permeability via dual tackling of its solubility and ... · contact rania h. fahmy...

10
Full Terms & Conditions of access and use can be found at http://www.tandfonline.com/action/journalInformation?journalCode=iddi20 Download by: [Rania Fahmy] Date: 14 May 2017, At: 03:48 Drug Development and Industrial Pharmacy ISSN: 0363-9045 (Print) 1520-5762 (Online) Journal homepage: http://www.tandfonline.com/loi/iddi20 Development of novel amisulpride-loaded liquid self-nanoemulsifying drug delivery systems via dual tackling of its solubility and intestinal permeability Wael Gamal, Rania H. Fahmy & Magdy I. Mohamed To cite this article: Wael Gamal, Rania H. Fahmy & Magdy I. Mohamed (2017): Development of novel amisulpride-loaded liquid self-nanoemulsifying drug delivery systems via dual tackling of its solubility and intestinal permeability, Drug Development and Industrial Pharmacy, DOI: 10.1080/03639045.2017.1322607 To link to this article: http://dx.doi.org/10.1080/03639045.2017.1322607 Accepted author version posted online: 27 Apr 2017. Published online: 11 May 2017. Submit your article to this journal Article views: 10 View related articles View Crossmark data

Upload: others

Post on 14-Sep-2019

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: permeability via dual tackling of its solubility and ... · CONTACT Rania H. Fahmy raniafahmy@gmail.com Department of Pharmaceutics and Industrial Pharmacy, Faculty of Pharmacy, Cairo

Full Terms & Conditions of access and use can be found athttp://www.tandfonline.com/action/journalInformation?journalCode=iddi20

Download by: [Rania Fahmy] Date: 14 May 2017, At: 03:48

Drug Development and Industrial Pharmacy

ISSN: 0363-9045 (Print) 1520-5762 (Online) Journal homepage: http://www.tandfonline.com/loi/iddi20

Development of novel amisulpride-loaded liquidself-nanoemulsifying drug delivery systemsvia dual tackling of its solubility and intestinalpermeability

Wael Gamal, Rania H. Fahmy & Magdy I. Mohamed

To cite this article: Wael Gamal, Rania H. Fahmy & Magdy I. Mohamed (2017): Developmentof novel amisulpride-loaded liquid self-nanoemulsifying drug delivery systems via dual tacklingof its solubility and intestinal permeability, Drug Development and Industrial Pharmacy, DOI:10.1080/03639045.2017.1322607

To link to this article: http://dx.doi.org/10.1080/03639045.2017.1322607

Accepted author version posted online: 27Apr 2017.Published online: 11 May 2017.

Submit your article to this journal

Article views: 10

View related articles

View Crossmark data

Page 2: permeability via dual tackling of its solubility and ... · CONTACT Rania H. Fahmy raniafahmy@gmail.com Department of Pharmaceutics and Industrial Pharmacy, Faculty of Pharmacy, Cairo

RESEARCH ARTICLE

Development of novel amisulpride-loaded liquid self-nanoemulsifying drugdelivery systems via dual tackling of its solubility and intestinal permeability

Wael Gamala, Rania H. Fahmya,b and Magdy I. Mohamedb

aDepartment of Pharmaceutics, Faculty of Pharmacy, Ahram Canadian University, Cairo, Egypt; bDepartment of Pharmaceutics and IndustrialPharmacy, Faculty of Pharmacy, Cairo University, Cairo, Egypt

ABSTRACTObjective: The aim of the current investigation was at enhancing the oral biopharmaceutical behavior;solubility and intestinal permeability of amisulpride (AMS) via development of liquid self-nanoemulsifyingdrug delivery systems (L-SNEDDS) containing bioenhancing excipients.Methods: The components of L-SNEDDS were identified via solubility studies and emulsification efficiencytests, and ternary phase diagrams were constructed to identify the efficient self-emulsification regions. Theformulated systems were assessed for their thermodynamic stability, globule size, self-emulsification time,optical clarity and in vitro drug release. Ex vivo evaluation using non-everted gut sac technique wasadopted for uncovering the permeability enhancing effect of the formulated systems.Results: The optimum formulations were composed of different ratios of CapryolTM 90 as an oil phase,CremophorVR RH40 as a surfactant, and TranscutolVR HP as a co-surfactant. All tested formulations werethermodynamically stable with globule sizes ranging from 13.74 to 29.19nm and emulsification time notexceeding 1min, indicating the formation of homogenous stable nanoemulsions. In vitro drug releaseshowed significant enhancement from L-SNEDDS formulations compared to aqueous drug suspension.Optimized L-SNEDDS showed significantly higher intestinal permeation compared to plain drug solutionwith nearly 1.6–2.9 folds increase in the apparent permeability coefficient as demonstrated by the ex vivostudies.Conclusions: The present study proved that AMS could be successfully incorporated into L-SNEDDS forimproved dissolution and intestinal permeation leading to enhanced oral delivery.

ARTICLE HISTORYReceived 4 January 2017Revised 4 April 2017Accepted 18 April 2017

KEYWORDSAmisulpride; bioenhancers;ex vivo permeability testing;non-everted gut sac model;self-nanoemulsifying; tern-ary phase diagrams

Introduction

Amisulpride (AMS) is a benzamide atypical antipsychotic drug mol-ecule that acts selectively on both dopamine D2 and D3 receptortypes with low affinity for any other receptor system leading toremarkable reduction in its side effects [1]. Additionally, AMS ischaracterized by its low risk of extrapyramidal symptoms and aunique ability in improving both positive and negative symptomsof schizophrenia [2].

Despite the distinctive neurochemical and pharmacological pro-files of AMS, it suffers from two major drawbacks; pH dependentsolubility and poor intestinal permeability leading to low oral bio-availability (approximately 48%). Regarding solubility, being aweakly basic drug of pKa ¼8.99 [3], AMS is highly soluble in acidicmedium (stomach pH conditions) but suffers from poor solubilityin neutral or high pH medium (intestinal pH conditions). This pHdependent solubility might cause drug precipitation lately afteroral administration when reaching the intestinal medium leadingto reduced drug absorption. Moreover, AMS suffers poor intestinalpermeability as it has been identified as a substrate for the intes-tinal P-glycoprotein (P-gp) efflux pump [4]. This barrier can beconsidered also as a main cause for the reduced oral bioavailabil-ity and clinical effectiveness of AMS.

Self-nanoemulsifying drug delivery systems are uniform mix-tures of oil, surfactant, co-surfactant and the investigated drug,

which upon contacting the aqueous gastrointestinal fluids, rap-idly emulsify into o/w nanoemulsion with minute droplet sizesand very high interfacial area. The formed nanoemulsions canbypass the dissolution step and introduce the drug to the GITin a completely solubilized state ready for absorption [5].Moreover, some of the excipients used in SNEDDS formulationsuch as TweenVR 80, CremophorVR RH40, CremophorVR EL,LabrasolVR and pluronics are claimed to be ‘bioenhancers.’ Suchbioenhancing properties arise from their inhibitory effect on theintestinal P-gp efflux pump activity leading to enhanced perme-ability and oral bioavailability of poorly absorbable drugs [6].Ghai and Sinha succeeded in improving the solubility and per-meability of the b1-adrenoreceptor blocker talinolol usingSNEDDS comprised of triacetin as the oily phase, Brij-721 as sur-factant and ethanol as co-surfactant [7]. Also, Bandyopadhyayet al. [8], formulated SNEDDS using Maisine 35-1 as long chaintriglyceride (LCT) or CapryolTM 90 as medium chain triglyceride(MCT) together with LabrasolVR and TweenVR 80 as surfactants forimproving the bioavailability of ezetimibe through enhancingthe poor solubility and permeability of the drug.

Literature lacks any evidence about the role of L-SNEDDS insolving both solubility and intestinal permeability problems ofAMS. Therefore, the present study targets the formulation, in vitroand ex vivo evaluation of stable L-SNEDDS of AMS for improvingits oral delivery.

CONTACT Rania H. Fahmy [email protected] Department of Pharmaceutics and Industrial Pharmacy, Faculty of Pharmacy, Cairo University, Kasr El-Aini,Cairo 11562, Egypt� 2017 Informa UK Limited, trading as Taylor & Francis Group

DRUG DEVELOPMENT AND INDUSTRIAL PHARMACY, 2017https://doi.org/10.1080/03639045.2017.1322607

Page 3: permeability via dual tackling of its solubility and ... · CONTACT Rania H. Fahmy raniafahmy@gmail.com Department of Pharmaceutics and Industrial Pharmacy, Faculty of Pharmacy, Cairo

Materials and methods

Materials

Amisulpride pure drug was a gift sample from Al Andalous forpharmaceutical industries (Cairo, Egypt). CapryolTM 90 (propyleneglycol monocaprylate), LabrafacTM Lipophile WL1349 (caprylic/cap-ric triglyceride), LabrasolVR (PEG-8 caprylic/capric glycerides),LabrafilVR M2125CS (linoleoyl macrogol-6 glycerides), PeceolTM (gly-ceryl monooleate) and TranscutolVR HP (diethylene glycol mono-ethyl ether) were kind gifts from Gattefosse (Saint-Priest, Lyon,France). MiglyolVR 812 (caprylic/capric triglyceride) and MiglyolVR

840 (propylene glycol dicaprylate/dicaprate) were purchased fromSasol (Witten, Germany). CremophorVR RH40 (polyoxyl-40 hydrogen-ated castor oil) was purchased from BASF (M€unster, Germany).CremophorVR EL (polyoxyl-35 hydrogenated castor oil) was pur-chased from Sigma-Aldrich (St. Louis, MO, USA). Isopropyl myris-tate (IPM) was purchased from Loba Chemie (Mumbai, India).Propylene glycol (PG), Polyethylene glycol 400 (PEG 400) andTweenVR 80 were purchased from El-Nasr pharmaceutical chemicalsCo. (Cairo, Egypt). Dialysis tubing cellulose membrane (Molecularweight cut off of 14 000) was purchased from Sigma-AldrichChemical Co. (St. Louis, MO, USA). All other chemicals and solventswere of analytical grade and used without further purification.

Determination of saturated solubility of AMS

The equilibrium solubility of AMS in various oils, surfactants andco-surfactants was determined. Excess AMS powder was added on2g of each of the investigated vehicles in a stoppered glass tube,followed by 30 s vortexing (BenchMixer; Benchmark Scientific Inc.,Edison, NJ, USA). The tubes were then left at 37 ± 0.5 �C for 48 h ina thermostatically controlled shaking water bath (Memmert; WNB14, Jakarta, Indonesia), then centrifuged at 4000 rpm for 5min.Samples were then filtered on a 0.22mm membrane filter and fil-trates were suitably diluted with ethanol and measured spectro-photometrically at 277.4 nm using ethanol as a blank (UV-1800;Shimadzu, Kyoto, Japan).

Emulsification ability screening of surfactants and co-surfactants

Different surfactants were screened for their emulsification abilityaccording to the method formerly described by Date andNagarsenker [9]. Briefly, 300mg of each surfactant was added to300mg of the oil phase and the mixtures were gently heated at50 ± 0.5 �C for homogenization of the components. Then, 50mgof each mixture was diluted with 50ml distilled water in a glass-stoppered conical flask. Ease of emulsification was determined bythe number of flask inversions required to yield homogenousemulsion. Emulsions were then allowed to stand for 2 h and theirpercentage transmittance (%) was measured at 638.2 nm usingdouble-beam UV spectrophotometer (UV-1800; Shimadzu, Kyoto,Japan) using distilled water as blank [10].

Regarding co-surfactants, mixtures of 100mg co-surfactant,200mg surfactant and 300mg oil were prepared and evaluated ina similar fashion as described in the screening of surfactants [9].

Construction of ternary phase diagrams

Ternary mixtures comprised of variable concentrations of selectedoil, surfactant and co-surfactant were prepared. The mixture compo-nents were always added up to 100%, and five ternary phase dia-grams were constructed and evaluated (Figure 1(A–E)). From eachmixture point on the phase diagram, 1 g of the ternary mixture wasprepared and homogenized by magnetic stirring with a rotationspeed of 125 rpm at 50 �C for 3min, then diluted up to 10ml with

distilled water, followed by immediate visual observation for investi-gating self-emulsification ability and presence of any phase separ-ation [11]. The diluted nanoemulsions were left for 24 h at 25 �C andreassessed regarding the previous criteria for stability purpose. Onlydispersions with clear or slightly bluish appearance were regardedas the nanoemulsion region of the diagram [12]. The ternary phasediagrams were plotted using SigmaPlotVR software (Version 12.5;Systat Software Inc., San Jose, CA, USA).

Determination of drug loading capacity of L-SNEDDS

Excess AMS powder was added on 2g of each of the investigatedSNEDDS systems (composition shown in Table 1) in a stopperedglass tube, followed by 30 s vortexing. The tubes were then left at37 ± 0.5 �C for 48 h in a thermostatically controlled shaking waterbath, then centrifuged at 4000 rpm for 5min. Samples were thenfiltered on a 0.22mm membrane filter and filtrates were suitablydiluted with ethanol and measured spectrophotometrically at277.4 nm using ethanol as a blank.

Formulation of AMS-loaded L-SNEDDS

Self-nanoemulsifying formulations that proved satisfactory AMSloading efficiency (�50mg/g) were selected for drug loading andfurther characterization. 50mg amounts of AMS (target dose) wereadded to the homogenous blends of oil, surfactant and co-surfac-tant, followed by magnetic stirring at 125 rpm and 50 �C (Stuart,SB162; Stone, Staffordshire, UK) till formation of a completely clearsystem. Prepared drug-loaded L-SNEDDS were stored at roomtemperature for upcoming evaluation.

Evaluation of AMS-loaded L-SNEDDS

Assessment of thermodynamic stability and globule size analysisThermodynamic stability of the formulated AMS-loaded L-SNEDDSwas assessed by visual observation of any signs of phase separ-ation or drug precipitation after centrifugation for 30min at3500 rpm, followed by six cooling heating cycles (4 �C and 45 �C)and finally three freeze thaw cycles (�21 �C and þ25 �C). Storageperiod at each temperature was not less than 48 h [13].

The mean globule size and polydispersity index (PDI) ofL-SNEDDS were determined by photon correlation spectroscopyusing a Malvern Zetasizer Nano ZS (Malvern Instruments Ltd.,Worcestershire, UK) after 100 times dilution with aqueous buffers(0.1 N HCl (pH 1.2) and phosphate buffer (pH 7.4)). Data were stat-istically analyzed using one-way ANOVA, followed by Post Hocleast significant difference test (LSD) to detect the source of thedifference (SPSS 23.0, Chicago, IL, USA). p Values less than .05were considered to be significant.

Transmission electron microscopyMorphology of nanoemulsion droplets was studied using JEOLelectron microscope (JEM-2100, Peabody, MA, USA). Formula (F2)was selected for examination under microscope as it showedacceptably minute globule size and low PDI values compared toother formulations. L-SNEDDS formula was diluted 100 timeswith distilled water, then 1–2 drops were taken onto TEM grid(400-mesh carbon coated grids). The sample was then negativelystained using a drop of 1% phosphotungistic acid for 30 s.

Determination of self-emulsification timeThe rate of emulsification was determined by adding 1 g of AMS-loaded L-SNEDDS to 200ml 0.1 N HCl (pH 1.2) at 37 �C and

2 W. GAMAL ET AL.

Page 4: permeability via dual tackling of its solubility and ... · CONTACT Rania H. Fahmy raniafahmy@gmail.com Department of Pharmaceutics and Industrial Pharmacy, Faculty of Pharmacy, Cairo

Figu

re1.

Ternaryph

asediagramsshow

ing

theself-nanoem

ulsification

regionsforsystem

scompo

sed

ofCapryolTM

90as

anoil,Transcutol

VRHPas

acosurfa

ctant,

and

different

surfactants(a)Crem

ophorVR

RH40,

(b)Crem

ophorVR

EL,(c)Tw

eenV

R

80,(d)

Crem

ophorVR

RH40:TweenV

R

80(1:1),(e)Crem

ophorVR

EL:TweenV

R

80(1:1).Areasof

nanoem

ulsion

region

sareshaded.

DRUG DEVELOPMENT AND INDUSTRIAL PHARMACY 3

Page 5: permeability via dual tackling of its solubility and ... · CONTACT Rania H. Fahmy raniafahmy@gmail.com Department of Pharmaceutics and Industrial Pharmacy, Faculty of Pharmacy, Cairo

magnetically stirred at 50 rpm. The time required for complete dis-appearance of the preconcentrate and formation of homogenoussingle phase system was assessed visually [14].

Spectroscopic characterization of optical clarityOptical clarity of the aqueous nano-dispersions produced fromSNEDDS emulsification was measured spectrophotometrically.AMS-loaded L-SNEDDS were diluted to 100 times with aqueousbuffers of different pH (0.1 N HCl (pH 1.2) and phosphate buffer(pH 7.4)). The absorbance of each solution was measured at400 nm using distilled water as a blank [14]. Data were statisticallyanalyzed using one-way ANOVA, followed by Post Hoc LSD todetect the source of the difference (SPSS 23.0, Chicago, IL, USA).p Values less than .05 were considered to be significant.

In vitro drug release studyThe selected AMS-loaded L-SNEDDS (F1, F2, F3, F4, F8, F13 andF14) (50mg/g) were placed, with 5ml of the dissolution mediumin a dialysis bag (MWCO 14 000 g/mole; Sigma Aldrich, St. Louis,MO, USA), and then the bag was securely tied from both endsusing a thermo-resistant thread. The loaded bags were then tiedto the paddles of the dissolution tester (apparatus II, rotating pad-dle, DIS 6000, Copley Scientific; Colwick, Nottingham, UK), andimmersed in 900ml freshly prepared dissolution medium (0.1 NHCl (pH 1.2) or phosphate buffer (pH 7.4)) rotated at 100 rpm andmaintained at 37± 0.5 �C. 5ml samples were withdrawn at regulartime intervals (0.25, 0.5, 1, 1.5, 2, 4, 6 and 8 h) and replaced by anequal amount of fresh dissolution medium in order to maintainsink conditions. Withdrawn samples were filtered through 0.22mmmembrane filter, suitably diluted and analyzed for the drug con-tent using UV spectrophotometer at 280 nm using the buffer asblank. The release patterns of AMS from the tested L-SNEDDSwere compared to that of the drug aqueous suspension (dialysisbag containing 50mg AMS powder in 5ml dissolution medium)using the similarity factor (f2) according to the following equation:

f2 ¼ 50� logf½1þ ð1=nÞXn

t¼1ðRt � TtÞ2��0:5 � 100g (1)

Ex vivo intestinal permeability study

Non-everted gut sac modelEx vivo intestinal permeation studies of AMS-loaded L-SNEDDSwere carried out using the non-everted gut sac technique. Thestudy protocol adhered to the European community standardsand the legislation for the protection of animals used for scientificpurposes (Directive 2010/63/EU) (European Union Law, 2010) andwas approved by the Research Ethical Committee of the Faculty ofPharmacy, Cairo University (approval number PI-959). Overnight-fasted male Wistar rats (200–250 g) were sacrificed by spinal dis-location, followed by removal of the intestine from the upper endof the duodenum to the lower end of the ileum, then manuallystripping the mesentery [15]. The removed intestinal part wascarefully washed with ice-cold Krebs solution (pH 6.5) using ablunt-ended syringe [16], then divided into 8 ± 0.5 cm long tubeswith 3.0 ± 0.5mm diameter. The prepared tubes were tied fromone end using braided silk suture and then filled with 1ml ofselected L-SNEDDS formula (equivalent to 1mg of AMS) via ablunt needle. The other end was then tied forming a closed sacwhich was immersed in 40ml oxygenated Krebs solution pre-warmed to 37 �C in a jacketed glass assembly. 3ml samples werewithdrawn from the serosal medium (Krebs solution) at predeter-mined time intervals (20, 40, 60, 80, 100 and 120min) and replen-ished with equal volume of fresh Krebs solution. The collectedsamples were filtered through 0.22 mm membrane filter and theamount of AMS permeated from mucosal to serosal medium wasdetermined by measuring the withdrawn samples using HPLCmethod. The same experiment was repeated (n¼ 3) using similarconcentration of AMS aqueous Krebs solution.

The concentration of AMS in the withdrawn samples was ana-lyzed by HPLC system using slightly modified HPLC method forAMS determination developed by Devadasu and Ravisankar andwas previously validated in terms of specificity, precision, linearity,accuracy, recovery, robustness and system suitability [17]. TheHPLC instrument (LC-2000 plus; Jasco Inc., Tokyo, Japan) wasequipped with a reversed-phase C18 column (25 cm �4.6mm; par-ticle size of 5 mm). The isocratic mobile phase (10mM phosphate

Table 1. Formulation composition and evaluation of L-SNEDDS selected from the optimum phase diagram.

L-SNEDDS formulanumber

Composition

Drug loadingcapacity (mg/g)±SD

Globule size (nm) ± SD

Self-emulsificationtime (s)±SD

CapryolTM

90 % (w/w)CremophorVR RH40%

(w/w)TranscutolVR

HP % (w/w) pH 1.2 pH 7.4

F1 10 20 70 69.79 ± 1.82 16.88 ± 0.77 16.56 ± 0.99 31 ± 1.41F2 10 30 60 64.12 ± 0.93 16.48 ± 1.22 16.25 ± 1.38 30 ± 2.12F3 10 40 50 59.88 ± 1.57 15.02 ± 1.12 16.30 ± 2.02 32 ± 0.71F4 10 50 40 54.93 ± 0.45 14.07 ± 2.45 13.74 ± 2.67 35 ± 2.83F5 10 60 30 49.65 ± 0.57

NA NAF6 10 70 20 48.33 ± 0.55F7 10 80 10 35.68 ± 1.93F8 20 30 50 61.62 ± 1.31 22.83 ± 0.44 21.42 ± 0.79 35 ± 2.82F9 20 40 40 48.56 ± 0.70

NA NAF10 20 50 30 44.73 ± 1.46F11 20 60 20 42.57 ± 2.10F12 20 70 10 37.97 ± 0.88F13 30 40 30 52.13 ± 1.41 21.02 ± 1.87 22.69 ± 0.68 34 ± 3.54F14 30 50 20 51.55 ± 1.82 29.19 ± 2.59 28.96 ± 5.80 45 ± 1.41F15 30 60 10 32.57 ± 1.52

NA NAF16 40 40 20 39.55 ± 1.18F17 40 50 10 38.71± 0.83

SD: standard deviation from the mean; NA: data not available due to low drug loading of less than 50mg/g.Data are the mean values (n¼ 3) ± SD.Bold italic values in Table 1 indicate drug loading of more than 50mg/g.

4 W. GAMAL ET AL.

Page 6: permeability via dual tackling of its solubility and ... · CONTACT Rania H. Fahmy raniafahmy@gmail.com Department of Pharmaceutics and Industrial Pharmacy, Faculty of Pharmacy, Cairo

buffer pH 4.0: Acetonitrile; 80:20) was run at a flow rate of 1.0mlper minute at room temperature and the column effluent wasmonitored using an ultraviolet (UV) detector set at 280 nm (UV-2070 Plus; Jasco Inc., Tokyo, Japan). The injection volume was20ml and the system was equipped with an interface operated byChromNAV data system software (v 2.0; Jasco Inc., Tokyo, Japan).Under a concentration range of 0.1–20mg/ml, retention time was5.6 ± 0.3min and the calibration curve equation of peak areaagainst AMS concentration (R2¼ 0.999) was:

Y ¼ 59067:6327xþ9561:5465

Calculation of apparent permeability coefficient (Papp)The apparent permeability coefficient (Papp) of plain drug solutionand AMS-loaded L-SNEDDS was calculated from the followingequation [18]:

Papp ¼ dQdt

� �� 1

A � Co

� �ðcm=sÞ (2)

where dQ/dt is the steady-state rate of drug entrance into theacceptor solution, A is the surface area of the prepared intestinalsacs, and Co is the initial concentration of AMS inside the sacs.dQ/dt was calculated by taking the slope of the linear portion ofthe graph plot between the cumulative amounts of drug perme-ated (mg) through the sac and the time of permeation (minutes).Considering the intestinal sacs are cylindrical in shape with 8 cmlength, 1ml volume and 0.30 cm inner diameter, the surface area(A) is calculated to be 7.54 cm2. Statistical analysis of the intestinalpermeation results was carried out using the Paired Student’st-test (p< .05).

Results and discussion

Determination of saturated solubility of AMS

The drug loading per formulation is a very critical factor in thedevelopment of nanoemulsion systems for poorly soluble drugs,which is dependent on the drug solubility in various formulationcomponents (oil, surfactant and co-surfactant). The solubilizingcapacity of oil is considered as a controlling criterion regarding itsselection for nanoemulsion formulation [19]. Results presented inTable 2 revealed that CapryolTM 90 had the highest solubilizingcapacity among the tested oils, which could be attributed to thepolarity of AMS molecule that might lead to more favorable solu-bilization in small/medium chain oils, and in mono-glyceridesrather than di- or tri-glycerides [20]. Therefore, CapryolTM 90 waschosen as the oil phase for further investigation.

Regarding surfactants, Table 2 demonstrated that LabrasolVR

had the highest drug solubility, followed by TweenVR 80,CremophorVR RH40, and finally CremophorVR EL. Moreover,TranscutolVR HP proved superiority in AMS solubilization (68.38mg/g)compared to the other tested co-surfactants (Table 2).

Since the selection of surfactants and co-surfactants shoulddepend on their emulsification efficiency not only on their drugsolubilizing ability [9], screening them for their emulsificationpower was a must.

Emulsification ability screening of surfactants and co-surfactants

Non-ionic surfactants were selected in this investigation since theyare generally regarded as safe and considered less toxic than ionicsurfactants. They also have lower CMCs which make them morepreferable for oral ingestion, they are less affected by pH and ionicstrength changes, and are biocompatible [19,20].

Moreover, the four nonionic surfactants selected in this study(LabrasolVR , TweenVR 80, CremophorVR RH40 and CremophorVR EL)offer reported bioactive effects such as the inhibitory effects onP-gp and CYP enzymes possessed by CremophorVR RH40 andCremophorVR EL [6,21].

After selection of CapryolTM 90 as the oily phase of highestdrug solubility, the goal was to identify the surfactant having thehighest emulsification efficiency with the chosen oil phase. Table 3represents variable transmittance values (%) of different oil-surfac-tant mixtures proving that the tested surfactants have variableemulsification efficiencies towards CapryolTM 90. The largest num-ber of flask inversions and the least UV transmittance (%) werereported for LabrasolVR , while on the other hand, very little numberof flask inversions and high UV percent transmittance werereported for CremophorVR RH40, CremophorVR EL, and TweenVR 80(99.77%, 98.86% and 90.16% respectively). Although, AMS solubil-ity in the three surfactants CremophorVR RH40, CremophorVR EL andTweenVR 80 was lower than in LabrasolVR as shown in Table 2,nevertheless, their higher emulsification efficiency motivated theirselection for further investigation.

Table 3. Percent UV transmittance and number of flask inversions for emulsions prepared using CapryolTM 90 as oilyphase.

Surfactants

CremophorVR RH40 CremophorVR EL TweenVR 80 LabrasolVR

UV transmittance (%) 99.77 ± 0.19 98.86 ± 0.52 90.16 ± 0.82 50 ± 1.41Number of flask inversions� 1 1 3 >20

Data are mean values (n¼ 3) ± SD.�Medians of number of flask inversions required for emulsion formation.

Table 2. Saturated solubility of AMS in various vehicles.

Type of vehicle Solubility (mg/g) ± SD

AMS aqueous solubility 0.18 ± 0.05

OilsCapryolTM 90 29.28 ± 0.66PeceolTM 17.67 ± 0.15MiglyolV

R

840 0.93 ± 0.26MiglyolVR 812 0.80 ± 0.21LabrafacTM Lipophile WL1349 0.74 ± 0.14Isopropyl myristate 0.21 ± 0.09

SurfactantsLabrasolVR 38.07 ± 0.67TweenVR 80 35.83 ± 0.50CremophorVR RH40 26.65 ± 0.46CremophorVR EL 17.13 ± 0.97

Co-surfactantsTranscutolVR HP 68.38 ± 1.54PEG 400 44.86 ± 0.26PG (Propylene glycol) 27.93 ± 0.30LabrafilVR M2125CS 5.28 ± 1.68

Data are the mean values (n¼ 3) ± SD.SD: standard deviation from the mean.

DRUG DEVELOPMENT AND INDUSTRIAL PHARMACY 5

Page 7: permeability via dual tackling of its solubility and ... · CONTACT Rania H. Fahmy raniafahmy@gmail.com Department of Pharmaceutics and Industrial Pharmacy, Faculty of Pharmacy, Cairo

Addition of a co-surfactant to the surfactant-containing formu-lations was reported to improve dispersibility and drug absorption,and obtain nanoemulsion systems at low surfactant concentration(to minimize the toxicity concerns) [5]. Three co-surfactants(TranscutolVR HP, PEG 400 and PG) were tested with combinationof CapryolTM 90 with the three selected surfactants (CremophorVR

RH40, CremophorVR EL and TweenVR 80), and compared regardingtheir percentage UV transmittance and number of flask inversions.The three investigated co-surfactants showed high emulsificationpower with all oil-surfactant combinations tested (Table 4), there-fore, the final selection of the co-surfactant was based on their dif-ferential solubilization of AMS. As TranscutolVR HP showed thehighest drug solubility (Table 2), and it was also known by its highdissolving power, good water solubility, as well as, its permeabilityenhancing effect through P-gp inhibition [6,22], it was thereforeselected as a promising co-surfactant for further investigation.

Construction of ternary phase diagrams

The self-nanoemulsifying formulations’ fields that could readilyemulsify upon dilution and gentle agitation were identified fromthe ternary phase diagrams of the multiple systems based onCapryolTM 90 as the oily phase, three surfactants (CremophorVR

RH40, CremophorVR EL and TweenVR 80), and TranscutolVR HP as co-surfactant. Five ternary phase diagrams were constructed, withtheir compositions shown in Figure 1(A–E). The ternary phase dia-grams revealed the optimum contents and ratios that could beused for the preparation of stable L-SNEDDS. Shaded regions onthe diagrams indicate the nanoemulsion regions; larger shadedregions indicate better self-nanoemulsifying ability.

With all the systems having CapryolTM 90 as oily phase andTranscutolVR HP as co-surfactant, system (A) having CremophorVR

RH40 as a surfactant, showed the largest nanoemulsion region asit could incorporate up to 40% (w/w) of the oily phase with theproduction of a clear and homogenous system. System (B), havingCremophorVR EL, showed smaller nanoemulsion region, while sys-tem (C), having TweenVR 80, showed the smallest nanoemulsionregion. This confirms the previous emulsification power results ofdifferent surfactants with CapryolTM 90, which was descendingfrom CremophorVR RH40 (highest), to CremophorVR EL, then TweenVR

80 (least) (Table 3).1:1 mixtures of either CremophorVR RH40 or CremophorVR EL

with TweenVR 80 (systems D and E) were tested in order to investi-gate if this will lead to significant improvement in the nanoemulsi-fication ability of the mixture compared to the single surfactant.However, the results presented in Figure 1 illustrated less oil incor-porated into the systems compared to using the correspondingsurfactant alone (CremophorVR RH40 or CremophorVR EL). This alsomay be attributed to the higher emulsification ability of any ofthese surfactants with CapryolTM 90 than TweenVR 80 (Table 3).

Furthermore, all the preconcentrates that gave clear nanoemul-sion after dilution showed no phase separation after a period of24 h storage at 25 �C. Conclusively, system A (CapryolTM 90/CremophorVR RH40/TranscutolVR HP) showed the largest nanoemul-sion region compared to the other prepared systems, and so wasselected as optimum system for further investigation.

Determination of drug loading capacity of L-SNEDDS

Seventeen self-nanoemulsifying formulations observed from thenanoemulsion region of system A (CapryolTM 90/CremophorVR

RH40/TranscutolVR HP) were assessed for their AMS-loading abilityas indicated by the drug saturated solubility in these formulations(Table 1). Generally, it was observed that for each oil concentra-tion, increasing TranscutolVR HP concentration was associated withcorresponding increase in AMS saturated solubility in the system,although CremophorVR RH40 concentration declined in the samemanner. This confirms the dominating solubilizing power of theco-surfactant rather than the surfactant in these systems as pre-sented in Table 2.

Moreover, Table 1 clarifies that all the formulated L-SNEDDSsolubilized more than 30mg/g AMS. However, only seven of theformulations, F1, F2, F3, F4, F8, F13 and F14, succeeded in solubi-lizing more than 50mg/g (50mg is the target dose) and therefore,selected for further investigation.

Evaluation of AMS-loaded L-SNEDDS

Assessment of thermodynamic stability and globule size analysisAll the seven tested L-SNEDDS (F1, F2, F3, F4, F8, F13 and F14)proved to be thermodynamically stable with no signs of drug pre-cipitation after centrifugation, heating–cooling and freeze–thawcycles.

Globule size measurement is a very critical step in L-SNEDDSevaluation as it gives a strong indication about the self-emulsification effectiveness which impact the rate and extent ofdrug dissolution, as well as, drug absorption. Due to considerablepH variations along the GIT, it was rational to observe the conse-quence of different media pH (1.2 and 7.4) on the formulatedSNEDDS. From the globule size analysis results (Table 1), it wasobserved that all the prepared systems possessed globule size<30 nm at different pH values which fulfill the SNEDDS criteria ofhaving a mean globule size value <200 nm. This might be attrib-uted to the use of the proper surfactant/co-surfactant mixturewhich reduced the free energy of the system, giving small globulesize [23]. It was also noted from Table 1 that the pH change didnot show significant effect on the globules sizes (p> .05).

Additionally, Table 1 showed that increasing oil concentrationfrom 10% (F1, F2, F3 and F4) to 20% and 30% (F8, F13 and F14)

Table 4. Percent UV transmittance and number of flask inversions for emulsions prepared by the investigated co-surfactantswith the tested oil-surfactant combinations.

Oil/surfactant mixture Co-surfactant UV transmittance (%) Number of flask inversions�CapryolTM 90/CremophorVR RH40 TranscutolVR HP 99.77 ± 0.47 1

PEG 400 99.54 ± 1.56 1PG (Propylene glycol) 99.77 ± 1.27 1

CapryolTM 90/CremophorVR EL TranscutolVR HP 99.08 ± 0.76 1PEG 400 99.77 ± 0.55 1PG (Propylene glycol) 99.31 ± 1.47 1

CapryolTM 90/TweenVR

80 TranscutolVR

HP 88.1 ± 1.27 5PEG 400 94.62 ± 2.47 3PG (Propylene glycol) 82.79 ± 2.45 9

Data are mean values (n¼ 3) ± SD.�Medians of number of flask inversions required for emulsion formation.

6 W. GAMAL ET AL.

Page 8: permeability via dual tackling of its solubility and ... · CONTACT Rania H. Fahmy raniafahmy@gmail.com Department of Pharmaceutics and Industrial Pharmacy, Faculty of Pharmacy, Cairo

resulted in a significantly larger globule size (p< .05). Similarresults were observed with Chen et al. [24], where they observedthat further addition of oil lead to a significant increase in themean globule size. This could be related to the expansion of oildroplets upon more oil addition [25].

Besides the globule size analysis, PDI was used for the assess-ment of the investigated diluted systems. PDI values are very use-ful in expressing the heterogeneity of globule size distribution inthe formed nanoemulsions [26]. PDI values at pH 1.2 ranged from0.092 to 0.244 with the formulations F13, F2 and F8 having theleast PDI values of 0.092, 0.168 and 0.193 respectively. Also, inmost of the formulations, changing pH from 1.2 to 7.4 resulted ina significant increase in PDI values. This could be attributed to thebasic nature of AMS that might lead to slight precipitation athigher pH values causing a broader range of globule sizedistribution.

Transmission electron microscopyMorphology of the diluted L-SNEDDS was observed using TEM.Photomicrograph in Figure 2 revealed the uniform size distributionand spherical shape of the nanoemulsion globules indicating astable state of the formed nanoemulsion. The sizes of globulesobserved in TEM were consistent with the data produced from theglobule size analysis.

Determination of self-emulsification timeThe rate of emulsification is considered an important index for theassessment of self-emulsification efficiency [22]. Usually, the forma-tion of liquid crystals and gel phases are the first steps ofthe emulsification process that considerably influence globule for-mation and accessibility of the interface for drug to partition[27,28]. As presented in Table 1, self-emulsification time resultsrevealed that all the seven tested formulations possessed rapidself-emulsification time ranging from 30 to 45 s, which is expectedto improve the drug absorption.

Spectroscopic characterization of optical clarityLower absorbance should be obtained with optically clear solu-tions because cloudier solutions will scatter more of the incidentradiation, resulting in higher absorbance [29]. To assess the opticalclarity quantitatively, UV–VIS spectrophotometer was used tomeasure the amount of light of a given wavelength transmittedby the solution. All the formulations showed very low (almost neg-ligible) absorbance values ranging from 0.007 to 0.098 at differentpH values (1.2 and 7.4). This indicates that highly clear and stablesystems were obtained by aqueous dilution of the preparedSNEDDS, which might be attributed to the small globule sizes ofthe prepared formulations (<30 nm).

In vitro drug release studyThe percentage dissolution-time profiles of the nanoemulsion pre-concentrates are shown in Figure 3. In 0.1 N HCl (pH 1.2), Figure3(A), approximately 70% or more of AMS was released after 4 hfrom all the tested systems. Also, no significant difference in drugrelease was observed between the tested formulations and theaqueous drug suspension, with f2 values of 52.21, 60.8, 63.61,59.57, 74.66 and 67.63 for the formulations F1, F2, F3, F8, F13 andF14, respectively, each compared to the aqueous drug suspension.Such good drug release profiles in 0.1 N HCl could be due to thebasic nature of AMS that makes it dissolve completely in theacidic pH.

However, results of drug release at pH 7.4 (Figure 3(B)) provedsignificantly higher drug release from the prepared L-SNEDDScompared to the aqueous drug suspension, with approximately22–32% increase in AMS release after the 8 h period. The f2 valuesat pH 7.4 were 36.06, 31.65, 40.77, 39.64, 37.52, 33.45 and 39.29for the formulations F1, F2, F3, F4, F8, F13 and F14, respectively,compared to the aqueous drug suspension. These findings suggestthat AMS-loaded L-SNEDDS could be used as a useful tool for pre-senting the drug in a solubilized form into the absorption sites

Figure 2. Transmission electron micrograph of the formulated L-SNEDDS (F2)after 100 times dilution with distilled water.

Figure 3. Release profiles of AMS from L-SNEDDS compared to drug suspension in pH 1.2 (a) and 7.4 (b) at 37± 0.5 �C.

DRUG DEVELOPMENT AND INDUSTRIAL PHARMACY 7

Page 9: permeability via dual tackling of its solubility and ... · CONTACT Rania H. Fahmy raniafahmy@gmail.com Department of Pharmaceutics and Industrial Pharmacy, Faculty of Pharmacy, Cairo

and avoiding its precipitation at neutral or basic pH mediuminside the GIT.

Ex vivo intestinal permeability study

A main objective of the current study was to improve the intes-tinal permeability of AMS as it is proven to be a substrate for theP-gp efflux pump, which might be responsible for the poor oralbioavailability of the drug [4]. Therefore, an ex vivo intestinal per-meability model was applied in order to investigate AMS permea-tion–enhancing impact of the formulated L-SNEDDS.

Since no significant differences were observed in the in vitrodrug release between all the seven tested self-nanoemulsifyingformulations in either pH 1.2 or pH 7.4, the selection of formula-tions for ex vivo study was based mainly on safety concernsthrough selection of formulations with minimum amount of sur-factant. Therefore, the self-nanoemulsifying formulations contain-ing the least surfactant concentrations at each oil category (10, 20and 30%) were selected; three formulations which are F2 (10%CapryolTM 90, 30% CremophorVR RH40 and 60% TranscutolVR HP), F8(20% CapryolTM 90, 30% CremophorVR RH40 and 50% TranscutolVR

HP) and F13 (30% CapryolTM 90, 40% CremophorVR RH40 and 30%TranscutolVR HP). (An exception was made for F2; selected insteadof F1 in the 10% oil-SNEDDS category as it had lower PDI values).

Ex vivo absorption models including everted and non-evertedgut sac techniques are widely used as more simple procedures forinvestigating various drug transport mechanisms and predictingin vivo drug absorption in humans [30]. Non-everted gut sac tech-nique was adopted in this study as it is characterized by severaladvantages including more simple procedure, less test samplerequired, easier successive collection of serosal samples and lessmorphological damage to the intestinal tissue compared to theeverted gut sac method [31].

Results of the intestinal permeation study (Figure 4(a)), illus-trated that the three tested L-SNEDDS formulations (F2, F8 andF13) exhibited significantly higher intestinal permeability com-pared to plain drug solution (p< .05). After the 120min-period,approximately 252.99, 410.79 and 427.43mg AMS were permeatedfrom F2, F8 and F13, respectively, compared to only 144.90 mg per-meated from the plain drug solution. The increase in AMS perme-ability among the three tested formulations might be related tothe increase in the oil percentage which could have a role in facili-tating drug penetration across the lipophilic intestinal barrier.

The apparent permeability (Papp) calculated and presentedin Figure 4(b), proved that Papp of plain drug solution(2.855� 10�6 cm/s) was significantly improved by nearly 1.6 times

using F2 (4.643� 10�6cm/s), 2.9 times using F8 (8.412� 10�6 cm/s),and 2.8 times using F13 (8.229� 10�6 cm/s).

Such significant improvement in AMS permeation from theL-SNEDDS formulations compared to the plain drug solution couldbe ascribed to (a) Formulation of AMS into L-SNEDDS, deliveringthe drug to its absorption sites in a completely solubilized form,(b) Presenting the drug in minute nanometric size droplets, whichpositively affected the permeability of AMS and (c) The use of bio-enhancing surfactants and co-surfactants (CremophorVR RH40 andTranscutolVR HP) in the formulation that probably improved theintestinal permeation of AMS through the inhibition of the intes-tinal P-gp efflux pump activity.

Conclusions

Amisulpride solubility and intestinal permeability were significantlyimproved through successful incorporation of the drug intoL-SNEDDS containing bioenhancing excipients (CremophorVR RH40and TranscutolVR HP).The Formulated systems with droplet sizes ofless than 30 nm proved a promising effect in delivering AMS intothe GIT in a soluble form with no risk of precipitation, while inhib-iting the intestinal P-gp efflux pump activity as proved by the exvivo studies. Further studies for the development and in vivoevaluation of solid dosage forms of the optimized AMS-SNEDDSare carried out in another part of this work, which is valuable forproviding patient acceptable AMS dosage forms for more effectivetreatment and enhanced oral bioavailability.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

[1] Schoemaker H, Claustre Y, Fage D, et al. Neurochemicalcharacteristics of amisulpride, an atypical dopamine D2/D3receptor antagonist with both presynaptic and limbicselectivity. J Pharmacol Exp Ther. 1997;280:83–97.

[2] Leucht S, Pitschel-Walz G, Engel RR, et al. Amisulpride, anunusual “atypical” antipsychotic: a meta-analysis of random-ized controlled trials. Am J Psychiatry. 2002;159:180–190.

[3] Schmitt U, Abou El-Ela A, Guo LJ, et al. Cyclosporine A(CsA) affects the pharmacodynamics and pharmacokineticsof the atypical antipsychotic amisulpride probably via

Figure 4. (a) Cumulative permeation of AMS from different self-nanoemulsifying formulations and plain AMS solution through non-everted gut sac technique. (b)Apparent permeability (Papp) of AMS-loaded L-SNEDDS formulations and plain AMS solution across the non-everted rat intestine (�10�6 cm/s), (�significant difference,p< .05 compared to plain drug solution). Composition of all L-SNEDDS is presented in Table 1.

8 W. GAMAL ET AL.

Page 10: permeability via dual tackling of its solubility and ... · CONTACT Rania H. Fahmy raniafahmy@gmail.com Department of Pharmaceutics and Industrial Pharmacy, Faculty of Pharmacy, Cairo

inhibition of P-glycoprotein (P-gp). J Neural Transm.2006;113:787–801.

[4] El Ela AA, Hartter S, Schmitt U, et al. Identification of P-glycoprotein substrates and inhibitors among psychoactivecompounds–implications for pharmacokinetics of selectedsubstrates. J Pharm Pharmacol. 2004;56:967–975.

[5] Porter CJ, Pouton CW, Cuine JF, et al. Enhancing intestinaldrug solubilisation using lipid-based delivery systems. AdvDrug Deliv Rev. 2008;60:673–691.

[6] Al-Mohizea A, Zawaneh F, Alam M, et al. Effect of pharma-ceutical excipients on the permeability of P-glycoproteinsubstrate. J Drug Deliv Sci Technol. 2014;24:491–495.

[7] Ghai D, Sinha VR. Nanoemulsions as self-emulsified drugdelivery carriers for enhanced permeability of the poorlywater-soluble selective beta(1)-adrenoreceptor blockerTalinolol. Nanomedicine 2012;8:618–626.

[8] Bandyopadhyay S, Katare O, Singh B. Optimized self nano-emulsifying systems of ezetimibe with enhanced bioavail-ability potential using long chain and medium chain trigly-cerides. Colloids Surf B Biointerfaces. 2012;100:50–61.

[9] Date AA, Nagarsenker MS. Design and evaluation of self-nanoemulsifying drug delivery systems (SNEDDS) forcefpodoxime proxetil. Int J Pharm. 2007;329:166–172.

[10] Elnaggar YSR, El-Massik MA, Abdallah OY. Self-nanoemulsify-ing drug delivery systems of tamoxifen citrate: design andoptimization. Int J Pharm. 2009;380:133–141.

[11] ElKasabgy NA. Ocular supersaturated self-nanoemulsifyingdrug delivery systems (S-SNEDDS) to enhance econazolenitrate bioavailability. Int J Pharm. 2014;460:33–44.

[12] Zhang P, Liu Y, Feng N, et al. Preparation and evaluation ofself-microemulsifying drug delivery system of oridonin. Int JPharm. 2008;355:269–276.

[13] Shafiq S, Shakeel F, Talegaonkar S, et al. Development andbioavailability assessment of ramipril nanoemulsion formu-lation. Eur J Pharm Biopharm. 2007;66:227–243.

[14] Basalious EB, Shawky N, Badr-Eldin SM. SNEDDS containingbioenhancers for improvement of dissolution and oralabsorption of lacidipine. I: development and optimization.Int J Pharm. 2010;391:203–211.

[15] Freag MS, Elnaggar YS, Abdallah OY. Lyophilized phytoso-mal nanocarriers as platforms for enhanced diosmin deliv-ery: optimization and ex vivo permeation. Int J Nanomed.2013;8:2385–2397.

[16] Al-Mohizea AM. Influence of intestinal efflux pumps on theabsorption and transport of furosemide. Saudi Pharm J.2010;18:97–101.

[17] Devadasu C, Ravisankar P. Validated RP-HPLC method forthe microgram determination of amisulpride in pure and

pharmaceutical formulations. Int J Res Pharm Biomed Sci.2011;2:202–211.

[18] Ruan L-P, Chen S, Yu B-Y, et al. Prediction ofhuman absorption of natural compounds by the non-everted rat intestinal sac model. Eur J Med Chem. 2006;41:605–610.

[19] Pouton CW, Porter CJ. Formulation of lipid-based deliverysystems for oral administration: materials, methods andstrategies. Adv Drug Deliv Rev. 2008;60:625–637.

[20] Azeem A, Rizwan M, Ahmad FJ, et al. Nanoemulsion compo-nents screening and selection: a technical note. AAPSPharmSciTech. 2009;10:69–76.

[21] Chen M-L. Lipid excipients and delivery systems forpharmaceutical development: a regulatory perspective. AdvDrug Deliv Rev. 2008;60:768–777.

[22] Wei L, Sun P, Nie S, et al. Preparation and evaluation ofSEDDS and SMEDDS containing carvedilol. Drug Dev IndPharm. 2005;31:785–794.

[23] Nepal PR, Han H-K, Choi H-K. Preparation and in vitro–invivo evaluation of WitepsolVR H35 based self-nanoemulsify-ing drug delivery systems (SNEDDS) of coenzyme Q 10. EurJ Pharm Sci. 2010;39:224–232.

[24] Chen H, Chang X, Weng T, et al. A study of microemulsionsystems for transdermal delivery of triptolide. J ControlRelease. 2004;98:427–436.

[25] Yuan Y, Li S-m, Mo F-k, et al. Investigation of microemulsionsystem for transdermal delivery of meloxicam. Int J Pharm.2006;321:117–123.

[26] Tripathi A, Gupta R, Saraf SA. PLGA nanoparticles of antitu-bercular drug: drug loading and release studies of awater in-soluble drug. Int J PharmTech Res. 2010;2:2116–2123.

[27] Constantinides PP. Lipid microemulsions for improving drugdissolution and oral absorption: physical and biopharma-ceutical aspects. Pharm Res. 1995;12:1561–1572.

[28] Porter CJ, Charman WN. In vitro assessment of oral lipidbased formulations. Adv Drug Deliv Rev. 2001;50:S127–S147.

[29] Singh SK, Prasad Verma PR, Razdan B. Glibenclamide-loadedself-nanoemulsifying drug delivery system: developmentand characterization. Drug Dev Ind Pharm. 2010;36:933–945.

[30] Sharma P, Chawla H, Panchagnula R. LC determination ofcephalosporins in in vitro rat intestinal sac absorptionmodel. J Pharm Biomed Anal. 2002;27:39–50.

[31] Dixit P, Jain DK, Dumbwani J. Standardization of an ex vivomethod for determination of intestinal permeability ofdrugs using everted rat intestine apparatus. J PharmacolToxicol Methods. 2012;65:13–17.

DRUG DEVELOPMENT AND INDUSTRIAL PHARMACY 9