You are on page 1of 16

Drug Delivery

ISSN: 1071-7544 (Print) 1521-0464 (Online) Journal homepage: https://www.tandfonline.com/loi/idrd20

Novel tablet formulation of amorphous


candesartan cilexetil solid dispersions involving P-
gp inhibition for optimal drug delivery: in vitro and
in vivo evaluation

Gurunath Surampalli, Basavaraj K. Nanjwade, P. A. Patil & Rakesh Chilla

To cite this article: Gurunath Surampalli, Basavaraj K. Nanjwade, P. A. Patil & Rakesh Chilla
(2016) Novel tablet formulation of amorphous candesartan cilexetil solid dispersions involving P-gp
inhibition for optimal drug delivery: in�vitro and in�vivo evaluation, Drug Delivery, 23:7, 2124-2138,
DOI: 10.3109/10717544.2014.945017

To link to this article: https://doi.org/10.3109/10717544.2014.945017

Published online: 31 Jul 2014.

Submit your article to this journal

Article views: 1367

View related articles

View Crossmark data

Citing articles: 6 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=idrd20
http://informahealthcare.com/drd
ISSN: 1071-7544 (print), 1521-0464 (electronic)

Drug Deliv, 2016; 23(7): 2124–2138


! 2014 Informa Healthcare USA, Inc. DOI: 10.3109/10717544.2014.945017

RESEARCH ARTICLE

Novel tablet formulation of amorphous candesartan cilexetil solid


dispersions involving P-gp inhibition for optimal drug delivery:
in vitro and in vivo evaluation
Gurunath Surampalli1, Basavaraj K. Nanjwade2, P. A. Patil3, and Rakesh Chilla4
1
Department of Pharmacology, Vaagdevi Institute of Pharma Sciences, Warangal, Andhra Pradesh, India, 2Faculty of Pharmacy, Department of
Pharmaceutics, Omer Al-Mukhtar University, Tobruk, Libya, 3Department of Pharmacology, International Medical Programme, USM-KLE, Belgaum,
Karnataka, India, and 4Department of Pharmaceutics, Vaagdevi Institute of Pharma Sciences, Warangal, Andhra Pradesh, India

Abstract Keywords
Objective: The aim of this study was to develop a novel tablet formulation of amorphous Candesartan, direct compression, naringin,
candesartan cilexetil (CAN) solid dispersion involving effective P-gp inhibition for optimal drug oral bioavailability
delivery by direct compression (DC) method.
Methods: To accomplish DC, formulation blends were evaluated for micromeritic properties. History
The Carr index, Hausner ratio, flow rate and cotangent of the angle a were determined. The
tablets with and without naringin prepared by DC technique were evaluated for average Received 26 May 2014
weight, hardness, disintegration time and friability assessments. The drug release profiles were Revised 11 July 2014
determined to study the dissolution kinetics. In vivo pharmacokinetic studies were conducted in Accepted 11 July 2014
rabbits. Accelerated stability studies were performed for tablets at 40 ± 2  C/75% RH ± 5% for
6 months.
Results: FTIR studies confirmed no discoloration, liquefaction and physical interaction between
naringin and drug. The results indicated that tablets prepared from naringin presented a
dramatic release (82%) in 30 min with a similarity factor (76.18), which is most likely due to the
amorphous nature of drug and the higher micromeritic properties of blends. Our findings
noticed 1.7-fold increase in oral bioavailability of tablet prepared from naringin with mean Cmax
and AUC0–12 h values as 35.81 ± 0.13 mg/mL and 0.14 ± 0.09 mg h/mL, respectively. The tablets
with and without naringin prepared by DC technique were physically and chemically stable
under accelerated stability conditions upon storage for 6 months.
Conclusion: These results are attractive for further development of an oral tablet formulation of
CAN through P-gp inhibition using naringin, a natural flavonoid as a pharmaceutical excipient.

Introduction Solid dispersions have been the focus of growing interest


during the last decades to develop poorly soluble compounds
Candesartan cilexetil (CAN) is an esterified prodrug of
like CAN to increase dissolution rate and impact saturation
candesartan and an angiotensin II receptor antagonist. Ester
solubility. In these dispersions, the drug is often present as
hydrolysis during absorption from gastrointestinal tract rapidly
partially amorphous or amorphous substance (Yonemochi
and completely bio-activates CAN to its active parent drug,
et al., 1997, 1999; Mura et al., 2002; Dai et al., 2007).
candesartan. A very low aqueous solubility (0.0003 mg/mL)
Previously, we reported that freeze-drying is the most suitable
is the major drawback of CAN to develop an oral dosage form
preparative technique among various pharmaceutical inter-
for its therapeutic application and efficacy. Incomplete
ventions for preparing amorphous solid dispersions of poorly
absorption from the gastrointestinal (GI) tract is reported due
water-soluble drug, CAN (Gurunath et al., 2014).
to the low solubility of CAN across the physiological pH range.
Together, we demonstrated that, improved aqueous
CAN is classified as a class II drug in the Biopharmaceutics
solubility followed by effective P-gp efflux pumps inhibition
Classification System (BCS) based on its solubility across
in the GI tract, the CAN oral bioavailability can significantly
physiologically relevant pH conditions and absorption char-
be enhanced through freeze-dried solid dispersions using
acteristics (Vijaykumar et al., 2009).
naringin, as a novel pharmaceutical excipient (Gurunath et al.,
2013, 2014).
Recently, Li et al. (2013, 2014) reported the acute and
chronic oral toxicity studies of naringin at a single dose of
Address for correspondence: Mr. S. Gurunath, H.No 2-10-910, Bank
Colony, Subedari, Wadepalli, Warangal 506370, Andhra Pradesh, India. 16 g/kg that did not produce any mortality, abnormal changes
Tel: +91 9966555091. Email: s.gurunath1979@gmail.com in body weights and food consumption, undesirable clinical
DOI: 10.3109/10717544.2014.945017 An in vitro and in vivo evaluation of novel tablet formulation 2125

signs, toxicologically related alterations in hematology, (Vijaykumar et al., 2009) but also its active metabolite,
clinical biochemistry and macroscopic findings upon expos- candesartan is a substrate of P-gp (also known as MDR1,
ure of rats. It is rational to establish that LD50 of rats is greater ABCB1; Zhou et al., 2009). The bioflavonoid, called
than 16 g/kg when treated with single oral dose of naringin. naringin, was identified as an extremely safe; naturally
Therefore, naringin could be considered as reasonably safe occurring novel pharmaceutical excipient to show effective
with Hodge and Sterner scale but non-toxic with Gosselin, P-gp-mediated cellular efflux inhibition of P-gp substrate,
Smith and Hodge scale. According to the study, the no candesartan (Lassoued et al., 2012; Gurunath et al., 2013,
observed adverse effect level (NOAEL) of naringin is 2014).
anticipated to be more than 1250 mg/kg/day to Sprague Hence, our present investigation aims:
dawley rats following daily oral administrations for 13 weeks (a) To prepare novel tablet formulation by DC method with
or 6 months. and without naringin to combine the two systems
P-gp-mediated drug efflux significantly intervenes the (improved poor water-solubility through amorphization
delivery of the drugs that requires a very small dose for their and effective P-gp efflux pump inhibition) to one system.
pharmacological actions or the drugs that show extremely (b) To evaluate the formulation blends for micromeritic
slow diffusion and dissolution rates. Those small amounts of properties.
drugs cannot reach the blood circulation in adequate quantity (c) To evaluate the physical characteristics of directly
as it decreases drug absorption and sometimes, that can be life compressed tablets with and without naringin.
threatening (Amin, 2013). (d) Pharmacokinetic evaluation of finished tablets with and
This impact on the bioavailability has been recognized in without naringin in rabbits.
research using mdr1a/1b knockout mice, where the absorption (e) To perform accelerated stability studies for tablets with
of paclitaxel, a substrate of P-gp, was greater than that in and without naringin at 40 ± 2  C/75% RH ± 5% upon
wild-type mice (Schinkel et al., 1997; Sparreboom et al., storage for 6 months.
1997; Woo et al., 2003). These findings established that about
half of paclitaxel dose administered is expelled out into the Materials and methods
gut lumen by P-gp and simply a small amount of drug is lost
by gut wall and liver metabolism. Thus, the effective Candesartan cilexetil (of purity more than 99%, melting
inhibition P-gp-mediated efflux can significantly improve temperature ¼ 163 ± 0.5  C) was kindly obtained as a gift
the oral bioavailability of paclitaxel. sample from the Dr. Reddy’s Labs (Hyderabad, India).
Traditionally, the oral route is the most frequently accepted Polyvinyl pyrrolidone K30 (PVP-K30) was obtained from
route of drug administration. Tablets are considered the most SD Fine Chemicals Ltd., Mumbai, India. Candesartan,
attractive dosage form by patients and industry for drug paracetamol, naringin, spray-dried lactose, Avicel PH 102,
delivery (Allen et al., 2005; Rudnic, 2005; Staniforth, 2007; mannitol, magnesium stearate, talc, Na2HPO4, KH2PO4,
Emery et al., 2009). NaOH, KCl, citric acid NaCl, CaCl2, MgCl2 and NaHCO3
Manufacturing tablets, a chiefly directly compressed tablet were obtained from Sigma Aldrich (Hyderabad, India).
is clear-cut manufacturing process that involves low cost, Acetonitrile, methanol of HPLC grade and orthophosphoric
hence attractive to pharmaceutical laboratories (Strydom acid were supplied by E-Merck India Ltd., Mumbai. All other
et al., 2011). chemicals were of analytical grade. All drug solutions were
The term ‘‘direct compression’’ (DC) is employed to freshly prepared before use.
describe the process where tablets are compressed directly
from the powder blends of active ingredients and suitable Drug–excipient interaction studies
excipients. No pre-treatment by wet or dry granulation is It was imperative to understand the compatibility between the
involved for the powder blends. The DC is more economical, drug under study and the novel excipient used within the
decreases the cycle time of the products and is simple in terms system while designing a tablet formulation. As a result, it is
of the requirements of good manufacturing practices. The DC necessary to conform that drug is not interacting with
is environment-friendly with less number of steps, without excipient for a period of 4 weeks under experimental
involving water and temperature, and finally the stability of conditions (40 ± 5  C and 75 ± 5% RH). The infrared absorp-
the final product can be improved (Gohel & Jogani et al., tion spectra were run between 4000 and 400 cm1 for the pure
2005; McCormick, 2005; Strydom et al., 2011). drug and the physical mixture of drug and excipient. The IR
To the best of our understanding, there are no available spectra of pure drug and a mixture of excipient and drug are
published reports or works demonstrating the effects of shown in Figure 1(a–c).
naringin as P-gp inhibitor as a novel concept of employing
naturally occurring bio-flavonoid as a pharmaceutical excipi-
Formulation of blends
ent with absorption-improving potential formulated as com-
pressed tablets with significant improvement in the Freeze-dried CAN-loaded (1:2 w/w, CAN: PVP-K30) solid
dissolution rate of CAN through amorphous forms. dispersions (Gurunath et al., 2014) and excipients, like
Consequently, our study involves CAN; an ester prodrug naringin, Avicel PH 102, spray-dried lactose, mannitol,
that rapidly de-esterifies in vivo during absorption from the were co-grounded in pestle mortar (without talc and magne-
GI tract to produce its active metabolite, candesartan sium stearate) and were passed through mesh no. 60. Finally,
(McClellan & Goa, 1998) was used as a model drug. CAN talc and magnesium stearate were added and mixed for 5 min
is not only an extremely poor lipophilic compound to obtain formulation blends (Table 1).
2126 G. Surampalli et al. Drug Deliv, 2016; 23(7): 2124–2138

Figure 1. FT-IR spectrum of (a) Candesartan


cilexetil (CAN), (b) Naringin (NGN) and (c)
a physical mixture of pure candesartan
cilexetil and Naringin.

Micromeritic studies of the formulation blends with The tests were performed with an automatic compactor
novel excipient (naringin) (Tapped Density Volumeter, Sisco, India). The bulk density
was determined by gradually pouring the samples into a
Bulk and tapped densities determination of the blends
100 mL graduated glass cylinder up to 60% of the container
Apparent density (b) and tapped density (t) were measured volume. The mass associated with the filled volume (60 mL)
through the apparent volume indirectly (Qiu et al., 2009). was evaluated on an analytical balance. The bulk density was
DOI: 10.3109/10717544.2014.945017 An in vitro and in vivo evaluation of novel tablet formulation 2127
Table 1. Composition of 100 mg tablets (FDS-NT and FDS-T). Tablets preparation by direct compression method
with FSD-NT and without FSD-T naringin
FDS-NT Quantity FDS-T Quantity
(with naringin) (mg) (without naringin) (mg) A preliminary assessment was made to test the ability of the
Freeze-dried solid 24 Freeze dried solid 24 blends to form tablets by DC with and without naringin.
dispersions dispersions An optimized tablet formulations were prepared as listed in
(1:2 w/w, drug to (1:2 w/w, drug to Table 1. Fifty tablets for each formulation were prepared.
carrier ratio) carrier ratio)
Naringin 15 b-Cyclodextrin 15
Table 1 shows the compositions and sample designations of
Spray-dried lactose 18 Spray-dried lactose 18 the tablets. The entire constituents as shown in Table 1 were
Mannitol 18 Mannitol 18 co-grounded for 5–10 min in a glass mortar pestle for each
Avicel PH 102 22 Avicel PH 102 22 formulation separately and were passed through sieve no. 60.
Magnesium stearate 01 Magnesium stearate 01
Talc 02 Talc 02 The mixed blends of excipients were compressed to produce
convex faced tablets weighing 100 mg each, respectively, with
5 mm diameter by means of a single punch tablet machine
(Cadmach, Ahmedabad). The compression force and time
calculated using the relation between the weight (g) and the
were 15 kN and varied from 30 to 34 s for both tablet
apparent volume (mL). The following formula was used to
formulations, respectively.
calculate the bulk density.
The tapped density was obtained by b ¼ VMb tapping a Physical evaluation of tablets
graduated glass cylinder containing a known amount of
samples. The volume after 100 taps (V100) and 500 taps (V500) According to official methods, all tablets (FSD-NT and FSD-
was recorded. In cases where the difference between V100 and T) were evaluated for weight uniformity, friability, thickness
V500 was more than 2%, an additional cycle of 500 taps was and hardness. Twenty tablets were used for weight variation
carried out until variations were less than 2%. determination using an electronic balance (Sartorius, India).
The tapped density (g/mL) was obtained from the mass (g) Twenty tablets (Jadhav et al., 2011) were weighed separately
and the final volume of solids (mL). The following formula and collectively using digital balance.
was used to calculate the tapped density. The average weight of one tablet was obtained by dividing
the total sum of 20 tablets. The weight variation test is an
M acceptable method for determining the drug content uniform-
t ¼
Vt ity of tablets. For an average weight equal to 130 mg or less,
the maximum of 10% weight variation is considered satis-
Hausner ratio and Carr’s index determination factory. Ten tablets were used to determine the tablet hardness
using a Monsanto tablet hardness tester (Campbell
The Hausner ratio (HR) is the association between t and b Electronics, Mumbai, India). About 40 N (1 kg ¼ 10 N),
and is linked to inter-particle resistance. It can be used to tablet breaking force is considered adequate for mechanical
evaluate the powder flow properties. The Carr index (CI%) is stability. Tablet friability was calculated as the percentage
expressed as a percentage which is the measure between the weight loss of 10 tablets at 25 rpm for 4 min using a friabilator
apparent and compact densities. The compressibility percent- (Sisco, India). The weight loss should not be greater than 1%
age indirectly provides a concept of the cohesion, size w/w. The disintegration time was determined at 37 ± 0.5  C
uniformity and surface area of powders or their mixtures in purified water with a disintegration tester (Sisco, India)
(Staniforth, 2007). The compressibility index (I) is calculated fitted with sintered disks. The disintegration time is recorded
as follows: for an average of six measurements. Results are shown in
t  b Tables 2 and 3.
I¼  100
t
Dissolution study
where t is the tapped density and b is the bulk density.
Hausner ratio (HR) is an indirect measure of ease of The USP II Paddle dissolution apparatus (TDT-08 L,
powder flow that is calculated as follows. Electrolab, India) was used to perform the dissolution studies.
The tablets were placed in 900 mL phosphate buffer (pH 6.8)
t maintained at 37 ± 0.5  C, stirring at 50 rpm. At periodical
Hr ¼
b intervals of 0, 2, 4, 6, 10, 15, 20, 25 and 30 min, samples were
where t is tapped density and b is bulk density. Smaller collected, filtered through 0.45 mm and replaced with a fresh
Hausner ratio (51.25) signifies improved flow properties than dissolution medium.
higher ones (41.25). Samples withdrawn at predetermined time intervals,
filtered in-line and assayed using an HPLC method.
Chromatographic separation was accomplished using a
Flow rate and the cotangent of angle a analysis
Shimadzu HPLC system (LC-20 A series) equipped with a
The blends flow rate was evaluated with a flow tester from UV detector, micro-volume double plunger pump (10 mL/
the flow time of the sample. The blends (50 g) were poured stroke) and manual injector type model 7725-port sample
through the funnel (12 mm) attached to the stiff equipment. injection valve was employed. The system was controlled
The funnel was opened and the time taken to release the through SCL-10Avp or SCL-10 A software (Conquer
samples was noted (Schüssele & Bauer-Brandl, 2003). Scientific Lab Equipment, San Diego, CA).
2128 G. Surampalli et al. Drug Deliv, 2016; 23(7): 2124–2138

Table 2. Micromeritic properties of formulation blends (Mean ± SD, n ¼ 3).

Parameters
Formulation Bulk Density (g/mL) Tapped Density (g/ml) Hausner’s ratio Carr’s index Time of flow (s) Cotangent of angle a
FDS-NT 0.567 ± 0.034 0.640 ± 0.023 1.13 ± 0.031 11.40 ± 2.442 2.1 ± 0.442 0.042 ± 0.072
FDS-T 0.618 ± 0.029a 0.696 ± 0.039a 1.12 ± 0.023a 11.20 ± 1.837a 2.3 ± 0.837a 0.041 ± 0.057a
a
p40.05 when compared to FDS-NT.

Table 3. Physical properties of FSD-NT and FSD-T tablets (Mean ± SD, n ¼ 3/6).

Parameters
Formulation Thicknessa (mm) Weighta (mg) Friabilitya (%) Hardnessa (kg/cm2) Disintegration time (min)b
FDS-NT 3.028 ± 0.045 100.4 ± 4.24 0.44 ± 0.047 3.8 ± 0.12 5.86 ± 1.4
FDS-T 3.049 ± 0.027c 100.9 ± 3.17c 0.46 ± 0.044c 3.8 ± 0.13c 6.03 ± 4.6c
a
Average of three determinations.
b
Average of six determinations.
c
p40.05 statistically insignificant when compared to FDS-NT.

HPLC analysis was performed under the following condi- Group I: Received tablet (FSD-T) in a randomized
tions: a Shim-pack VP-ODS or equivalent ODS column sequence (100 mg each; p.o.; n ¼ 3) in two occasions with
(Particle size 5 mm, column dimension: I.D 4.6 mm  length wash out period.
150 mm) maintained at room temperature. An isocratic Group II: Received tablet (FSD-NT) in a randomized
mobile phase composed of acetonitrile:methanol (60:40% sequence (100 mg each; p.o.; n ¼ 3) in two occasions with
v/v) was adjusted for pH 6.0 with orthophosphoric acid at wash out period.
ambient temperature and the injection volume was 20 mL at a The prior approval for conducting the experiments in
flow rate of 1 mL/min. The detective wavelength was 255 nm. rabbits was obtained from our Institutional Animal
The mobile phase was filtered through a 0.45 mm filter in a Committee with registration number 1533/PO/a/11/CPCSEA.
Millipore solvent-filtration apparatus before use.
HPLC method was validated prior to assay as previously Sampling procedures
reported (Gurunath et al., 2013), which was carried out for Blood samples (2 mL) were collected at 0.0, 0.25, 0.5, 0.75, 1,
linearity, precision, accuracy, limit of detection and quanti- 1.5, 2, 4, 6, 8 and 12 h of drug administration from a marginal
fication according to US Pharmacopoeia (The United States ear vein into K3 EDTA tubes (J. K. Diagnostics, Rajkot, India)
Pharmacopoeia, 2007) and International Conference on and placed on ice, protected from light. Samples were stored
Harmonization (ICH) guidelines [ICH (Q2R1), 2007] at 20  C until assayed.
The similarity factor (f2) was calculated to compare the
release profiles of two formulations under similar conditions Preparation of plasma samples
of dissolution media, agitation rate, pH and temperature
employed, to rule out or understand the influence of naringin Liquid–liquid extraction technique was employed for all the
on drug release as novel excipient. The similarity between plasma samples. An aliquot of 500 mL plasma mixed with
two dissolution profiles was considered significant when its 25 mL of internal standard (paracetamol, 10 mg/mL) was
value falls between 50 and 100. Drug release studies were spiked with 3 mL of HPLC grade acetonitrile by vortex
carried out in triplicate. mixing for 3 min. The mixture was centrifuged at 10 500 rpm
for 10 min using a Remi-Model RM 12 centrifugal device, the
supernatant was filtered through 0.45 mm filter (Millipore)
Pharmacokinetic study
and the organic supernatant was evaporated to dryness at
Drug administration room temperature. The residue was reconstituted with 200 mL
of mobile phase, then the solution was centrifuged at
Six male (2.8–3.1 kg) healthy adult albino rabbits were
15 000 rpm for 10 min; finally, an aliquot of 20 mL of the
divided into two groups. Each group consisting of three
solution was injected into the HPLC (Shimadzu, LC- 20 A
rabbits each. Each rabbit was housed individually in separate
series) equipped with a UV detector at a flow rate of 1 mL/
cages under environmentally controlled conditions (25 ± 2  C,
min and the run time was 5 min. The mobile phase comprised
12 h light/dark cycle). Prior to each experiment, the rabbits
60% acetonitrile: 40% methanol at pH 6.0 was determined at
were starved for 24 h and were allowed free access to tap
UV wavelength of 255 nm.
water. Tablets (FSD-NT and FSD-T, each of 100 mg with
5 mm diameter) were orally administered into the stomach of
HPLC analysis
rabbits using a gastric intubation tube (made of silicone
rubber) with one tablet placed on the tip of tube. Care was The chromatograms were acquired using LC solutions
taken to prevent any airway obstruction. The study was software (Shimadzu, Japan) from a Shimadzu HPLC system
conducted as two-periods, two treatments crossover design for (LC-20 A series) equipped with a UV detector. The retention
2 weeks as follows. times for candesartan and internal standard (IS, paracetamol)
DOI: 10.3109/10717544.2014.945017 An in vitro and in vivo evaluation of novel tablet formulation 2129

were approximately 3.93 and 2.85 min, respectively. Linear equivalent to 10 mg of drug (CAN) was weighed and
regression analysis was employed for standard curves dissolved in 10 mL of methanol, filtered using 0.45 mm filter
obtained from drug/internal standard peak area ratios as a paper and an aliquot of 20 mL of the resulting filtrate was
function of plasma concentration versus time data was injected into HPLC to determine the concentration of CAN
analyzed and the oral pharmacokinetic data were developed. from tablet samples. All assays were carried out in triplicate.

Accelerated stability studies Hardness testing


In this study, the accelerated stability studies were performed The hardness or crushing strength of the tablets was measured
for tablets prepared with and without naringin because these using a Monsanto Hardness Tester (Campbell Electronics,
studies are more convenient and less time-consuming over Mumbai, India). Three tablets from each formulation
long-term stability studies, which will take longer time to (FDT-NT and FDT-T; 100 mg each) were tested randomly at
evaluate the physical and chemical stability of the product initial (0), 1, 3 and 6 months, respectively, and the average
under normal conditions of temperature and humidity. reading was noted. The hardness is measured in kg/cm2.
Moreover, tablet formulations containing active pharma-
ceutical ingredient (API) in any high-energy amorphous form X-ray powder diffraction studies
(including solid dispersions) are expected to enhance solu- X-ray powder diffraction (XRPD) analyses were performed
bility, dissolution rate and consequently, oral bioavailability for crushed powder samples of two tablets (FDT-NT and
of poor water-soluble drugs. However, the amorphous regions FDT-T; 100 mg each) at initial (0), 1, 3 and 6 months,
of the dispersed drug in the carrier are thermodynamically respectively, using an X-ray diffractometer (PW 1729, Philips,
unstable and are therefore susceptible to recrystallization Netherlands). The samples were irradiated with monochro-
upon storage (Hecq, 2005). matized Cu Ka radiation (1.5406 Å) and analyzed between 5
Hence, in this study, the prepared tablets (FDT-NT and and 50 2y. The voltage and current used were 40 kV and
FDT-T) were subjected to accelerated stability studies where 40 mA, respectively. The range and the chart speed were
the product is stored under extreme condition of temperature 2  103 CPS and 10 mm/degree 2y, respectively.
and humidity of 40 ± 2  C/75% RH ± 5% RH for 6 months,
respectively.
Degradation studies using HPLC
Tablets (FDT-NT and FDT-T; 100 mg each) were wrapped
in aluminum foil to prevent the formulation from exposure to Three tablets from each formulation (FDT-NT and FDT-T;
light to simulate the aluminum packaging, i.e. Alu–Alu 100 mg each) were tested for degradation studies at initial (0),
packing, of drug products and stored in airtight containers 3 and 6 months, respectively. Each tablet was weighed
which is impermeable to solid, liquid and gases. These packed individually and crushed to a powder in a glass mortar. The
samples of tablets in airtight containers were subjected to powder blend equivalent to 10 mg of pure drug (CAN) was
accelerated stability studies for a period of 6 months as per weighed and dissolved in 10 ml of methanol, filtered using
ICH guidelines [ICH-Q1A (R2) 2003] at the extreme 0.45 mm filter paper and an aliquot of 20 mL of the resulting
temperature and relative humidity conditions of 40 ± 2  C/ solution was injected into the HPLC to observe the peaks
75% RH ± 5% RH in a humidity-controlled oven (Thermolab, corresponding to the pure drug (CAN) and degradation
India). The samples from airtight containers were withdrawn products, respectively. All assays were carried out in
at 1, 3 and 6 months and evaluated for physical appearance, triplicate.
drug content, hardness, disintegration time and in vitro drug
release at 30 min against initial (0 month) values. In vitro dissolution studies
The physical stability was analyzed to monitor for
Drug release studies were carried out in triplicate for two
potential re-crystallization of the solid dispersions in prepared
tablets (FDT-NT and FDT-T; 100 mg each) at initial (0), 3 and
tablets with and without naringin using XRPD studies, while
6 months, respectively, as described above (refer ‘‘Dissolution
the chemical stability was studied with HPLC to evaluate for
study’’ section). The similarity factor (f2) was calculated to
the formation of any degradation product/s under storage
compare the release profiles of two tablets under similar
conditions for 1, 3 and 6 months, respectively. Any physical
conditions of dissolution media, agitation rate, pH and
and chemical changes observed during storage for 6 months
temperature employed, to understand the influence of storage
were compared against initial (0-month) diffractrograms and
conditions on drug release profiles for a period of 6 months.
the peak areas, respectively.
The similarity between two dissolution profiles at initial
(0), 3 or 6 months was considered significant when its value
Drug content determination falls between 50 and 100.
The content uniformity was assessed according to USP
requirements. The test is used to ensure that every tablet Data analysis
contains the amount of drug substance intended with a
Pharmacokinetic analysis
negligible variation among tablets within a batch upon
storage. Three tablets from each formulation (FDT-NT and Plasma concentrations of candesartan in rabbits were plotted
FDT-T; 100 mg each) at initial (0), 1, 3 and 6 months were against time and non-compartmental analysis was used to
tested randomly. Each tablet was weighed individually and determine the pharmacokinetic parameters using Kinetica 5.0
crushed to a powder in a glass mortar. The powder blend Software (Adept Scientific Ltd., Amor Way, Letchworth,
2130 G. Surampalli et al. Drug Deliv, 2016; 23(7): 2124–2138

Garden City, Herts, United Kingdom). The area under the Physical characterization of tablets
plasma concentration versus time curve up to the last
Results of average weight, hardness and friability
quantifiable time point, AUC0–t was obtained by the linear
trapezoidal method. The AUC0–t was extrapolated to infinity Direct compression (DC) method was used due to its ease of
(AUC0–1) by adding the quotient Clast/Kel, where Clast manufacture and lower cost. We optimized (Table 1) the
represents the last measured concentration and Kel represents tablet formulation for 100 mg prepared from naringin for
the apparent terminal rate constant. The half-life of the hardness value, 3.8 kg/cm2, which was significantly (p50.05)
terminal elimination phase was obtained using the relation- greater than the hardness value, 1.5 kg/cm2 for 85 mg tablet
ship t1/2 ¼ 0.693/Kel. The Cmax and Tmax were obtained without naringin. This signifies that naringin formulates the
directly from the data. Oral clearance (Cl) was calculated as tablet harder than the moderate value (1.5 kg/cm2) without
dose divided by AUC0–1. The apparent volume of distribu- naringin. The significant (p50.05) change in the hardness
tion was obtained from the equation Vd ¼ D/(AUC0–1)  Kel. formulated with naringin may be attributed to the glycosidic
Mean residence time (MRT) was determined by division of linkages present in the flavonoid glycoside. However, the
area under the first moment curve (AUMC) by AUC0–1. mechanism is unclear and needs further elucidation.
Since, there is no such official specification specified in
Statistical analysis any pharmacopoeia with reference to hardness of material and
hardness test for uncoated tablets. Formulator is supposed to
Data was expressed as mean ± standard deviation (±SD). decide and maintain the range of hardness during preparation
The statistical significance was determined using based upon requirements in addition to type of formulation.
One-way ANOVA followed by Dunnett’s test or by a However, minimum acceptable hardness of uncoated tablets is
Student’s t-test. Results were considered statistically signifi- 40 N (4.08 kg/cm2) approximately (Banker, 1987).
cant from the control when p50.05 and very significant when Therefore, to equilibrate the hardness (3.8 kg/cm2) of
p50.01. tablets prepared from naringin, hardness of the tablets without
naringin was increased with increased concentration of
Results b-cyclodextrin at constant compression force (15 kN) for
both tablets (FSD-NT and FSD-T) in our preliminary trials.
FTIR spectroscopy
Consequently, the tablets were assessed for their weight
Figure 1(a–c) shows the FT-IR spectra of pure CAN, naringin variation, thickness, disintegration time, hardness and friabil-
and their physical mixture under study. The spectrum of CAN ity. The results in Table 3 reveal that the tablets at hardness
shows prominent peaks at 3410, 2941, 2860,1753, 1716, 3.8 kg/cm2, the disintegration time, friability and thickness
1614, 1548 and 1282 cm1 corresponding to N–H stretching, remained invariable for both tablets (FSD-NT and FSD-T)
C–H aromatic stretching, C–H aliphatic stretching, C ¼ O with no significant change in their weight variation.
stretching in ester, C ¼ O stretching in acid, C ¼ N stretching,
C ¼ C stretching and –C–O–ester. Dissolution study
The IR spectrum of naringin exhibited characteristic IR
Figure 2 features the dissolution release profiles from tablet
bands at around 3408 (broad), 3387 (broad), 2920, 1691 and
formulations (FSD-NT and FSD-T) in the dissolution
1616 cm1 due to more than one O–H aromatic stretching,
medium. In vitro dissolution results showed that tablets
C–H aliphatic stretching, C ¼ O and C ¼ C bonds,
(FSD-NT and FSD-T) provided an immediate release of the
respectively.
drug with no significant differences in their release profiles
The physical mixture showed the superimposed spectra of
under similar conditions of pH medium, agitation rate,
CAN crystals and naringin. The crystalline form of CAN and temperature and dissolution media, respectively.
naringin has sharper and significant peaks at all above-
mentioned prominent regions. No shifting of peaks was
observed with either pure drug or naringin in the physical
mixture spectra. It might be speculated that the no intermo-
lecular bonding was formed between the molecules of two
systems indicating absence of any chemical interaction
between CAN and excipient (naringin).

Evaluation of micromeritic properties


Results of flowability, bulk and tapped density, Hauser’s ratio
and Carr’s index
According to the results presented in Table 2, it is possible to
conclude that the blended powder has flow properties and
compaction performance appropriate for pharmaceutical
applications. The results show that naringin as an excipient
displayed a good flow, in view of the fact that the values of
Figure 2. Dissolution profiles of FSD-NT and FST-T tablets showing
flow time and cotangent of angle a were less than 10 s and cumulative amount of drug release in 900 mL of phosphate buffer
0.1, respectively (Table 2). (pH 6.8). Mean ± SD (n ¼ 3).
DOI: 10.3109/10717544.2014.945017 An in vitro and in vivo evaluation of novel tablet formulation 2131

To describe the release behavior of CAN from the However, there was no change in the time to reach peak
formulations, the data was fitted into a first-order kinetic concentration (tmax) for both formulations (FSD-NT and
model. The tablet formulations (FSD-NT and FSD-T) showed FSD-T, 2 h), which was consistently similar to our previously
a good correlation to first-order kinetics equation (r2 ¼ 0.992) demonstrated pharmacokinetic studies, indicating that cande-
suggesting that the existence of linear correlation between sartan could be absorbed more rapidly after its prodrug, CAN
concentration and ratios of peak areas. was formulated into freeze-dried solid dispersion.
In reality, the release profiles were superimposable under In particular, areas under the concentration–time curves
analogous conditions tested. The similarity factor (f2) was (AUC0–12 h) for FSD-NT (644.03 ± 78.6 mg h/mL, respect-
used to compare the dissolution profiles that reached to a ively) were enhanced by 1.5-fold over the area under the
value of 76.18 in phosphate buffer (pH 6.8) at 50 rpm. concentration–time profiles of FSD-T (428.51 ± 67.8 mg
h/mL; p50.01) paralleled with a reduction in candesartan
Pharmacokinetic study clearance from 3235.64 ± 231.7 to 1878.62 ± 115.7 mL/min.
In vivo studies in rabbits were conducted in order to evaluate In addition, a significant change (p50.01) in the elimin-
whether an immediate plasma concentration profile of ation half life was observed between two tablet formulations
candesartan from the newly compressed tablet formulations (FSD-NT and FSD-T) from 7.15 ± 3.9 to 4.86 ± 2.5 h,
(FSD-NT and FSD-T) as a solid dosage form could be respectively. This revealed that the oral bioavailability of
achieved. candesartan could be dramatically increased in the presence
It has been reported that CAN must be transformed to of naringin. The significantly enhanced oral absorption of
candesartan to achieve antihypertensive effects (Vijaykumar candesartan from FSD-NT was consistent after its direct
et al., 2009). The plasma concentrations of candesartan were delivery to the intestine, which indicated that the intestinal
determined by HPLC method to evaluate the pharmacokinetic absorption was predominant in the improved oral absorption
behavior of FSD-NT against FSD-T. The plasma concentra- of CAN.
tion–time profiles of candesartan after single-dose oral The therapeutic effect of CAN occurred at 4–6 h after the
administration of each tablet with and without naringin is oral administration of free-CAN tablets (Israili, 2000; Gleiter
presented in Figure 3 and the corresponding pharmacokinetic & Mörike, 2002; Husain et al., 2011). Hence, the more rapid
parameters are summarized in Table 4. The experimental absorption of candesartan from FSD-NT tablet formulation
results showed a significant (p50.001) difference occurred could be helpful in treatment of hypertension or heart failure
between the pharmacokinetic profiles of FSD-NT in com- in clinical therapeutics.
parison to that of FSD-T.
Accelerated stability studies
Taken together, at each time point, the plasma concentra-
tion of candesartan from FSD-NT was higher than those The results of the accelerated stability studies showed, no
measured for the plain FSD-T. These are evidenced by our significant differences in drug content uniformity, hardness,
previous reports (Gurunath et al., 2013, 2014), which showed disintegration time and in vitro dissolution profiles before
high apparent permeability coefficient in vitro and also higher (0-month) and after the storage periods (1, 3 or 6 months) for
Cmax in vivo with freeze-dried CAN solid dispersions in the tablets with and without naringin at extreme condition of
presence of naringin. temperature and relative humidity of 40 ± 2  C/75% RH ± 5%.
The peak concentrations (Cmax) of FSD-NT and FSD-T The stability data of tablets (FDT-NT and FDT-T) are
were found to be 88.44 ± 13.8 and 122.23 ± 34.5 mg/mL, shown in Tables 5–7 for percentage of drug content, hardness
respectively, which was significantly improved (p50.01) over and disintegration time to evaluate the effect of storage for
1.38 when compared to the peak concentration of FSD-T.
Table 4. Pharmacokinetic parameters (Mean ± SD, n ¼ 6) of candesartan
following oral administration of FSD-T and FSD-NT tablets.

Parameters FSD-T (100 mg) FSD-NT (100 mg)


AUC(0–t) mg/mL h 428.51 ± 67.8 644.03 ± 78.6a
AUC(0–in) mg/mL h 515.09 ± 83.4 887.17 ± 93.4a
AUC(extra) mg/mL h 85.58 ± 13.4 243.14 ± 64.8a
AUMC (0–t) mg/mL h  h 1792.04 ± 180.4 2816.16 ± 217.8a
AUMC (0–in) mg/mL h  h 3438.48 ± 234.3 8243.05 ± 545.7a
AUMC (extra) mg/mL h  h 1646.43 ± 169.2 5426.89 ± 287.8a
Cmax (mg/mL) 88.44 ± 13.8 122.23 ± 34.5a
Tmax (h) 2.0 ± 1.4 2.0 ± 1.2
t1/2 (h) 4.86 ± 2.5 7.15 ± 3.9a
Vd (L) 1.36 ± 0.8 1.16 ± 0.9a
Clearance (mL/min) 3235.64 ± 231.7 1878.62 ± 115.7a
MRT (h) 6.67 ± 2.7 9.29 ± 4.8a
K (h  1) 0.142 ± 0.31 0.09 ± 0.42a
RB (%) – 66.5

Data are presented as mean ± SD (n ¼ 6). Cmax, maximum plasma


Figure 3. Plasma concentration–time profiles (mean ± SD, n ¼ 6) of concentration; Tmax, time to reach peak concentration; AUC, area
candesartan following oral administration of FSD-NT and FSD-T tablet under the plasma concentration–time curve; t1/2, elimination half-life.
a
(100 mg) formulations. Significant at p50.01 compared to FSD-T.
2132 G. Surampalli et al. Drug Deliv, 2016; 23(7): 2124–2138

Table 5. Percentage of drug content for FDT-T and FDT-NT tablets from accelerated stability studies (Mean ± SD, n ¼ 3).

Drug content (%)


Parameter
Initial 1st-month at 3rd-month at 6th-month at
Formulation (0-month) 40 ± 2  C/75% RH ± 5% RH 40 ± 2  C/75% RH ± 5% RH 40 ± 2  C/75% RH ± 5% RH
FDT-T 98.4 ± 0.44 99.2 ± 0.28a 97.4 ± 0.96a 97.1 ± 1.75a
FDT-NT 98.7 ± 1.83 99.7 ± 0.60a 98.3 ± 0.61a 97.8 ± 1.67a
a
Average of three determinations with p40.05 statistically insignificant when compared to initial (0-month) values.

Table 6. Hardness of FSD-T and FSD-NT tablets from accelerated stability studies (Mean ± SD, n ¼ 3).

Hardness (kg/cm2)
Parameter
Initial 1st-month at 3rd-month at 6th-month at
Formulation (0-month) 40 ± 2  C/75% RH ± 5% RH 40 ± 2  C/75% RH ± 5% RH 40 ± 2  C/75% RH ± 5% RH
FDT-T 3.80 ± 0.12 4.01 ± 0.62a,b 4.08 ± 0.32a,b 4.05 ± 0.52a,b
FDT-NT 3.80 ± 0.13 3.81 ± 0.43a 3.83 ± 0.73a 3.85 ± 0.23a
a
Average of three determinations with p40.05 statistically insignificant when compared to initial (0-month) values.
b
p50.01 when compared to initial (0-month) values.

Table 7. Disintegration time of FSD-T and FSD-NT tablets from accelerated stability studies (Mean ± SD, n ¼ 3).

Disintegration time (min)


Parameter
Initial 1st-month at 3rd-month at 6th-month at
Formulation (0-month) 40 ± 2  C/75% RH ± 5% RH 40 ± 2  C/75% RH ± 5% RH 40 ± 2  C/75% RH ± 5% RH
FDT- T 5.86 ± 1.4 6.95 ± 1.12a,b 7.12 ± 1.36a,b 7.65 ± 1.67a,b
FDT-NT 6.03 ± 1.6 6.17 ± 1.17a 6.26 ± 1.42a 6.12 ± 1.63a
a
Average of three determinations with p40.05 statistically insignificant when compared to initial (0- month) values.
b
p50.01 when compared to initial (0-month) values.

6 months. Percentage drug content of two tablets with and compared to initial (0-month) profiles in phosphate buffer
without naringin were found within the range of 97.1–99.7% (pH 6.8) at 50 rpm.
which meets the USP guidelines as shown in Table 5. The No significant effect on the dissolution profiles of tablets
results in Table 6 reveal no significant (p40.05) difference in without naringin (FDT-T) was observed due to the changes in
hardness values (3.8 kg/cm2) of tablets prepared from hardness and disintegration time upon storage for 6 months.
naringin (FSD-NT) as compared to initial 0-month. This was clearly evidenced from the dissolution profiles of
However, for tablets without naringin (FDT-T), there was FDT-T tablets (Figure 4a) with no significant (p40.05)
significant difference in hardness (p50.01) values at tem- changes as compared to initial (0-month) release profile with
perature and relative humidity of 40 ± 2  C/75% RH ± 5%, the similarity factors (f2) as 65.23 and 71.43 for 3 and 6
respectively, wherein hardness was increased with time, months, respectively.
prolonging the disintegration time of the tablets (Table 7); In addition, the Figure 5(a,b) depicts the XRPD patterns of
but, in all cases, disintegration time was within the specified pure drug (CAN), hydrophilic polymer (PVP-K30), FDT-NT
limit (within 15 min) of official monographs for uncoated and FDT-T tablets, respectively, to evaluate the effect of
tablets (Indian Pharmacopoeia, 2007). storage at 40 ± 2  C/75% RH ± 5% for different time periods.
Figure 4 features the dissolution release profiles from The diffraction pattern of pure drug (CAN) exhibited intensity
tablets (FDT-NT and FDT-T) in the dissolution medium at peaks at 2y values of 9.8 , 9.9 , 10 , 10.1 , 11.6 , 17.2 ,
initial (0), 3 and 6 months, respectively. In vitro dissolution 20.3 , 23.3 , 24.9 , 27.7 and 29.2 (Figure 5). The spectrum
results showed that tablets (FDT-NT and FDT-T) provided an of PVP-K30 exhibited complete absence of any diffraction
immediate release of the drug with no significant differences peaks, which is characteristic of an amorphous compound
in their release profiles upon storage for 6 months as showing the XRPD halo (Figure 5).
compared to initial (0-month) release under the similar Neither FDT-NT nor FDT-T showed any signs of
conditions of pH medium, agitation rate, temperature and re-crystallization during the storage period as depicted in
dissolution media, respectively. Figure 5(a,b). The diffractrograms of tablets (FDT-NT and
In fact, the release profiles of initial (0), 3 or 6 months FDT-T) at initial (0-month), 1, 3 and 6 months showed
were superimposable under analogous conditions tested. The peaks similar to PVP-K30 and absence of any major
similarity factor (f2) was used to compare the dissolution diffraction peaks corresponding to pure drug (CAN) indicat-
profiles for 3 and 6 months, respectively, that reached to a ing that the drug is present in an amorphous state, which is in
value of 65.23 (3-month; FDT-T); 71.43 (6-month; FDT-T) agreement with our previously published report (Gurunath
and 75.78 (3-month; FDT-NT); 76.86 (6-month; FDT-NT) as et al., 2014).
DOI: 10.3109/10717544.2014.945017 An in vitro and in vivo evaluation of novel tablet formulation 2133

Figure 4. Dissolution profiles of (a) FSD-T and (b) FSD-NT tablets showing cumulative amount of drug release in 900 mL of phosphate buffer (pH 6.8)
over storage for 3 and 6 months, respectively (Mean ± SD, n ¼ 3).

The HPLC results in Figure 6 showed no peaks other than than 23% and up to 35% are acknowledged as poor flow
those corresponding to pure drug (CAN) at retention time materials.
(3.95 ± 0.04 min) for tablets (FDT-NT and FDT-T) upon The Hauser’s ratio (HR) obtained as a result of the ratio
storage for 6 months. The observed peak areas of both tablets between the bulk density and compacted density of powders is
corresponding to pure drug (CAN) upon storage for 6 months correlated with the cohesion and adhesion forces of particles.
were similar to those of their initial (0-month) peak areas with Hausner demonstrated that powders such as coarse spheres,
similar retention time values. had ratios of approximately 1.25 with low inter-particle
friction, whereas Hausner ratios greater than 1.5 showed more
Discussion
cohesive, less free-flowing powders such as flakes.
Candesartan cilexetil (CAN), an anti-hypertensive agent, is Therefore, it is established that values lower than 1.25 are
widely indicated for the treatment of hypertension or heart related to a good flow, whereas values higher than 1.25 are
failure in clinical therapeutics. Data available on its poor related to a poor flow, in view of the fact that flow and
solubility, limited intestinal absorption and disposition cohesive properties are inversely proportional (Staniforth,
showed encouraging results with variable absorption, poor 2007; Ofori-Kwakye et al., 2010).
efficacy and bioavailability profiles. Numerous factors such The Carr’s index and Hausner ratio based on the
as low aqueous solubility, poor dissolution properties and cohesiveness between the particles only reflect the potential
poor apparent permeability due to intrinsically low absorptive for consolidation of powders. To determine the ease and speed
membrane permeability contribute to the low and erratic oral of flow, it is imperative to determine the flow rate by means
bioavailability of drugs. More recently, secretory membrane of the time and cotangent of a angle.Since the values of flow
transporters (e.g. P-gp and BCRP) have also been implicated time and cotangent of angle a were less than 10 s and 0.1,
in controlling the oral bioavailability and variability of drug respectively, the formulation blends exhibited good flow
absorption. As a result, novel formulations of CAN exhibiting properties.
better oral absorption and bioavailability need to be Hardness is an element of interest to the compression
developed. process that can be used for equipment adjustment and
Our FT-IR drug-excipient interaction studies showed calibration to obtain the quality of the finished product. In
no discoloration, liquefaction and physical interaction general, higher the compression force, greater will be the
between the drug and the novel excipient used. No consolidation and deformation of the powder in the diet. This
significant shift in the peak was observed which revealed will results in the greater the contact points in the material
that both the drug and excipient are compatible with each being compacted with the low porosity of the tablet. There
other. The FTIR spectra of drug, excipient and mixture of will be an inclination to increase the hardness and disinte-
drug and excipient are shown in the Figure 3(a–c), gration time.
respectively. Alternatively, low hardness will be an indirect measure of
The formulation blends illustrated good packing and the inefficiency of the process of consolidation and compac-
flowability as indicated by Hausner’s ratio value and Carr’s tion of the powders in tablets. Harder tablets will have inferior
index. The Carr’s compressibility index can be used as an friability with a difficulty to be ejected from the press (Kuentz
indication to produce a uniform blend that predicts the flow & Leuenberger, 2000; Armstrong, 2007; Staniforth, 2007;
properties. Compressibility index values that fall between Niazi, 2009).
5 and 15% indicate excellent flow whereas those fall between In our preliminary trials, the ability of b-cyclodextrin as
12 and 16% represent good flow. Conversely, values greater effective diluent (Late & Banga, 2010) in place of naringin to
2134 G. Surampalli et al. Drug Deliv, 2016; 23(7): 2124–2138

Figure 5. (a) XRPD diffractograms of single components of CAN and PVP-K30 together with tablets prepared (a) without naringin–FSD-T at initial
(0 month), FSD-T at 1st month, FSD-T at 3rd month and FSD-T at 6th month, respectively, upon storage for 6 months at 40 ± 2  C/75% RH ± 5% and
(b) with naringin–FSD-NT at initial (0 month), FSD-NT at 1st month, FSD-NT at 3rd month and FSD-NT at 6th month, respectively, upon storage for
6 months at 40 ± 2  C/75% RH ± 5%.
DOI: 10.3109/10717544.2014.945017 An in vitro and in vivo evaluation of novel tablet formulation 2135

Figure 6. Typical HPLC chromatograms of FSD-T and FSD-NT upon storage for 6 months at 40 ± 2  C/75% RH ± 5%.

formulate tablets containing freeze-dried CAN-loaded out or understand the influential role of naringin on drug
solid dispersions prepared by the DC was demonstrated release.
in order to equilibrate the hardness of FSD-NT to that of Besides, the increase in the oral bioavailability of hydro-
FSD-T tablets at constant compression force and time to rule phobic drugs of low water solubility using water-soluble
2136 G. Surampalli et al. Drug Deliv, 2016; 23(7): 2124–2138

excipients, these diluents (naringin or b-cyclodextrin) can also improved through wetting of drug particles and localized
act as fillers for such tablets. solubilization by hydrophilic carriers.
The analogous disintegration time values for FSD-NT and In addition, there was also concomitant decrease in plasma
FSD-T formulations may also attribute to the superior water clearance and volume of distribution with increased the
solubility nature of naringin (i.e. 1 mg of naringin dissolved in plasma concentration and AUC as shown in Figure 3 and
mL of water) due to the hydroxyl groups and glycosidic Table 1, with FSD-NT tablet formulation. The small volume
linkages present in the flavonoid and hydrophilic nature of of distribution, low oral clearance and increased mean
b-cyclodextrin, respectively. residence time obtained in this study indicated that the drug
Although hardness is not an attribute that can be directly distributes primarily in the extracellular spaces.
correlated with drug release, a greater hardness value may be The above-mentioned facts may attribute to high protein
beneficial for maintaining tablet integrity and controlling binding at physiological pH, as previously reported (Gleiter &
drug delivery. From a comparison of the release profiles, Mörike, 2002; Husain et al., 2011). The significant change in
increasing or equilibrating the hardness of the FSD-T to that MRT and Kel suggests that the increased time for which the
of FSD-NT using b-cyclodextrin facilitated the release of drug remains in the body.
CAN, most likely because the b-cyclodextrin is a hydrophilic The relative absorption (RA) is defined as the AUC0–t
polymer, allowing the penetration of the dissolution medium values of FSD-NT compared to that of FSD-T, which was
and thus showing similarity in the rate of drug release to that relatively higher for FSD-NT as depicted in Table 2. The
of tablet formulation with naringin (FSD-NT). increased rate of absorption could be attributed to the
Both FSD-NT and FSD-T tablets offered dramatic concomitant and synergistic inhibition of P-gp present in
improvements in the rate as well as the extent of drug release the intestine as a plausible explanation for prominent increase
(Figure 5) and presented the highest (80%) drug release in in oral bioavailability of CAN.
30 min. In the early 10-min period, there were about 54 and Accelerated stability studies for the tablet formulations
57% increase of drug release from both tablets, respectively. (FDT-NT and FDT-T) did not show any significant changes in
This improved drug release attributed to the presence of their physical appearance, drug contents, disintegration time
amorphous form of CAN, as confirmed by our earlier and in vitro drug release profiles at 30 min (p40.01). Upon
reported DSC and PXRD studies (Gurunath et al., 2014). 6 months of storage at 40 ± 2  C/75% RH ± 5% RH (Figure 4),
Other factors such as increased wettability, solubilization of the tablets (FSD-NT and FSD-T) demonstrated similar
the drug by the carrier at the diffusion layer, the reduction or dissolution profiles as compared to that of initial 0-month
absence of aggregation, agglomeration increased the rate release profiles indicating that CAN was stable in the
of dissolution and higher micromeritic properties exerted by presence of PVP-K30 and other excipients like naringin.
excipients might have been contributed to improved drug Tablets prepared from naringin showed no significant
release from these tablet formulations. This can clearly be changes in the hardness and disintegration time as observed
evidenced by superimposable drug release profiles with for tablets without naringin. Although significant changes
similarity factor value between 50 and 100 (76.18). were noticed in the hardness and disintegration time of FDT-T
Based on all of the above findings, the possible mechan- tablets probably due to the loss of moisture, but no significant
isms for an increased release rate from both tablets have been change was observed in their dissolution profiles upon storage
postulated and are summarized as follows: (i) reduction of for 6 months indicating that both formulations (FSD-NT and
crystallinity, (ii) increased water absorption by the hydro- FSD-T) were fairly stable at storage time and conditions.
philic carrier and excipients which leads to increased Figure 5 represents XRPD diffractrograms of tablet
wettability, (iii) solubilization effect by the carrier and formulations (1, 3 and 6 months over storage at 40 ± 2  C/
excipients, (iv) phase transition from crystalline to amorphous 75% RH ± 5% RH) showing no signs of re-crystallization in
nature and (v) higher micromeritic properties exerted by the samples. Hence, there was no influence of storage time as
naringin or b-cyclodextrin. well as storage conditions on the XRPD patterns of the drug
The results of the in vivo pharmacokinetic study demon- in the samples of tablets, exhibiting similar XRPD patterns
strated that rabbits exhibited the moderate inter-individual that overlay the initial (0-month) patterns, thus retaining the
variation in the plasma concentrations of candesartan, which amorphous nature of the drug.
may be reflective of absorption, metabolism and elimination Furthermore, the results of the chemical stability by HPLC
differences between individuals. As shown in Figure 3, analyses showed no peaks other than those corresponding to
FSD-NT was absorbed significantly from the intestinal pure drug (CAN) at retention time (2.85 ± 0.04 min) for
regions and reached a maximum peak plasma concentration tablets (FDT-NT and FDT-T) upon storage for 6 months
at about 2 h after the drug treatment as compared to FSD-T suggesting that the tablets did not undergo decomposition or
indicating that the presence of naringin contributed signifi- degradation upon storage over 6 months’ study, indicating that
cantly for the enhanced intestinal absorption of candesartan the assay procedure is selective for the analysis of CAN
from its tablet formulation as a P-gp inhibitor. without interference of the excipients. This could be attributed
All the Cmax and AUC0t values of FSD-NT were much to the evolved interactions between CAN and PVP-K30
greater than the Cmax and AUC0t values of FSD-T at the (Barmpalexis et al., 2013) or glycosidic linkages of naringin,
same dose of 100 mg, which may explain that the dissolution which requires further elucidation.
of CAN along with effective P-gp inhibition significantly Oral formulations of CAN having better efficacy and less
contributed to improved intestinal absorption. This is also toxicity could be developed using appropriate P-gp inhibitor.
attributed to the presence of solid state (amorphous) of drug However, further studies are required to compare the
DOI: 10.3109/10717544.2014.945017 An in vitro and in vivo evaluation of novel tablet formulation 2137

effectiveness of naringin with other novel P-gp inhibitors to Armstrong NA. (2007). Tablet manufacture. In: Swarbrick J, ed.
Encyclopedia of pharmaceutical technology, 3rd ed. New York,
enable pharmacokinetic modulation of candesartan or its USA: Informa Healthcare, 3653–83.
prodrug, CAN. Genetic differences mediated by P-gp in both Banker GS, Anderson NR. (1987). Tablets. In: Lachman L, Lieberman
hepatic metabolism and intestinal transport may be an HA, Kanig JL, eds. The theory and practice of industrial pharmacy.
important contributing factor in the wide inter-individual 3rd ed. Philadelphia, PA: Lea & Febiger, 293–345.
Barmpalexis P, Koutsidis I, Karavas E, et al. (2013). Development of
variability in the absorption and disposition of candesartan. PVP/PEG mixtures as appropriate carriers for the preparation of drug
These results suggest that the administration of CAN in the solid dispersions by melt mixing technique and optimization of
form of tablet dosage form formulated with safe and naturally dissolution using artificial neural networks. Eur J Pharm Biopharm
occurring P-gp inhibitor may improve the therapeutic index of 85:1219–31.
Dai WG, Dong LC, Song YQ. (2007). Nanosizing of a drug/carrageenan
CAN for the treatment of hypertension and other associated complex to increase solubility and dissolution rate. Int J Pharm 342:
complications of hypertension. 201–7.
Emery E, Oliver J, Pugsley T, et al. (2009). Flowability of moist
pharmaceutical powders. Powder Technol 189:409–15.
Conclusion Gleiter CH, Mörike KE. (2002). Clinical pharmacokinetics of cande-
sartan. Clin Pharmacokinet 41:7–17.
The success of DC method in the design of oral dosage forms Gohel MC, Jogani PD. (2005). Review of co-processed directly
requires the study of the physical properties of the formula- compressible excipients. J Pharm Pharm Sci 8:76–93.
tion blends. The blends formed more suitable tablets with and Gurunath S, Baswaraj KN, Patil PA. (2013). Oral bioavailability and
without naringin, as a result of improved flow and compress- intestinal absorption of candesartan cilexetil: role of naringin as
p-glycoprotein inhibitor. Drug Dev Ind Pharm (Early Online: 1–7).
ibility of the novel excipients. The compression time, force Available at http://www.informapharmascience.com/doi/abs/ 10.3109/
and amount of excipients used did not influence friability, 03639045.2013.850716.
hardness and drug release. The dissolution profiles demon- Gurunath S, Baswaraj KN, Patil PA. (2014). Enhanced solubility and
strated that the proposed formulations produced more than intestinal absorption of candasartan cilexetil solid dispersions using
everted rat intestinal sacs. Saudi Pharm J 22:246–57.
80% of drug release within 30 min. No change in dissolution Hecq J, Deleers M, Fanara D, et al. (2005). Preparation and character-
profiles and release kinetics were noticed between two ization of nanocrystals for solubility and dissolution rate enhancement
formulations, most likely due to the similarity in nature of of nifidipine. Int J Pharm 299:167–77.
the excipients. Another improvement that can be associated Husain A, Azim MS, Mitra M, Bhasin PS. (2011). A review on
candesartan: pharmacological and pharmaceutical profile. J Appl
with the developed tablets was that no other additional Pharm Sci 1:12–17.
ingredients in the formulations were required since naringin ICH (Q2R1). (2007). Guideline on validation of analytical procedures:
acted as filler in the DC process. Consequently, the current text and methodology. In: Proceedings of the International Conference
results demonstrate that the intestine plays an important role on Harmonization, Geneva.
ICH-Q1A (R2). (2003). ICH harmonised tripartite guideline. Cover Note
in the net absorption and disposition of CAN, which showed for Revision of Q1A(R) Stability Testing of New Drug Substances and
that naringin increased significantly the oral bioavailability of Products.
candesartan, suggesting the involvement of P-glycoprotein in Indian Pharmacopoeia. (2007). Tablet Disintegration Test Apparatus.
CAN disposition. It also revealed that naringin, a Indian Pharmacopoeia Commission, 6th ed. Vol. I:187–9.
Israili ZH. (2000). Clinical pharmacokinetics of angiotensin II (AT1)
P-glycoprotein inhibitor can be used as pharmaceutical receptor blockers in hypertension. J Hum Hypertens 14:S73–86.
excipient for developing oral dosage forms to increase Jadhav SB, Kaudewar DR, Kaminwar GS, et al. (2011). Formulation and
intestinal permeability and in vivo pharmacokinetic perform- evaluation of dispersible tablets of diltiazem hydrochloride. Int J
ance of CAN to enhance oral bioavailability. Also, under the Pharm Tech Res 3:1314–21.
Kuentz M, Leuenberger H. (2000). A new theoretical approach to tablet
accelerated stability conditions evaluated in this study, tablets strength of a binary mixture consisting of a well and a poorly
prepared with and without naringin involving CAN solid compactable substance. Eur J Pharm Biopharm 49:151–9.
dispersions prepared by DC process were physically and Lassoued MA, Sfar S, Bouraoui A, Khemiss F. (2012). Absorption
chemically stable over 6 months as compared to initial enhancement studies of clopidogrel hydrogen sulphate in rat everted
gut sacs. J Pharm Pharmacol 64:541–52.
(0-month) physical and chemical parameters. Late SG, Banga AK. (2010). Response surface methodology to optimize
However, due to the complexity of the in vivo model, the novel fast disintegrating tablets using b-cyclodextrin as diluent. AAPS
role of P-gp and other putative secretory efflux transporters is PharmSciTech 11:1627–35.
not yet clear and needs to be further elucidated. Hence, Li P, Wang S, Guan X, et al. (2014). Six months chronic toxicological
evaluation of naringin in Sprague–Dawley rats. Food Chem Toxicol
further studies are required for comprehensive and multi- 66:65–75.
disciplinary evaluation of various claims to make effective Li P, Wang S, Guan X, et al. (2013). Acute and 13 weeks subchronic
use of these products. toxicological evaluation of naringin in Sprague-Dawley rats. Food
Chem Toxicol 60:1–9.
McClellan KJ, Goa KL. (1998). Candesartan cilexetil: a review of its use
Declaration of interest in essential hypertension. Drugs 56:847–69.
McCormick D. (2005). Evolution in direct compression. Pharm Technol
Authors declare no conflicts of interest in the publication of 4:52–62.
this research work. Mura P, Cirri M, Faucci MT, et al. (2002). Investigation of the effects of
grinding and co-grinding on physicochemical properties of glisentide.
J Pharm Biomed Anal 30:227–37.
References Niazi SK. (2009). Handbook of pharmaceutical manufacturing formu-
lations: compressed solid products, 2 ed. Vol. 1. New York, USA:
Allen LV, Popovich NG, Ansel HC. (2005). Ansel’s pharmaceutical Informa Health Care, 62–81.
dosage forms, 8th ed. Baltimore, USA: Lippincot Williams & Wilkins. Ofori-Kwakye K, Asantewaa Y, Kipo SL. (2010). Physicochemical and
Amin L. (2013). P-glycoprotein inhibition for optimal drug delivery. binding properties of cashew tree gum in metronidazole tablet
Drug Target Insights 7:27–34. formulations. Int J Pharm Pharm Sci 2:105–9.
2138 G. Surampalli et al. Drug Deliv, 2016; 23(7): 2124–2138

Qiu Y, Chen Y, Zhang GGZ, et al. (2009). Developing solid oral dosage Strydom SJ, Ottoa DP, Liebenberg W, et al. (2011). Preparation and
forms: pharmaceutical theory & practice. 1st ed. USA: Elsevier, characterization of directly compactible layer-by-layer nanocoated
163–70. cellulose. Int J Pharm 404:57–65.
Rudnic E. (2005). Oral solid dosage forms. In: Gennaro AR, ed. The United State Pharmacopoeia 30/NF 25. (2007). Asian Ed: the
Remington: the science and practice of pharmacy. 21st ed. official compendia of standard. Rockville: The United States
Philadelphia, PA: Lippincott Williams & Wilkins, 1615–49. Pharmacopoeial Convection Inc., 621–720.
Schinkel AH, Mayer U, Wagenaar E, et al. (1997). Normal viability Vijaykumar N, Raviraj P, Venkateshwarlu V, Harisudhan T. (2009).
and altered pharmacokinetics in mice lacking mdr1-type Development and characterization of solid oral dosage form
(drug-transporting) P-glycoproteins. Proc Natl Acad Sci USA 94: incorporating candesartan cilexetil. Pharm Dev Technol 14:290–8.
4028–33. Woo JS, Lee CH, Shim CK, Hwang SJ. (2003). Enhanced oral
Schüssele A, Bauer-Brandl A. (2003). Note on the measurement of bioavailability of paclitaxel by coadministration of the P-glycoprotein
flowability according to the European pharmacopoeia. Int J Pharm inhibitor KR30031. Pharm Res 20:24–30.
257:301–4. Yonemochi E, Kitahara S, Maeda S, et al. (1999). Physicochemical
Sparreboom A, van Asperen J, Mayer U, et al. (1997). Limited oral properties of amorphous clarithromycin obtained by grinding and
bioavailability and active epithelial excretion of paclitaxel (Taxol) spray drying. Eur J Pharm Sci 7:331–8.
caused by P-glycoprotein in the intestine. Proc Natl Acad Sci USA 94: Yonemochi E, Ueno Y, Ohmae T, et al. (1997). Evaluation of amorphous
2031–5. ursodeoxycholic acid by thermal methods. Pharm Res 14:798–803.
Staniforth J. (2007). Powder flow. In: Aulton M, ed. Aulton’s pharma- Zhou L, Chen X, Gu Y, Liang J. (2009). Transport characteristics of
ceutics: the design and manufacture of medicines. 3rd ed. Canada: candesartan in human intestinal Caco-2 cell line. Biopharm Drug
Elsevier, 152–80. Dispos 30:259–64.

You might also like