</strong> anhydrated α- <strong>and</strong> β-forms.<br />

In the equilibrium HTS method (pseudo)polymorphic changes occurred with four model<br />

drugs; piroxicam, indomethacin, dicl<strong>of</strong>enac sodium <strong>and</strong> spironolactone (Table 11) (Figure<br />

11). From XRPD <strong>and</strong> DSC results it was obvious that the starting materials in the shakeflask<br />

<strong>and</strong> HTS methods were not <strong>of</strong> the same form. Also spironolactone is known to<br />

crystallize from different solvents (e.g. methanol, acetonitrile, ethanol, ethyl acetate <strong>and</strong><br />

benzene) in different polymorphic forms (El-dalsh et al., 1983; Beckstead et al., 1993).<br />

During the HTS solubility tests no polymorphic changes were, however, observed.<br />

Although some materials had (pseudo)polymorphic changes during the solubility testing,<br />

these changes were not problematic for the interpretation <strong>of</strong> the results, because both the<br />

shake-flask <strong>and</strong> 96-well plate methods do measure the equilibrium solubility <strong>and</strong> the<br />

possible transformations are easily observed. Still, possible polymorphic changes <strong>of</strong> drug<br />

c<strong>and</strong>idates may have an influence on the solubility results in solubility screening <strong>and</strong>, later<br />

on, on the drug development (Bauer et al., 2001).<br />

48


Table 11. Starting <strong>and</strong> raw material results from the XRPD (polymorphic form) <strong>and</strong> DSC<br />

(melting temperatures) measurements (n=3). Modified from paper IV.<br />

Compound XRPD DSC Reported melting<br />

temperature<br />

Dicl<strong>of</strong>enac sodium<br />

raw material (Shake-flask)<br />

starting material (HTS)<br />

Indomethacin<br />

raw material (Shake-flask)<br />

starting material (HTS)<br />

Piroxicam<br />

raw material (Shake-flask)<br />

starting material (HTS)<br />

I, III<br />

II, III<br />

γ<br />

α, γ<br />

α, β<br />

monohydrate<br />

49<br />

288.3 o C<br />

285.5 o C<br />

160.0 o C<br />

154.0 o C, 160.5 o C<br />

202.4 o C<br />

201.5 o C<br />

a Llinas et al., 2007; b Chen et al., 2002a; c Vrecer et al., 2003<br />

5.3 Human intestinal fluid (HIF)<br />

5.3.1 Characterization <strong>of</strong> fasted state HIF (IV)<br />

II=180.5 o C a<br />

γ=160-161 o C b<br />

α=152-154 o C b<br />

α=202.6 o C c<br />

β=199.7 o C c<br />

monohydr.=202.0 o C c<br />

HIF was collected from five healthy volunteers’ duodenum. Inter- <strong>and</strong> intraindividual<br />

variations were observed in the study (Tables 12 <strong>and</strong> 13), like in the earlier studies<br />

(Lindahl et al., 1997; Clarysse et al., 2009b). The mean pH value measured from the<br />

pooled HIF was 6.24 compared to values from 6.6 to 7.5 <strong>and</strong> the mean osmolality was 205<br />

mOsmkg -1 compared to values <strong>of</strong> 200–271 mOsmkg -1 (Table 6).<br />

The mean <strong>of</strong> the total bile salt concentration in fasted HIF was 3.63 mM, which is in<br />

line with the earlier literature values from 2.0 to 3.5 mM (Table 6). Instead, the<br />

phospholipid concentration (1.8 mM, median 1.1 mM) was higher than in the earlier<br />

studies (0.2–0.6 mM) (Table 6) because <strong>of</strong> one sample, which was very concentrated. The<br />

samples vary between individuals <strong>and</strong> individual concentrations <strong>of</strong> up to 2.7 mM have<br />

been found in fasted upper small intestine earlier (Clarysse et al., 2009b). The mean<br />

cholesterol level <strong>of</strong> HIF was 1.8 mM, which was close to the level found by Mansbach et<br />

al. (1975) (1.5 mM).


Table 12. Summary <strong>of</strong> volunteers’ (V1-5) characteristics <strong>and</strong> human intestinal fluid<br />

collection from duodenum.<br />

Volunteer Gender Age BMI<br />

(kg/m 2 Volume pH Osmolality<br />

) (ml)<br />

(mOsm/kg)<br />

V1 Male 25 21.7 70 6.62 176<br />

V2 Female 39 20.8 100 5.19 219<br />

V3 Male 37 23.1 100 6.60 209<br />

V4 Female 24 21.6 40 6.27 213<br />

V5 Female 31 20.9 35 6.51 207<br />

Mean<br />

- 29 21.6 69<br />

6.24 205<br />

±s.d.<br />

8<br />

0.9 31<br />

0.60 17<br />

Table 13. Composition <strong>of</strong> fasted HIF from five healthy volunteers (V1-5). In the table<br />

DC=deoxycholate, C=cholate, CDC=chenodeoxycholate <strong>and</strong><br />

UDC=ursodeoxycholate<br />

Bile<br />

salts (mM)<br />

Phospholipids<br />

(mM)<br />

Cholesterol<br />

(mM)<br />

50<br />

DC<br />

(%)<br />

C<br />

(%)<br />

CDC<br />

(%)<br />

V1 2.45 1.35 1.48 29.5 42.5 23.8 4.2<br />

V2 2.62 1.11 1.68 24.6 56.7 18.1 0.6<br />

V3 4.54 2.18 2.27 3.4 64.6 25.3 6.7<br />

V4 1.65 0.99 1.71 0.8 57.7 35.4 6.1<br />

V5 8.60 4.69 1.48 7.1 66.6 25.6 0.7<br />

5.3.2 HIF in solubility testing (IV)<br />

UDC<br />

(%)<br />

Equilibrium HTS solubility tests were performed using the fasted state HIF (pH 6.24) for<br />

13 model drugs <strong>and</strong> the results were compared to the HTS solubility results in the aqueous<br />

buffer solution at pH 6.8 (Figure 12) – resulting in a fair correlation (R 2 =0.630). More<br />

importantly, the absolute values <strong>of</strong> compound solubility differed only modestly (less than<br />

6.9-fold in all cases), even though 1000-fold range <strong>of</strong> solubilities were studied. Slightly<br />

increased solubility values in HIF compared to the buffer solution pH 6.8 were seen for<br />

piroxicam, naproxen, rifampicin <strong>and</strong> dicl<strong>of</strong>enac sodium. Vice versa, the solubilities in HIF


compared to pH 6.8 buffer solution were decreased for ibupr<strong>of</strong>en <strong>and</strong> ketopr<strong>of</strong>en. No<br />

major differences were seen between the solubility in buffer solution pH 6.8 <strong>and</strong> HIF<br />

suggesting that buffer is an adequate medium when studying dissolution <strong>and</strong> solubility<br />

with compounds in the 100 μg/ml–100mg/ml solubility range. However, we can not rule<br />

out the possibility that more significant solubility differences would be seen for the<br />

compounds with solubility below 100 μg/ml. For example, preliminary solubility test <strong>of</strong><br />

beclomethasone dipropionate, showed 150 times higher solubility in HIF (0.6 μg/ml) than<br />

in the buffer at pH 6.8 (0.004 μg/ml). Obviously, more detailed studies are needed to<br />

determine the limits <strong>and</strong> utilization potential <strong>of</strong> HIF in drug solubility <strong>and</strong> dissolution<br />

testing, especially in the case <strong>of</strong> the compounds with very low solubility. It is worth<br />

noting, however, that the simple buffer is useful medium over surprisingly wide solubility<br />

range that extends at least to 10 2 μg/ml.<br />

HIF was, found to be a suitable solubility/dissolution medium for the 96-well plate<br />

testings. The small volumes make it possible to utilize the HIF also as a medium in drug<br />

discovery process, which may have an even more pronounced importance for very low<br />

solubility compounds. HIF is the most relevant solubility medium, <strong>and</strong> in this system it<br />

can be used instead <strong>of</strong> the buffers or simulated intestinal fluids, if needed.<br />

Figure 12 The solubilities in HIF (y-axis) <strong>and</strong> buffer solution pH 6.8 (x-axis). Modified from<br />

paper IV.<br />

51


6 Conclusions<br />

The main conclusions drawn based on the results <strong>of</strong> this study are:<br />

1. Channel flow method was introduced to the dissolution testing. The method is<br />

applicable on drug dissolution testing <strong>and</strong> enables especially kinetic follow-up<br />

at the very beginning <strong>of</strong> the dissolution process. The results from the channel<br />

flow method <strong>and</strong> the intrinsic dissolution method correlated well with each<br />

other. Later the channel flow method has been further developed to study the<br />

solvent-mediated solid-state transformation during dissolution (Aaltonen et al.,<br />

2006; Lehto et al., 2008).<br />

2. A micro dissolution method was used to determine equilibrium solubilities <strong>of</strong><br />

model compounds with smaller volumes than by the shake-flask method. The<br />

solubility results from the micro dissolution were reliable indicating utilization<br />

potential <strong>of</strong> the micro dissolution method instead <strong>of</strong> shake-flask method.<br />

<strong>Miniaturization</strong> was successful to the 96-well plates.<br />

3. A 96-well plate method was developed for measuring the equilibrium drug<br />

solubilities <strong>and</strong> dissolution pr<strong>of</strong>iles as a function <strong>of</strong> time in microlitre scale.<br />

The method is useful for screening purposes at early stages <strong>of</strong> drug<br />

development. Also polymorphic changes <strong>of</strong> compounds can be studied. The<br />

methodology developed is the first true equilibrium solubility measuring HTS<br />

system utilizing 96-well plates, meaning that in the beginning the compound<br />

was in solid form.<br />

4. A novel surface tension based microtensiometric method was presented for<br />

kinetic HTS <strong>of</strong> drug solubility properties. The method was tested by<br />

determining physiologically relevant parameters on the solubility <strong>of</strong> model<br />

compounds. The method is well suited for classifying purposes <strong>and</strong> solubility<br />

screening tests.<br />

5. HIF from duodenum was used as a medium for the first time in 96-well plate<br />

equilibrium solubility testings. Microliter volumes make it possible to utilize<br />

HIF, the most relevant solubility medium, in dissolution testing at drug<br />

discovery stage. This may be even more important in the future with very low<br />

solubility compounds. Still, the aqueous buffer solutions are the first choice<br />

due to good prediction capacity, low costs, ease <strong>of</strong> access <strong>and</strong> good correlation<br />

with model drugs.<br />

52


7 References<br />

Aaltonen, J., Heinänen, P., Peltonen, L., Kortejärvi, H., Tanninen, V.P., Christiansen, L.,<br />

Hirvonen, J., Yliruusi, J., Rantanen, J., 2006. In situ measurement <strong>of</strong> solvent-mediated<br />

phase transformation during dissolution testing. J Pharm Sci 95: 2730-2737.<br />

Aaltonen, J., Rades, T., 2009. Towards physico-relevant dissolution testing: The<br />

importance <strong>of</strong> solid-state analysis in dissolution. Dissol Tech 16: 47-54.<br />

Achanta, A.S., Gray, V.A., Cecil, T.L., Grady, L.T., 1995. Evaluation <strong>of</strong> the performance<br />

<strong>of</strong> prednisone <strong>and</strong> salicylic acid USP dissolution calibrators. <strong>Drug</strong> Dev Ind Pharm 21:<br />

1171-1182.<br />

Alelyunas, Y.W., Liu, R., Pelosi-Kilby, L., Shen, C., 2009. Application <strong>of</strong> a dried-DMSO<br />

rapid throughput 24-h equilibrium solubility in advancing discovery c<strong>and</strong>idates. Eur J<br />

Pharm Sci 37: 172-182.<br />

Allen, F.H., 2002. The Cambridge Structural Database: a quarter <strong>of</strong> a million crystal<br />

structures <strong>and</strong> rising. Acta Cryst B58: 380-388.<br />

Alsenz, J., Kansy, M., 2007. High throughput solubility measurement in drug discovery<br />

<strong>and</strong> development. Adv <strong>Drug</strong> Del Rev 59: 546-567.<br />

Alsenz, J., Meister, E., Haenel, E., 2007. Development <strong>of</strong> a partially automated solubility<br />

screening (PASS) assay for early drug development. J Pharm Sci 96: 1748-1762.<br />

Amidon, G.L., Lennernäs, H., Shah, V.P., Crison, J.R., 1995. A theoretical basis for a<br />

biopharmaceutic drug classification: the correlation <strong>of</strong> in vitro drug product dissolution<br />

<strong>and</strong> in vivo bioavailability. Pharm Res 12: 413-420.<br />

Arm<strong>and</strong>, M., Borel, P., Pasquier, B., Dubois, C., Senft, M., Andre, M., Peyrot, J.,<br />

Salducci, J., Lairon, D., 1996. Physicochemical characteristics <strong>of</strong> emulsions during fat<br />

digestion in human stomach <strong>and</strong> duodenum. Am J Physiol 271: G171-G183.<br />

Artursson, P., Palm, K., Luthman, K., 2001. Caco-2 monolayers in experimental <strong>and</strong><br />

theoretical predictions <strong>of</strong> drug transport. Adv <strong>Drug</strong> Deliv Rev 46: 27-43.<br />

Ashford, M., 2007. Gastrointestinal tract – physiology <strong>and</strong> drug absorption. In: Aulton’s<br />

pharmaceutics. The design <strong>and</strong> manufacture <strong>of</strong> medicines, Aulton, M.E. (Ed),<br />

Churchill Livingstone Elsevier, pp. 270-285.<br />

Avdeef, A., 2007. <strong>Solubility</strong> <strong>of</strong> sparingly-soluble ionisable drugs. Adv <strong>Drug</strong> Del Rev 59:<br />

568-590.<br />

Avdeef, A., Berger, C.M., Brownell, C., 2000. pH-metric solubility. 2: Correlation<br />

between the acid-base titration <strong>and</strong> the saturation shake-flask solubility-pH methods.<br />

Pharm Res 17: 85-89.<br />

Back, D.J., Rogers, S.M., 1987. Review: first-pass metabolism by the gastrointestinal<br />

mucosa. Aliment Pharmacol Ther 1: 339-357.<br />

Bakatselou, V., Oppenheim, R.C., Dressman, J.B., 1991. Solubilization <strong>and</strong> wetting<br />

effects <strong>of</strong> bile salts on the dissolution <strong>of</strong> steroids. Pharm Res 8: 1461-1469.<br />

Balakin, K.V., Ivnenkov, Y.A., Skorenko, A.V., Nikolsky, Y.V., Savchuk, N.P.,<br />

Ivashchenko, A.A., 2004. In silico estimation <strong>of</strong> DMSO solubility <strong>of</strong> organic<br />

compounds for bioscreening. J Biomol Screen 9:22-31.<br />

Banaker, U.V., 1992. Pharmaceutical dissolution testing. Marcel Dekker Inc., New York.<br />

Bard, B., Martel, S., Carrupt, P.-A., 2008. High throughput UV method for the estimation<br />

<strong>of</strong> thermodynamic solubility <strong>and</strong> the determination <strong>of</strong> the solubility in biorelevant<br />

media. Eur J Pharm Sci 33: 230-40.<br />

Bauer, J., Spanton, S., Henry, R., Quick, J., Dziki, W., Porter, W., Morris, J., 2001.<br />

Ritonavir: An extraordinary example <strong>of</strong> conformational polymorphism. Pharm Res 18:<br />

859-866.<br />

53


Beckett, A.H., Quach, T.T., Kurs, G.S., 1996. Improved hydrodynamics for USP<br />

apparatus 2. Dissol Tech 3: 7-18.<br />

Beckstead, H.D., Neville, G.A., Shurvell, H.F., 1993. Differentiation <strong>of</strong> solvated<br />

spironolactone samples by FT-raman <strong>and</strong> FT-IR diffuse reflectance spectroscopy.<br />

Fresenius J Anal Chem 345: 727-732.<br />

Benet, L.Z., Amidon, G.L., Barends, D.M., Lennernäs, H., Polli, J.E., Shah, V.P.,<br />

Stavchansky, S.A., Yu, L.X., 2008. The use <strong>of</strong> BDDCS in classifying the permeability<br />

<strong>of</strong> marketed drugs. Pharm Res 52: 483-488.<br />

Berger, C.M., Tsinman, O., Voloboy, D., Lipp, D., Stones, S., Avdeef, A., 2007.<br />

Technical note: Miniaturized intrinsic dissolution rate (Mini-IDR TM ) measurement <strong>of</strong><br />

grise<strong>of</strong>ulvin <strong>and</strong> carbamazepine. Dissol Tech 14: 39-41.<br />

Bergström, C.A.S., Norinder, U., Luthman, K., Artursson, P., 2002. Experimental <strong>and</strong><br />

computational screening models for prediction <strong>of</strong> aqueous drug solubility. Pharm Res<br />

19: 182-188.<br />

Bevan, C.D., Lloyd, R.S., 2000. A high-throughput screening method for the<br />

determination <strong>of</strong> aqueous drug solubility using laser nephelometry in microtiter plates.<br />

Anal Chem 72: 1781-87.<br />

Blanquet, S., Zeijdner, E., Beyssac, E., Meunier, J.-P., Denis, S., Havenaar, R., Alric, M.,<br />

2004. A dynamic artificial gastrointestinal system for studying the behavior <strong>of</strong> orally<br />

administered drug dosage forms under various physiological conditions. Pharm Res<br />

21: 585-591.<br />

Blasko, A., Leahy-Dios, A, Nelson, W.O., Austin, S.A., Killion, R.B., Visor, G.C.,<br />

Massey, I.J., 2001. Revisiting the solubility concept <strong>of</strong> pharmaceutical compounds.<br />

Monatsh Chem 132: 789-798.<br />

Bordner, J., 1984. Piroxicam monohydrate: a zwitterionic form, C15H13N3O4S.H2O. Acta<br />

Cryst C40: 989-990.<br />

Brewster, M.E., L<strong>of</strong>tsson, T., 2007. Cyclodextrins as pharmaceutical solubilizers. Adv<br />

<strong>Drug</strong> Deliv Rev 59: 645-666.<br />

Brouwers, J., Tack, J., Lammert, F., Augustjins, P., 2006. Intraluminal drug <strong>and</strong><br />

formulation behaviour <strong>and</strong> integration in in vitro permeability estimation: a case study<br />

with amprenavir. J Pharm Sci 95: 372-383.<br />

Brown, C.A., Compton, R.G., Narramore, C.A., 1993. The kinetics <strong>of</strong> calcite<br />

dissolution/precipitation. J Colloid Interface Sci 160: 372-379.<br />

Cakiryildiz, C., Mehta, P.J., Rahmen, W., Schoenleber, D., 1975. <strong>Dissolution</strong> studies with<br />

a multichannel continuous-flow apparatus. J Pharm Sci 64: 1692-1697.<br />

Castellari, C., Ottani, S., 1997. Two monoclinic forms <strong>of</strong> dicl<strong>of</strong>enac acid. Acta Cryst C53:<br />

794-797.<br />

Charman, W.N., Rogge, M.C., Boddy, A.W., Berger, B.M., 1993. Effect <strong>of</strong> food <strong>and</strong> a<br />

monoglyceride emulsion formulation on danazol bioavailability. J Clin Pharmacol 33:<br />

381-387.<br />

Chen, X., Morris, K.R., Griesser, U.J., Byrn, S.R., Stowell, J.G., 2002a. Reactivity<br />

differences <strong>of</strong> indomethacin solid forms with ammonia gas. J Am Chem Soc 124:<br />

15012-15019.<br />

Chen, T.-M., Shen, H., Zhu, C., 2002b. Evaluation <strong>of</strong> a method for high throughput<br />

solubility determination using a multi-wavelength UV plate reader. Comb Chem High<br />

Throughput Screen 5: 575-81.<br />

Chen, X.-Q., Venkatesh, S., 2004. Miniature device for aqueous <strong>and</strong> non-aqueous<br />

solubility measurements during drug discovery. Pharm Res 21: 1758-61.<br />

Cheng, C.-L., Yu, L.X., Lee, H.-L., Yang, C.-Y., Lue, C.-S., Chou, C.-H., 2004.<br />

Biowaiver extension potential to BCS class II high solubility-low permeability drugs:<br />

54


idging evidence for metformin immediate-release tablet. Eur J Pharm Sci 22: 297-<br />

304.<br />

Clarysse, S., Psachoulias, D., Brouwers, J., Tack, J., Annaert, P., Duchateau, G., Reppas,<br />

C., Augustijns, P., 2009a. Postpr<strong>and</strong>ial changes in solubilizing capacity <strong>of</strong> human<br />

intestinal fluids for BCS class II drugs. Pharm Res 26: 1456-1466.<br />

Clarysse, S., Tack, J., Lammert, F., Duchateau, G., Reppas, C., Augustijns, P., 2009b.<br />

Postpr<strong>and</strong>ial evolution in composition <strong>and</strong> characteristics <strong>of</strong> human duodenal fluids in<br />

different nutritional states. J Pharm Sci 98: 1177-1192.<br />

Compton, R.G., Pilkington, M.B.G., Stearn, G.M., 1988. Mass transport in channel<br />

electrodes: the application <strong>of</strong> the backward implicit method to electrode reactions (EC,<br />

ECE <strong>and</strong> DISP) involving coupled homogenous kinetics. J Chem Soc Faraday Trans 1:<br />

2155-2171.<br />

Compton, R.G., Harding, M.S., Pluck, M.R., Atherton, J.H., Brennan, C.M., 1993.<br />

Mechanisms <strong>of</strong> solid/liquid interfacial reactions. The dissolution <strong>of</strong> benzoic acid in<br />

aqueous solution. J Phys Chem 97: 10416-10420.<br />

Dai, W.-G., Pollock-Dove, C., Dong, L.C., Li, S., 2008. Advanced screening assays to<br />

rapidly identify solubility-enhancing formulations: high-throughput, miniaturization<br />

<strong>and</strong> automation. Adv <strong>Drug</strong> Deliv Rev 60: 657-72.<br />

Davis, W.W., Parke, T.V., 1942. A nephelometric method for determination <strong>of</strong> solubilities<br />

<strong>of</strong> extremely low order. J Am Chem Soc 64: 101-107.<br />

Davis, J.M., Rowley, S.D., Braine, H.G., Piantadosi, S., Santos, G.W., 1990. Clinical<br />

toxicity <strong>of</strong> cryopreserved bone marrow graft infusion. Blood 75: 781-786.<br />

Dehring, K.A., Workman, H.L., Miller, K.D., M<strong>and</strong>agere, A., Pole, S.K., 2004.<br />

Automated robotic liquid h<strong>and</strong>ling/laser-based nephelometry system for high<br />

throughput measurement <strong>of</strong> kinetic aqueous solubility. J Pharm Biomed Anal 36: 447-<br />

56.<br />

Di, L., Kerns, E.H., 2006. Biological assay challenges from compound solubility:<br />

strategies for bioassay optimization. <strong>Drug</strong> Discovery Today 11:446-451.<br />

Dideborg, O., Dupont, L., 1972. La structure crystalline et moleculaire de la<br />

spironolactone (7α-acetylthio-3-oxo-17α-4-pregnene-21,17βcarbolactone). Acta Cryst<br />

B28: 3014-3022.<br />

Dokoumetzidis, A., Macheras, P., 2006. A century <strong>of</strong> dissolution research: From Noyes<br />

<strong>and</strong> Whitney to the biopharmaceutics classification system. Int J Pharm 321: 1-11.<br />

Dressman, J.B., Berardi, R.R., Dermentzoglou, L.C., Russell, T.L., Schmaltz, S.P.,<br />

Barnett, J.L., Jarvenpaa, K.M., 1990. Upper gastrointestinal (GI) pH in young, healthy<br />

men <strong>and</strong> women. Pharm Res 7: 756-761.<br />

Dressman, J.B., Amidon, G.L., Reppas, C., Shah, V., 1998. <strong>Dissolution</strong> testing as a<br />

prognostic tool for oral drug absorption: Immediate release dosage forms. Pharm Res<br />

15: 11-22.<br />

El-dalsh, S.S., El-Sayed, A.A., Badawi, A.A., 1983. Studies on spironolactone<br />

polymorphic forms. <strong>Drug</strong> Dev Ind Pharm 9: 877-894.<br />

EMEA, 2002. Note for Guidance on the Investigation <strong>of</strong> Bioavailability <strong>and</strong><br />

Bioequivalence. Committee for Medicinal Products for Human Use (CHMP).<br />

EMEA, 2008. Annex 7 to note for evaluation <strong>and</strong> recommendation <strong>of</strong> pharmacopoeial<br />

texts for use in the ICH regions on dissolution test general chapter. Committee for<br />

Medicinal Products for Human Use (CHMP).<br />

European Pharmacopoeia, 2009, 6 th edition, Druckerei C.H. Beck, Nördlingen.<br />

Fallingborg, J., Christensen, L.A., Ingeman-Nielsen, M., Jacobsen, B.A., Abildgaard, K.,<br />

Rasmussen, H.H., 1989. pH-pr<strong>of</strong>ile <strong>and</strong> regional transit times <strong>of</strong> the normal gut<br />

measured by a radiotelemetry device. Aliment Pharmacol Therap 3: 605-613.<br />

55


FDA, 1997a. Guidance for industry: <strong>Dissolution</strong> testing <strong>of</strong> immediate release solid oral<br />

dosage forms, retrieved November 2009, from http://www.fda.gov/downloads<br />

/<strong>Drug</strong>s/GuidanceComplianceRegulatoryInformation/Guidances/ucm070237.pdf<br />

FDA, 1997b. Guidance for industry: Extended release oral dosage forms: development,<br />

evaluation <strong>and</strong> application <strong>of</strong> in vitro/in vivo correlations, retrieved November 2009,<br />

from http://www.fda.gov/downloads/<strong>Drug</strong>s/GuidanceComplianceRegulatoryInformati<br />

on/Guidances/UCM070239.pdf<br />

FDA, 2000. Guidance for industry: Waiver <strong>of</strong> in vivo bioavailability <strong>and</strong> bioequivalence<br />

studies for immediate release solid oral dosage forms based on a biopharmaceutics<br />

classification system, retrieved November 2009, from http://www.fda.gov/ downloads<br />

/<strong>Drug</strong>s/GuidanceComplianceRegulatoryInformation/Guidances/ucm070246.pdf<br />

Fini, A., Fazio, G., Fern<strong>and</strong>ez-Hervas, M.J., Holgado, M.A., Rabasco, A.M., 1995.<br />

Influence <strong>of</strong> crystallization solvent <strong>and</strong> dissolution behaviour for a dicl<strong>of</strong>enac salt. Int J<br />

Pharm Sci 121: 1926.<br />

Fleischer, D., Li, C., Zhou, Y., Pao, L.-H., Karim, A., 1999. <strong>Drug</strong>, meal <strong>and</strong> formulation<br />

interactions influencing drug absorption after oral administration. Clin Pharmacokinet<br />

36: 233-254.<br />

Fligge, T.A., Schuler, A., 2006. Integration <strong>of</strong> a rapid automated solubility classification<br />

into early validation <strong>of</strong> hits obtained by high throughput screening. J Pharm Biomed<br />

Anal 42: 449-54.<br />

Galia, E., Nicolaides, E., Hörter, D., Löbenberg, R., Reppas, C., Dressman, J.B., 1998.<br />

Evaluation <strong>of</strong> various dissolution media for predicting in vivo performance <strong>of</strong> class I<br />

<strong>and</strong> II drugs. Pharm Res 15: 698-705.<br />

Garbacz, G., Wedemeyer, R.-S., Nagel, S., Giessmann, T., Mönnikes, H., Wilson, C.G.,<br />

Siegmund, W., Weitschies, W., 2008. Irregular absorption pr<strong>of</strong>iles observed from<br />

dicl<strong>of</strong>enac extended release tablets can be predicted using a dissolution test apparatus<br />

that mimics in vivo physical stresses. Eur J Pharm Sci 70: 421-423.<br />

Garbacz, G., Blume, H., Weitschies, W., 2009. Investigation <strong>of</strong> the dissolution<br />

characteristics <strong>of</strong> nifedipine extended-release formulations using USP apparatus 2 <strong>and</strong><br />

a novel dissolution apparatus. Dissol Tech 16: 7-13.<br />

Glomme, A., März, J., Dressman, J.B., 2005. Comparison <strong>of</strong> a miniaturized shake-flask<br />

solubility method with automated potentiometric acid/base titrations <strong>and</strong> calculated<br />

solubilities. J Pharm Sci 94: 1-16.<br />

Gu, L., Huynh, O., Becker, A., Peters, S., Nguyen, H., Chu, N., 1987. Preformulation<br />

selection <strong>of</strong> a proper salt for a weak acid-base (RS-82856) – a new positive inotropic<br />

acid. <strong>Drug</strong> Dev Ind Pharm 13: 437-448.<br />

Gurantz, D., Laker, M.F., H<strong>of</strong>fman, A.F., 1981. Enzymatic measurement <strong>of</strong> cholinecontaining<br />

phospholipids in bile. J Lipid Res 22: 373-376.<br />

Hauss, D.J., 2007. Oral lipid-based formulations. Adv <strong>Drug</strong> Deliv Rev 59: 667-676.<br />

Havenaar, R., Menikus, M., 1994. In vitro model <strong>of</strong> an in vivo digestive tract. JP, US,<br />

European Patent PCT/NL, 93/00225.<br />

Higuchi, T., 1961. Rate <strong>of</strong> release <strong>of</strong> medicaments from ointment bases containing drugs<br />

in suspension. J Pharm Sci 50: 874-875.<br />

Hill, J.W., Petrucci, R.H., 1999. General Chemistry, 2 nd edition, Prentice Hall, New<br />

Jersey.<br />

Ho, N.F.H., Merkle, H.P., Higuchi, W.I., 1983. Quantitative, mechanistic <strong>and</strong><br />

physiologically realistic approach to the biopharmaceutical design <strong>of</strong> oral drug<br />

delivery systems. <strong>Drug</strong> Dev Ind Pharm 9: 1111-1184.<br />

Hoelke, B., Gieringer, S., Arlt, M., Saal, C., 2009. Comparison <strong>of</strong> nephelometric, UVspectroscopic,<br />

<strong>and</strong> HPLC methods for high-throughput determination <strong>of</strong> aqueous drug<br />

solubility in microtiter plates. Anal Chem 81: 3165-3172.<br />

56


Homon, C.A., Nelson R.M., 2006. High-throughput screening: Enabling <strong>and</strong> influencing<br />

the process <strong>of</strong> drug discovery. In: The process <strong>of</strong> new drug discovery <strong>and</strong><br />

development, Smith, C.G., O’Donnell, J.T. (Eds.), 2 nd edition, Informa healthcare,<br />

New York, pp. 79-102.<br />

Int.Ph. 2009, 4 th edition, retrieved November 2009 from http://apps.who.int/phint/en/p<br />

/docf<br />

Jaiboon, N., Yos-In, K., Ruangchaithaweesuk, S., Chaichit, N., Thutivoranath, R.,<br />

Siritaedmukul, K., Hannongbua, S., 2001. New orthorhombic form <strong>of</strong> 2-((2,6dichlorophenyl)amino)<br />

benzeneacetic acid (dicl<strong>of</strong>enac acid). Anal Sci 17: 1465-1466.<br />

Jantratid, E., Janssen N., Reppas, C., Dressman, J., 2008. <strong>Dissolution</strong> media simulating<br />

conditions in the proximal human gastrointestinal tract: An update. Pharm Res 25:<br />

1663-1676.<br />

Japanese Pharmacopoeia 2009, 15 th edition, retrieved November 2009 from<br />

http://jpdb.nihs.go.jp/jp15e/<br />

Johnson, K.C., Swindell, A.C., 1996. Guidance in the setting <strong>of</strong> drug particle size<br />

specifications to minimize variability in absorption. Pharm Res 13: 1795-1798.<br />

Kalantzi, L., Goumas, K., Kalioras, V., Abrahamsson, B., Dressman, JR., Reppas, C.,<br />

2006a. Characterization <strong>of</strong> the human upper gastrointestinal contents under conditions<br />

simulating bioavailability/bioequivalence studies. Pharm Res 23: 165-176.<br />

Kalantzi, L., Persson, E., Polentarutti, B., Abrahamsson, B., Goumas, K., Dressman, J.,<br />

Reppas, C., 2006b. Canine intestinal contents vs. simulated media for the assessment<br />

<strong>of</strong> solubility <strong>of</strong> two weak bases in the human small intestinal contents. Pharm Res 23:<br />

1373-1381.<br />

Kawakami, K., 2009. Current status <strong>of</strong> amorphous formulation <strong>and</strong> other special dosage<br />

forms as formulations for early clinical phases. J Pharm Sci 98: 2875-2885.<br />

Kerns, E.H., 2001. High throughput physicochemical pr<strong>of</strong>iling for drug discovery. J<br />

Pharm Sci 90: 1838-1858.<br />

Kibbey, C.E., Poole, S.K., Robinson, B., Jackson, J.D., Durham, D., 2001. An integrated<br />

process for measuring the physicochemical properties <strong>of</strong> drug c<strong>and</strong>idates in a<br />

preclinical discovery environment. J Pharm Sci 90: 1164-1175.<br />

Kistenmacher, T.J., Marsh, R.E., 1972. Crystal <strong>and</strong> molecular structure <strong>of</strong> an<br />

antiinflammatory agent, indomethacin, 1-(p-chlorobenzoyl)-5-methoxy-2methylindole-3-acetic<br />

acid. J Am Chem Soc 94: 1340-1345.<br />

Kojic-Prodic, B., Ruzic-Toros, Z., 1982. Structure <strong>of</strong> the anti-inflammatory drug 4hydroxy-2-methyl-N-2-pyridyl-2H-1λ<br />

6 ,2-benzothiazine-3-carboxamide 1,1-dioxide<br />

(piroxicam). Acta Cryst B38: 2948-2951.<br />

Kortejärvi, H., Urtti, A., Yliperttula, M., 2007. Pharmacokinetic simulation <strong>of</strong> biowaiver<br />

criteria: The effects <strong>of</strong> gastric emptying, dissolution, absorption <strong>and</strong> elimination rates.<br />

Eur J Pharm Sci 30: 155-166.<br />

Kortejärvi, H., Shawahna, R., Koski, A., Malkki, J., Ojala, K., Yliperttula, M., 2010. Very<br />

rapid dissolution is not needed to guarantee bioequivalence for biopharmaceutics<br />

classification system (BCS) I drugs. J Pharm Sci 99: 621-625.<br />

Kortejärvi, H., Yliperttula, M., Dressman, J.B., Junginger, H.E., Midha, K.K., Shah, V.P.,<br />

Barends, D.M., 2005. Biowaiver monographs for immediate release solid oral dosage<br />

forms: ranitidine hydrochloride. J Pharm Sci 94: 1617-1625.<br />

Kramer, C., Heinisch, T., Fligge, T., Beck, B., Clark, A., 2009. A consistant dataset <strong>of</strong><br />

kinetic solubilities for early-phase drug discovery. J Med Chem 4: 1529-1536.<br />

Lehto, P., Aaltonen, J., Niemelä, P., Rantanen, J., Hirvonen, J., Tanninen, V.P., Peltonen,<br />

L., 2008. Simultaneous measurement <strong>of</strong> liquid-phase <strong>and</strong> solid-phase transformation<br />

kinetics in rotating disc <strong>and</strong> channel flow cell dissolution devices. Int J Pharm 363: 66-<br />

72.<br />

57


Lindahl, A., Ungell, A.-L., Knutson, L., Lennernäs L., 1997. Characterization <strong>of</strong> fluids<br />

from the stomach <strong>and</strong> proximal jejunum in men <strong>and</strong> women. Pharm Res 14: 497-502.<br />

Lindenberg, M., Wieg<strong>and</strong>, C., Dressman, J., 2005. Comparison <strong>of</strong> the adsorption <strong>of</strong><br />

several drugs to typical filter materials. Dissol Tech 12: 22-25.<br />

Lipinski, C.A., 2000. <strong>Drug</strong>-like properties <strong>and</strong> the causes <strong>of</strong> poor solubility <strong>and</strong> poor<br />

permeability. J Pharmacol Toxicol Methods 44: 235-249.<br />

Lipinski, C., 2003. Aqueous solubility in discovery, chemistry, <strong>and</strong> assay changes. In:<br />

<strong>Drug</strong> bioavailability; estimation <strong>of</strong> solubility, permeability, absorption <strong>and</strong><br />

bioavailability, van de Waterbeemd, H., Lennernäs, H., Artursson, P. (Eds.), Wiley-<br />

VCH Verlag GmbH, Weinheim, pp. 215-231.<br />

Lipinski, C.A., 2004. <strong>Solubility</strong> in water <strong>and</strong> DMSO: Issues <strong>and</strong> potential solutions. In:<br />

Biotechnology: Pharmaceutical aspects, vol. 1, Borchardt, R.T., Kerns, E.H., Lipinski,<br />

C.A., Thakler, D.R., Wang, B. (Eds.), American association <strong>of</strong> pharmaceutical<br />

sciences, Arlington VA, pp. 93-125.<br />

Lipinski, C.A., Lombardo, F., Dominy, B.W., Feeney, P.J., 1997. Experimental <strong>and</strong><br />

computational approaches to estimate solubility <strong>and</strong> permeability in drug discovery<br />

<strong>and</strong> development settings. Adv <strong>Drug</strong> Deliv Rev 23: 3-25.<br />

Llinas, A., Burley, J.C., Box, K.J., Glen, R.C., Goodman J.M., 2007. Dicl<strong>of</strong>enac<br />

solubility: Independent determination <strong>of</strong> the intrinsic solubility <strong>of</strong> three crystal forms. J<br />

Med Chem 50: 979-983.<br />

L<strong>of</strong>tsson, T., Hreinsdottir, D., 2006. Determination <strong>of</strong> aqueous solubility by heating <strong>and</strong><br />

equilibration. AAPS Pharm Sci Tech 7: E1-E4.<br />

Lorimer, J.W., Cohen-Adad, R., 2003. Thermodynamics <strong>of</strong> solubility. In: The<br />

experimental determination <strong>of</strong> solubilities, Hefter, G.T., Tomkins, R.P.T. (Eds.),<br />

volume 6, John Wiley <strong>and</strong> Sons Ltd, Chichester, pp. 17-76.<br />

Lütjohann, D., Hahn, C., Prange, W., Sudhop, T., Axelson, M., Sauerbruch, T., von<br />

Bergmann, K., Reichel, C., 2004. Influence <strong>of</strong> rifampin on serum markers <strong>of</strong><br />

cholesterol <strong>and</strong> bile acid synthesis in men. Int J Clin Pharmacol Ther 42: 307-313.<br />

Mansbach, M.C., Cohen, R.S., Leff, P.B., 1975. Isolation <strong>and</strong> properties <strong>of</strong> the mixed lipid<br />

micelles present in intestinal content during fat digestion in man. J Clin Invest 56: 781-<br />

791.<br />

Martinez, M.N., Amidon, G.L., 2002. A mechanistic approach to underst<strong>and</strong>ing the<br />

factors affecting drug absorption: A review <strong>of</strong> fundamentals. J Clin Pharmacol 42:<br />

620-643.<br />

Mathus-Vliegen, E.M.H., van Ierl<strong>and</strong>-van Leeuwen, M.L., Terpstra, A., 2004. Lipase<br />

inhibition by orlistat: effects on gall-bladder kinetics <strong>and</strong> cholecystokinin release<br />

obesity. Aliment Pharmacol Ther 19: 601-611.<br />

Mauger, J.W., 1996. Physicochemical <strong>and</strong> fluid mechanical factors related to dissolution<br />

testing. Dissol Tech 3: 7-11.<br />

Mauger, J., Ballard, J., Brockson, R., De, S., Gray, V., Robinson, D., 2003. Intrinsic<br />

dissolution performance testing <strong>of</strong> the USP dissolution apparatus 2 (rotating paddle)<br />

using modified salicylic acid calibrator tablets: pro<strong>of</strong> <strong>of</strong> principle. Dissol Tech 10: 6-<br />

15.<br />

Mithani, S., Bakatselou, V., TenHoor, C.N., Dressman, J.B., 1996. Estimation <strong>of</strong> the<br />

increase in solubility <strong>of</strong> drugs as a function <strong>of</strong> bile salt concentration. Pharm Res 13:<br />

163-167.<br />

Moreno de la Cruz, M.P., Oth, M., Deferme, S., Lammert, F., Tack, J., Dressman, J.,<br />

Augustijns, P., 2006. Characterization <strong>of</strong> fasted-state human intestinal fluids collected<br />

from duodenum <strong>and</strong> jejunum. J Pharm Pharmacol 58: 1079-1089.<br />

Muangsin, N., Prajuabsook, M., Chimsook, P., Chantarasiri, N., Siraleartmukul, K.,<br />

Chaichit, N., Hannongbua, S., 2004. Structure determination <strong>of</strong> dicl<strong>of</strong>enac in a<br />

58


dicl<strong>of</strong>enac-containing chitosan matrix using conventional X-ray powder diffraction<br />

data. J Appl Crystallogr 37: 288-294.<br />

Nicklasson, M., Brodin, A., 1984. The relationship between intrinsic dissolution rates <strong>and</strong><br />

solubilities in the water-ethanol binary solvent system. Int J Pharm 18: 149-156.<br />

Nicklasson, M., Brodin, A., Nyqvist, H., 1981. Studies on the relationship between<br />

solubility <strong>and</strong> intrinsic rate <strong>of</strong> dissolution as a function <strong>of</strong> pH. Acta Pharm Suec 18:<br />

119-128.<br />

Nicklasson, M., Fyhr, P., Magnusson, A.-B., Gunnvald, K. A., 1988. Preformulation study<br />

on the in vitro dissolution characteristics <strong>of</strong> the organophosphorus poisoning antidote<br />

HI-6. Int J Pharm 46: 247-254.<br />

Noyes, A.A., Whitney, W.R., 1897. The rate <strong>of</strong> solution <strong>of</strong> solid substances in their own<br />

solutions. J Am Chem Soc 19: 930-934.<br />

O’Donnell, J.R., Burnett, A.K., Sheehan, T., Tansey, P., McDonald, G.A., 1981. Safety <strong>of</strong><br />

dimethylsulphoxide. Lancet 28: 498.<br />

Omelczuk, M.O., McGinity, J.W., 1993. The influence <strong>of</strong> thermal treatment on the<br />

physical-mechanical <strong>and</strong> dissolution properties <strong>of</strong> tablets containing poly(DL-lactide<br />

acid). Pharm Res 10: 542-548.<br />

Oprea, T.I., Davis, A.M., Teague, S.J., Leeson, P.D., 2001. Is there a difference between<br />

leads <strong>and</strong> drugs? A historical perspective. J Chem Inf Comput Sci 41: 1308-1315.<br />

Pan, L., Ho, Q., Tsutsui, K., Takahashi, L., 2001. Comparison <strong>of</strong> chromatographic <strong>and</strong><br />

spectroscopic methods used to rank compounds for aqueous solubility. J Pharm Sci 90:<br />

521-29.<br />

Pernarowski, M., Woo, W., Searl, R.O., 1968. Continuous flow apparatus for the<br />

determination <strong>of</strong> the dissolution characteristics <strong>of</strong> tablets <strong>and</strong> capsules. J Pharm Sci 57:<br />

1419-1421.<br />

Persson, E.M., Gustafsson, A., Carlsson, A.S., Nilsson, R.G., Knutson, L., Forsell, P.,<br />

Hanisch, G., Lennernäs, H., Abrahamsson, B., 2005. The effects <strong>of</strong> food on the<br />

dissolution <strong>of</strong> poorly soluble drugs in human <strong>and</strong> in model small intestinal fluids.<br />

Pharm Res 22: 2141-2151.<br />

Persson, A.M., Baumann, K., Sundelöf, L.-O., Lindberg, W., Sokolowski, A., Pettersson,<br />

C., 2008. Design <strong>and</strong> characterization <strong>of</strong> a new miniaturized rotating disk equipment<br />

for in vitro dissolution rate studies. J Pharm Sci 97: 3344-3355.<br />

Porter, C.J.H., Charman, W.N., 2001. Lipid-based formulation for oral administration:<br />

opportunities for bioavailability enhancement <strong>and</strong> lipoprotein targeting <strong>of</strong> lipophilic<br />

drugs. In <strong>Drug</strong> targeting technology: physical, chemical, biological methods, vol. 115,<br />

Schreier, H., (Eds.), Marcel Dekker Inc., New York, pp. 85-130.<br />

Pudipeddi, M., Serajuddin, A.T.M., 2005. Trends in solubility <strong>of</strong> polymorphs. J Pharm Sci<br />

94: 929-939.<br />

Qureshi, S.A., McGilveray, I.J., 1999. Typical variability in drug dissolution testing: study<br />

with USP <strong>and</strong> FDA calibrator tablets <strong>and</strong> a marketed drug (glibenclamide) product.<br />

Eur J Pharm Sci 7: 249-258.<br />

Qureshi, S.A., Shabnam, J., 2003. Applications <strong>of</strong> new device (spindle) for improved<br />

characterization <strong>of</strong> drug release (dissolution) <strong>of</strong> pharmaceutical products. Eur J Pharm<br />

Sci 19: 291-297.<br />

Rapoport, A.P., Rowe, J.M., Packman, C.H., Ginsberg S.J., 1991. Cardiac arrest after<br />

autologous marrow infusion. Bone Marrow Transplant 7: 401-403.<br />

Röst, M., Quist, P.-O., 2003. <strong>Dissolution</strong> <strong>of</strong> USP prednisone calibrator tablets effects <strong>of</strong><br />

stirring conditions <strong>and</strong> particle size distribution. J Pharm Biomed Anal 31: 1129-1143.<br />

Santos, N.C., Figueira-Coelho, J., Martins-Silva, J., Saldanha, C., 2003. Multidisciplinary<br />

utilization <strong>of</strong> dimethyl sulfoxide: pharmacological, cellular <strong>and</strong> molecular aspects.<br />

Biochem Pharmacol 65: 1035-1041.<br />

59


Savolainen, M., Kogermann, K., Heinz, A., Aaltonen, J., Peltonen, L., Strachan, C.,<br />

Yliruusi, J., 2009. Better underst<strong>and</strong>ing <strong>of</strong> dissolution behaviour <strong>of</strong> amorphous drugs<br />

by in situ solid-state analysis using Raman spectroscopy. Eur J Pharm Biopharm 71:<br />

71-79.<br />

Serajuddin, A.T.M., 2007. Salt formation to improve drug solubility. Adv <strong>Drug</strong> Del Rev<br />

59: 603-616.<br />

Shah, A.C., Peot, C.B., Ochs, J.F., 1973. Design <strong>and</strong> evaluation <strong>of</strong> a rotating filterstationary<br />

basket in vitro dissolution test apparatus I: Fixed fluid volume system. J<br />

Pharm Sci 62: 671-677.<br />

Sheng, J.J., McNamara D.P., Amidon, G.L., 2009. Toward an in vivo dissolution<br />

methodology: A comparison <strong>of</strong> phosphate <strong>and</strong> bicarbonate buffers. Mol Pharm 6: 29-<br />

39.<br />

Shepherd, R.E., Price, J.C., Luzzi, L.A., 1972. <strong>Dissolution</strong> pr<strong>of</strong>iles for capsules <strong>and</strong> tablets<br />

using a magnetic basket dissolution apparatus. J Pharm Sci 61: 1152-1156.<br />

Singh, T., Dutt, J., 1985. Cyclic voltammetry at the turbular graphite electrode. Reversible<br />

process (theory). J Electroanal Chem 190: 65-73.<br />

Souliman, S., Blanquet, S., Beyssac, E., Cardot, J.-M., 2006. A level A in vitro/in vivo<br />

correlation in fasted <strong>and</strong> fed states using different methods: Applied to solid immediate<br />

release oral dosage forms. Eur J Pharm Sci 27: 72-79.<br />

Stenberg, P., Luthman, K., Artursson, P., 2000. Virtual screening <strong>of</strong> intestinal drug<br />

permeability. J Control Release 65: 231-243.<br />

Stroncek, D.F., Fautsch, S.K., Lasky, L.C., Hurd, D.D., Ramsay, N.K.C., McCullough, J.,<br />

1991. Adverse reactions in patients transfused with cryopreserved marrow.<br />

Transfusion 31: 521-526.<br />

Stuart, M., Box, K., 2005. Chasing equilibrium: Measuring the intrinsic solubility <strong>of</strong> weak<br />

acids <strong>and</strong> bases. Anal Chem 77: 983-990.<br />

Styler, M.J., Topolsky, D.L., Crilley, P.A., Covalesky, V., Bryan, R., Bulova, S., Brodsky,<br />

I., 1992. Transient high grade heart block following autologous bone marrow infusion.<br />

Bone Marrow Transplant 10: 435-438.<br />

Sunesen, V.H., Pedersen, B.L., Kristensen, H.G., Müllertz, A., 2005. In vivo in vitro<br />

correlations for a poorly soluble drug, danazol, using the flow-through dissolution<br />

method with biorelevant dissolution media. Eur J Pharm Sci 24: 305-313.<br />

Suomalainen, P., Johans, C., Söderlund, T., Kinnunen, P.K.J., 2004. Surface activity<br />

pr<strong>of</strong>iling <strong>of</strong> drugs applied to the prediction <strong>of</strong> blood-brain barrier permeability. J Med<br />

Chem 47: 1783-88.<br />

Szejtli, J., 1988. Cyclodextrin technology. Kluwer Academic Publishers, Dordrecht.<br />

Takano, R., Sugano, K., Higashida, A., Hayashi, Y., Machida, M., Aso, Y., Yamashita, S.,<br />

2006. Oral absorption <strong>of</strong> poorly water-soluble drugs: computer simulation <strong>of</strong> fraction<br />

absorbed in humans from a miniscale dissolution test. Pharm Res 23: 1144-56.<br />

Taub, M., Kristensen, L., Frokjaer, S., 2002. Optimized conditions for MDCK<br />

permeability <strong>and</strong> turbidimetric solubility studies using compounds representative <strong>of</strong><br />

BCS classes I-IV. Eur J Pharm Sci 15: 331-340.<br />

Tenho, M., Aaltonen, J., Heinänen, P., Peltonen, L., Lehto, V.-P., 2007. Effect <strong>of</strong> texture<br />

on the intrinsic dissolution behaviour <strong>of</strong> acetylsalicylic acid <strong>and</strong> tolbutamide compacts.<br />

J Appl Cryst 40: 857-864.<br />

Tsuji, A., Tamai, I., 1996. Carrier-mediated intestinal transport <strong>of</strong> drugs. Pharm Res 13:<br />

963-977.<br />

Uesu, N.Y., Pineda, E.A.G., Hechenleitner, A.A.W., 2000. Microcrystalline cellulose<br />

from soybean husk: Effects <strong>of</strong> solvent treatments on its properties as acetylsalicylic<br />

acid carrier. Int J Pharm 206: 85-96.<br />

60


USP 2003/2009, 26 th /32 nd edition, United States Pharmacopoeial Convention, Rockville<br />

MD.<br />

Vertzoni, M., Dressman, J., Butler, J., Hempenstall, J., Reppas, C., 2005. Simulation <strong>of</strong><br />

fasting gastric conditions <strong>and</strong> its importance for the in vivo dissolution <strong>of</strong> lipophilic<br />

compounds. Eur J Pharm Biopharm 60: 413-417.<br />

Vrecer, F., Vrbinc, M., Meden, A., 2003. Characterization <strong>of</strong> piroxicam crystal<br />

modifications. Int J Pharm 256: 3-15.<br />

WHO, 2006. Proposal to waive in vivo bioequivalence requirements for WHO model list<br />

<strong>of</strong> essential medicines immediate-release, solid oral dosage forms, in WHO Technical<br />

report series, N:o 937.<br />

Wood, J.H., Syarto, J.E., Letterman, H.J., 1965. Improved holder for intrinsic dissolution<br />

rate study. J Pharm Sci 54: 1068.<br />

Wu, C.-Y., Benet, L.B., 2005. Predicting drug disposition via application <strong>of</strong> BCS:<br />

Transport/Absorption/Elimination interplay <strong>and</strong> development <strong>of</strong> a biopharmaceutics<br />

drug disposition classification system. Pharm Res 22: 11-23.<br />

Yazdanian, M., Briggs, K., Yankovsky, C., Hawi, A., 2004. The ‘high solubility’<br />

definition on the current FDA guidance on biopharmaceutical classification system<br />

may be too strict for acidic drugs. Pharm Res 21: 293-299.<br />

Yu, L.X., Amidon, G.L., Polli, J.E., Zhao, H., Mehta, M.U., Conner, D.P., Shah, V.P.,<br />

Lesko, L.J., Chen, M.-L., Lee, V.H.L., Hussain, A.S., 2002. Biopharmaceutics<br />

classification system: The scientific basis for biowaiver extensions. Pharm Res 19:<br />

921-925.<br />

Yu, L.X., Carlin, A.S., Amidon, G.L., Hussain, A.S., 2004. Feasibility studies <strong>of</strong> utilizing<br />

disk intrinsic dissolution rate to classify drugs. Int J Pharm 270: 221-227.<br />

Zhou, L., Yang, L., Tilton, S., Wang, J., 2007. Development <strong>of</strong> a high throughput<br />

equilibrium solubility assay using miniaturized shake-flask method in early drug<br />

discovery. J Pharm Sci 96: 3052-71.<br />

61

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!