Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Materials Science and Engineering C 75 (2017) 25–33

Contents lists available at ScienceDirect

Materials Science and Engineering C

journal homepage: www.elsevier.com/locate/msec

A new bioavailability enhancement strategy of curcumin via


self-assembly nano-complexation of curcumin and bovine
serum albumin
Hong Yu a, Minh-Hiep Nguyen b, Wean Sin Cheow c, Kunn Hadinoto a,⁎
a
School of Chemical and Biomedical Engineering, Nanyang Technological University, 637459, Singapore
b
Faculty of Applied Science, Ton Duc Thang University, Ho Chi Minh City, Vietnam
c
Singapore Institute of Technology, 138683, Singapore

a r t i c l e i n f o a b s t r a c t

Article history: Amorphous drug nanoparticles have recently emerged as a superior bioavailability enhancement strategy for
Received 5 October 2016 poorly soluble drugs in comparison to the conventional microscale amorphous solid dispersions. In particular,
Received in revised form 16 December 2016 amorphous drug nanoparticle complex (or nanoplex) represents an attractive bioavailability enhancement strat-
Accepted 6 February 2017
egy of curcumin (CUR) - a medicinal herb known for its wide-ranging therapeutic activities - attributed to the
Available online 10 February 2017
high payload, cost-effective preparation, and supersaturation generation of the nanoplex. To address the poor
Keywords:
colloidal stability of conventional nanoplex formulations, we herein developed a new class of CUR nanoplex by
Drug-protein complexation complexation of CUR with bovine serum albumin (BSA). The effects of two key variables in drug-protein com-
Albumin nanoparticles plexation, i.e. pH and mixing ratio (MBSA/CUR), on the physical characteristics and preparation efficiency were in-
Therapeutic proteins vestigated. While the CUR-BSA nanoplex preparation was found to favor acidic pH and MBSA/CUR below unity, the
nanoplex's physical characteristics were minimally affected by pH and MBSA/CUR. At the optimal condition, CUR-
BSA nanoplex with size ≈90 nm, zeta potential ≈27 mV, and payload ≈70% were produced at nearly 100% CUR
utilization rate and ≈80% yield. The nanoplex produced a prolonged supersaturation level at ≈9× of the satura-
tion solubility for 4 h. The dissolution rate could be modulated by thermal treatment of the nanoplex post its
preparation. The long-term amorphous state stability, storage colloidal stability, and preserved bioactivity of
the nanoplex were successfully established. Lastly, the CUR-BSA nanoplex was found to be superior to the con-
ventional nanoplex in its size, supersaturation generation, colloidal stability, and yield.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction bioavailability caused by its poor solubility in the gastrointestinal fluid


and rapid degradation at the intestinal pH [3]. While nanoencapsulation
The therapeutic properties of curcumin (CUR) - a natural flavonoid ex- has been extensively investigated as a bioavailability enhancement strat-
tracted from turmeric plants - have been well established ranging from egy of CUR, CUR nanocapsules exhibit two major drawbacks in (a) their
anti-inflammatory and antimicrobial to anticancer and anti-diabetic [1]. costly and intricate preparation, and (b) low CUR payload [4,5].
The importance of CUR as therapeutics was also evident from the exten- To address these issues, we previously developed amorphous CUR-
sive research efforts in the development of novel nanomaterials for accu- chitosan nanoparticle complex (or nanoplex in short) exhibiting a
rate determination of the amount of CUR in the plasma [2]. The high payload (N 80%) prepared by a simple, highly efficient, cost-
effectiveness of CUR therapy, however, is limited by its low oral effective method based on drug-polysaccharide complexation [6].
Most importantly, the nanoplex was capable of producing high appar-
ent solubility of CUR upon dissolution as a result of the supersaturation
Abbreviations: AA, acetic acid; BSA, bovine serum albumin; CE, complexation
generation of its amorphous state, thus making it an ideal bioavailability
efficiency; CFU, colony forming unit; CHI, chitosan; CUR, curcumin; FESEM, field
emission scanning electron microscope; FTIR, Fourier transform infrared spectroscopy; enhancement strategy [6]. The CUR-chitosan nanoplex, however, exhib-
HPLC, high performance liquid chromatography; HPMC, hydroxypropylmethylcellulose; ited a tendency to agglomerate in its aqueous suspension form during
MBSA/CUR, mixing ratio of BSA to CUR (w/w); MCHI/CUR, mixing ratio of CHI to CUR (w/w); short-term storage (i.e. after 6 h), hence necessitating its immediate
MHB, Mueller Hinton broth; MIC, minimum inhibitory concentration; MW, molecular transformation to dry powders [6].
weight; OD600, optical density at 600 nm; PBS, phosphate buffered saline; PCS, photon cor-
relation spectroscopy; pI, isoelectric pH; UV–vis, ultraviolet visible.
Herein we hypothesized that the use of proteins, in place of poly-
⁎ Corresponding author. saccharides, as the complexation partner for CUR could improve the
E-mail address: kunnong@ntu.edu.sg (K. Hadinoto). colloidal stability of the resultant CUR nanoplex. Bovine serum

http://dx.doi.org/10.1016/j.msec.2017.02.018
0928-4931/© 2017 Elsevier B.V. All rights reserved.
26 H. Yu et al. / Materials Science and Engineering C 75 (2017) 25–33

albumin (BSA) was used as the model protein because of its low cost, nanoplex produced, from which the optimal preparation condition was
biodegradability, non-toxicity, and wide acceptance for pharmaceu- determined. Subsequently, the second objective was to analyze and com-
tical uses [7–9]. Most importantly, several studies have reported the pare the (1) physical characteristics, (2) preparation efficiency, (3) colloi-
ability of BSA to provide excellent steric and electrostatic stabiliza- dal stability, (4) dissolution rate, and (5) supersaturation generation
tion of nanoparticles by virtue of the presence of adsorbed protein capability of the CUR-BSA nanoplex prepared at the optimal condition,
corona on the nanoparticle's surface [10–12]. Furthermore, BSA has with those exhibited by the conventional CUR-chitosan nanoplex. In addi-
been known to possess high binding capacity for various drugs, in- tion, we assessed whether the bioactivities of CUR were adversely affect-
cluding CUR and other flavonoids [13,14], where CUR binding with ed by its complexation with BSA by using antimicrobial assay as the
serum albumins (including BSA) has been found to improve the representative bioactivity test among the many bioactivities of CUR.
chemical stability of CUR in physiological condition [15–17]. Lastly, the third objective of the present work was to equip the CUR-
The schematic of the CUR-BSA nanoplex formation is presented in BSA nanoplex with modulated release functionality achieved by post-
Fig. 1. The high affinity of CUR to BSA attributed to the hydrophobic preparation thermal treatment of the nanoplex. As proteins, BSA un-
binding of CUR to the hydrophobic cavities of BSA [18] led to the forma- folds upon heating causing the disruption of its compact helical struc-
tion of soluble CUR-BSA complex. Depending on the preparation pH, the tures [21,22]. We hypothesized that the conformational change of BSA
electrostatic binding between BSA and the oppositely charged CUR in the nanoplex after the thermal treatment would enable us to modu-
could also play a role in the CUR-BSA complex formation [16]. Similar late the release of CUR from the nanoplex. For this purpose, the effect of
to the particle formation mechanism in drug-polysaccharide or heating on the nanoplex's structural integrity was examined first in
protein-polyelectrolyte complexation [19,20], aggregates of the soluble order to determine the onset of thermal denaturation of BSA that led
CUR-BSA complex subsequently formed due to the hydrophobic inter- to the structural collapse of the nanoplex.
actions among the bound CUR and BSA molecules (Fig. 1). The aggre-
gates of CUR-BSA complex then precipitated upon reaching a critical 2. Materials and methods
concentration to produce the CUR-BSA nanoplex.
In the present work, the first objective was to investigate the effects of 2.1. Materials
the two key variables in drug-protein complexation (i.e. pH and mixing
ratio of BSA to CUR) on the (1) physical characteristics (i.e. size, zeta po- CUR (≥ 98% purity) and Mueller Hinton broth (MHB) were pur-
tential, payload, and amorphous state) and (2) preparation efficiency chased from Alfa Aesar (USA) and BD Diagnostics (USA), respectively.
(i.e. curcumin utilization rate and production yield) of the CUR-BSA BSA (MW = 69 kDa, ≥ 96% purity), chitosan (CHI) (MW =

Fig. 1. Schematic of CUR complexation with BSA driven by their hydrophobic and electrostatic interactions resulting in the formation of CUR-BSA nanoplex.
H. Yu et al. / Materials Science and Engineering C 75 (2017) 25–33 27

50–190 kDa, 75–85% deacetylation), potassium hydroxide (KOH), gla- (PXRD) patterns from 5° to 70° (2θ) with a step size of 0.02°/s using
cial acetic acid (AA), ethanol, and phosphate buffered saline (PBS) D8 Advance X-ray diffractometer equipped with Cu Kα radiation
(pH 7.4) were purchased from Sigma-Aldrich (Singapore). Clinically de- (Bruker, Germany). The PXRD analysis was performed for the native
rived strain of P. aeruginosa was provided by the clinical microbiology CUR, native BSA, and for the CUR-BSA nanoplex 48-h after six-month
laboratory of National University Hospital (Singapore). storage at 25 °C and 55% relative humidity.

2.2. Methods 2.2.3. Dissolution rate and supersaturation generation


The dissolution time-profile of the nanoplex under sink condition
2.2.1. Preparation of CUR-BSA nanoplex was reported in terms of % Dissolution defined in Eq. (3), where the
Acidic CUR having pKa of 8.4, 9.9, and 10.5 [23] was dissolved in amount of CUR recovered in the dissolution medium represented the
0.1 M KOH (pH 13) at 5 mg/mL to produce negatively charged CUR mol- net difference between the amount of CUR dissolved from the nanoplex
ecules. Amphoteric BSA with pI in the range of 4.8 to 5.6 [24] was dis- and the amount of the dissolved CUR that had undergone hydrolytic
solved in aqueous AA solution at AA concentration between 0.4% (v/v) degradation in PBS [26].
(pH 2.95) and 1.2% (v/v) (pH 2.71) to produce BSA molecules with net
positive charges. Equal volumes of the CUR and BSA solutions were Mass of CUR recovered in the dissolution medium
%Dissolution ¼
then mixed immediately after their preparation to minimize alkaline Mass of CUR in the nanoplex prior to dissolution
degradation of CUR [23], as well as acidic degradation of BSA in strong  100 ð3Þ
acids (pH ≤ 2.0) [25]. The resultant CUR-BSA nanoplex suspension was
then washed by two cycles of centrifugation at 14,000 × g for 25 min, Briefly, 0.1 mL of the nanoplex suspension was placed inside a
followed by its re-suspension in deionized water for characterizations. microtube filled with 1.9 mL PBS, where one tube was used for each
Four independent replicates from different days were used for the time interval investigated. The nanoplex suspension in PBS was then
characterizations. placed in a shaking incubator at 37 °C. At fixed time interval over a
To study the pH effect, the mixing ratio of BSA to CUR (MBSA/CUR) was 30-min-period, 0.5 mL of the nanoplex suspension was withdrawn
fixed at 1.2 (w/w), where a value N1.0 for MBSA/CUR was arbitrarily cho- and filtered after which the amount of the dissolved CUR was deter-
sen to ensure sufficient BSA available for complexation with CUR. Sub- mined by High Performance Liquid Chromatography (HPLC). The
sequently, MBSA/CUR was varied between 0.2 and 2.5 (w/w) at a fixed HPLC analysis was performed using ZORBAX Eclipse Plus C18 column
amount of CUR (i.e. 5 mg/mL) to study the effect of MBSA/CUR at the op- (250 × 4.6 mm, 5 μm particle size) at detection wavelength of
timal pH. For the CUR-CHI nanoplex, the detailed preparation method 423 nm. Aqueous ethanol solution 90% (v/v) was used as the mobile
had been presented in Nguyen, Yu, Kiew and Hadinoto [6], hence the phase at 1 mL/min to produce CUR's retention time of approximately
method was not repeated here for brevity. 2.5 min.
The preparation efficiency was characterized by the complexation The supersaturation generation capability of the nanoplex was char-
efficiency (CE) and the yield defined in Eqs. (1) and (2), respectively. acterized in triplicates by the ratio of the supersaturated concentration
CE representing the CUR utilization rate was determined by taking the generated upon dissolution (C) to the thermodynamic saturation solu-
difference between the initial mass of CUR added and the mass of CUR bility of CUR (CSat) determined previously to be equal to 4.15 μg/mL
recovered in the supernatant after the first centrifugation. The mass of [6]. The supersaturation was generated by adding the nanoplex in ex-
CUR recovered in the supernatant was quantified by UV–vis spectro- cess (i.e. 15 × CSat) to 8.5 mL PBS in a shaking incubator at 37 °C. At a
photometry (UV Mini-1240, Shimadzu, Japan) at 423 nm. The yield fixed time interval over a 24-h-period, 0.2 mL aliquot was withdrawn,
representing the utilization rate of both CUR and BSA was determined filtered, and immediately diluted tenfold with aqueous 80% (v/v) etha-
using the freeze-dried mass of the nanoplex produced after the washing nol solution to prevent CUR precipitation from the supersaturated solu-
step. tion. The amount of CUR in the aliquot was then determined by HPLC
following the same protocols described earlier. The effect of the inclu-
Mass of CUR that formed nanoplex sion of hydroxypropylmethylcellulose (HPMC) as crystallization inhibi-
CE ð%Þ ¼  100 ð1Þ
Initial mass of CUR added tor in the dissolution medium on the supersaturation level (C/CSat) was
investigated at HPMC = 2 mg/mL.
Mass of CUR−BSA nanoplex produced
Yield ð%Þ ¼  100 ð2Þ
Initial mass of CUR and BSA added 2.2.4. Antimicrobial assay and colloidal stability
Antimicrobial activity of the CUR-BSA nanoplex was evaluated from
the minimum inhibitory concentration (MIC) against clinically derived
2.2.2. Physical characterizations of CUR-BSA nanoplex isolate of P. aeruginosa. The MIC was determined by microbroth dilution
The size and zeta potential of the nanoplex were characterized by method according to the Clinical and Laboratory Standards Institute
photon correlation spectroscopy (PCS) using Brookhaven 90 Plus Nano- (CLSI) guidelines [27]. The detailed MIC protocols were provided in
particle Size Analyzer (Brookhaven Instruments Corporation, USA). The the Supplementary materials for brevity. For the nanoplex prepared at
particle morphology was examined by Field Emission Scanning Electron the optimal pH and MBSA/CUR, its colloidal stability during ambient stor-
Microscope (FESEM) (JSM-6700F, JEOL, USA) using the freeze-dried age was examined in triplicates by monitoring the changes in the size
sample of the nanoplex sputter coated with platinum. The payload and zeta potential of the nanoplex at fixed time intervals over a 24-h-
was characterized by dispersing a known mass of the freeze-dried period.
nanoplex in 80% (v/v) aqueous ethanol solution after which the amount
of the dissolved CUR was determined by UV–vis spectrophotometry at 2.2.5. Post-preparation thermal treatment of CUR-BSA nanoplex
423 nm. The payload was defined here as the ratio of the mass of CUR The thermal treatment was performed on the CUR-BSA nanoplex
in the nanoplex to the total mass of the nanoplex. prepared at MBSA/CUR of 0.6, 1.0, and 2.5. Briefly, 1 mL of the nanoplex
The existence of CUR in the nanoplex was verified by Fourier Trans- suspension was added to 9 mL of deionized water that had been
form Infrared Spectroscopy (FTIR) using Spectrum One (Perkin-Elmer, preheated at an elevated temperature in the range of 65 °C to 100 °C.
USA). The FTIR was performed between 400 and 4500 cm−1 at The resultant suspension was then incubated in a water bath at the ele-
1 cm− 1 spectral resolution for the native CUR, native BSA, CUR-BSA vated temperature for 10 min after which aliquot of the heat-treated
nanoplex, and physical mixture of the two natives. The amorphous nanoplex suspension was withdrawn for measurements of its size and
state of the nanoplex was assessed from its Powder X-ray Diffraction zeta potential. The rest of the nanoplex suspension was then
28 H. Yu et al. / Materials Science and Engineering C 75 (2017) 25–33

centrifuged, followed by freeze drying of the sediments. The payload of Not unlike the trends in the size and zeta potential, the payload was
the heat-treated nanoplex was determined using the same method de- not greatly affected either by the preparation pH (Fig. 2B). The payloads
scribed in Section 2.2.2. The dissolution rate of the nanoplex after the at pH 11.6, 5.1, and 4.9 were similar in magnitude at 53 ± 7%, 58 ± 4%,
thermal treatment was characterized over a 30-min-period by the and 59 ± 4%, respectively. The payloads were found to be slightly higher
same method described in Section 2.2.3. at 70 ± 8% and 71 ± 9%, respectively, upon decreasing the pH to 4.7 and
4.4. The variation in payload, however, was found to be statistically in-
3. Results significant (two-tailed Student's t-test at significance level of 0.05).
In terms of the preparation efficiency, the CUR-BSA nanoplex pre-
3.1. Effects of pH at fixed MBSA/CUR pared at pH 11.6 exhibited low CE of 39 ± 4% and low yield of 7 ± 1%
denoting a very small amount of CUR was transformed to the nanoplex
At fixed MBSA/CUR of 1.2, the CUR-BSA nanoplex was successfully pre- (Fig. 2B). In contrast, at pH 5.1 to 4.4, the CE was significantly higher
pared for all the pH values investigated (i.e. pH 4.4 to 11.6) as indicated ranging from 92 ± 1% at pH 5.1 to 83 ± 5% at pH 4.4 (Fig. 2B), denoting
by the results of the size measurement in Fig. 2A. On this note, the pH high CUR utilization rates at acidic preparation pH. Likewise, the yield
displayed in Fig. 2A represented the pH of the mixed CUR and BSA solu- was also improved greatly to reach 62 ± 2% and 47 ± 2% at pH 5.1
tions. The nanoplex exhibited the smallest size (i.e. 67 ± 2 nm) at and 4.4, respectively (Fig. 2B).
pH 11.6 and the largest size (i.e. 103 ± 4 nm) at pH 4.9. The nanoplexes The optimal pH for the CUR-BSA nanoplex preparation was thus de-
prepared at the other pH showed little size variation as a function of pH termined to be at pH 5.1 (i.e. AA = 0.6% v/v) based on the high CE of
with sizes lying in the narrow range of 80 to 90 nm (Fig. 2A). The zeta N90%, high yield of N 60%, and small nanoplex size (i.e. 84 ± 6 nm) ob-
potential of the nanoplex was also relatively constant as a function of tained under this condition. While comparable CE and yield were pro-
the preparation pH in the range of −19 to −25 mV denoting its good duced at pH 4.9, the nanoplex size was larger (N100 nm) with a lower
colloidal stability (Fig. 2A). Significantly, the negative zeta potentials zeta potential rendering it inferior to the nanoplex produced at pH 5.1.
of the nanoplex after washing in a neutral pH condition indicated that To confirm that pH 5.1 represented the optimal condition, we carried
BSA, which carried net negative charges above its pI [20], had a predom- out an additional run at AA = 0.5% (v/v) corresponding to the prepara-
inant presence on the nanoplex's surface as expected. tion pH of 5.6 (data not plotted for clarity). While the CUR-BSA
nanoplex produced exhibited similar physical characteristics as the
nanoplex prepared at pH 5.1 (i.e. 98 ± 1 nm, −20 ± 7 mV), the prepa-
ration efficiency was highly inferior with CE of and yield of 67 ± 10%
and 35 ± 6%, respectively.
The successful preparation of the CUR-BSA nanoplex at pH 5.1 was
verified by the FESEM image in Fig. 3A, which showed the presence of
nanoparticles with slightly elongated shapes having sizes b100 nm,
hence they were in agreement with the size measurements by PCS.
The PXRD patterns of the CUR-BSA nanoplex after six-month storage re-
vealed the amorphous state of the nanoplex as reflected by the appear-
ance of broad amorphous halos, which were in contrast to the sharp
crystalline peaks observed in the PXRD pattern of the native CUR
(Fig. S1 of Supplementary materials). In this regard, the strong binding
between CUR and BSA inhibited the former from assembling into or-
dered crystalline structures upon precipitation of the CUR-BSA complex,
resulting in the amorphous state formation.
The presence of CUR in the CUR-BSA nanoplex was verified by the
appearance of the characteristics bands of CUR at 1626, 1508, and
1272 cm− 1 in the FTIR spectrum of the nanoplex (Fig. 3B). These
bands, which also appeared in the FTIR spectra of the native CUR and
the physical mixture of CUR and BSA, corresponded to the stretching vi-
brations of the C_ C\\Cring, C _O, and enol C\\O bonds of CUR, respec-
tively [28]. On this note, the characteristic bands of BSA were typically
presented at 1656 and 1545 cm− 1 corresponding to the C _O
stretching (amide I) and C\\N stretching coupled with N\\H bending
(amide II), respectively [13]. The amide bands, however, were not dis-
tinctly visible in the FTIR spectra of the nanoplex and the physical mix-
ture because they overlapped with the characteristic bands of CUR
(Fig. 3B).

3.2. Effects of MBSA/CUR at the optimal pH

At pH 5.1, the nanoplex size was found to be minimally affected by


MBSA/CUR at 0.4 ≤ MBSA/CUR ≤ 2.5, where the sizes fell within the narrow
range of 75 to 90 nm (Fig. 4A). The nanoplex prepared at MBSA/
CUR ≤ 0.4, on the other hand, were considerably larger at 126 ± 13 nm
and 116 ± 18 nm for MBSA/CUR = 0.2 and 0.3, respectively (Fig. 4A).
The zeta potential was also found to be relatively constant as a function
of MBSA/CUR with a majority of the runs produced nanoplexes having
zeta potentials that fell within the narrow range of − 21 to − 27 mV
Fig. 2. Effects of preparation pH on the (A) size, zeta potential; (B) payload, CE, and yield of (Fig. 4A). The exception was for the nanoplex produced at MBSA/
the CURBSA nanoplex produced. CUR = 0.7 in which a higher zeta potential of − 31 ± 0.4 mV was
H. Yu et al. / Materials Science and Engineering C 75 (2017) 25–33 29

Fig. 3. (A) FESEM image and (B) FTIR spectra of the CUR-BSA nanoplex prepared at the
optimal pH.

observed. In contrast, the payload exhibited a distinct trend as a func-


tion of MBSA/CUR, where it gradually decreased with increasing MBSA/
Fig. 4. Effects of MBSA/CUR on the (A) size, zeta potential; (B) payload, CE, and yield of the
CUR from 80 ± 3% at MBSA/CUR = 0.2 to 62 ± 5% at MBSA/CUR = 1.0, before
CUR-BSA nanoplex produced.
it further decreased to 46 ± 4% at MBSA/CUR = 2.5 (Fig. 4B).
The effects of MBSA/CUR on the preparation efficiency were only evi-
dent for 1.0 ≤ MBSA/CUR ≤ 2.5, where a decrease in the CE and yield MBSA/CUR variations, a slight change in the pH from 5.1 to 5.6 as AA
with increasing MBSA/CUR was observed, hence suggesting that prepara- was varied from 0.6% to 0.5% (v/v) caused a significant drop in both
tion at MBSA/CUR N 1.0 was inefficient. Specifically, the CE and yield the CE and yield. Likewise, a slight pH change from 4.9 to 4.7 as AA
steadily decreased from 92 ± 2% and 60 ± 3% at MBSA/CUR = 1.0, respec- was varied from 0.8 to 1.0% (v/v) caused a considerable drop in the
tively, to 46 ± 4% and 24 ± 2% at MBSA/CUR = 2.5 (Fig. 4B). At MBSA/ yield. In contrast, the impacts of the MBSA/CUR were not as significant,
CUR b 1.0, CE was not affected by the MBSA/CUR variation, where it where varying MBSA/CUR from 0.2 to 2.5 (N12-fold increase) only had
remained at nearly 100% denoting excellent CUR utilization rates at an impact on the CE and yield at MBSA/CUR N 1.0. Hence, the preparation
pH 5.1 (Fig. 4B). Similarly, after taking into account the experimental efficiency was more sensitive to pH than MBSA/CUR.
uncertainties, the yield was fairly constant at approximately 80% at
0.2 ≤ MBSA/CUR ≤ 0.8, before it began to decrease at MBSA/CUR = 0.9 to 3.3. Physical characteristics of the optimal CUR-BSA nanoplex formulation
71 ± 3% (Fig. 4B).
Considering that MBSA/CUR b 0.4 resulted in larger nanoplex and 3.3.1. Colloidal stability after preparation
MBSA/CUR N 0.9 led to lower preparation efficiency, the optimal MBSA/ Aqueous suspension of the CUR-BSA nanoplex prepared at the opti-
CUR for the CUR-BSA nanoplex preparation was thus determined to lie mal condition was found to exhibit prolonged colloidal stability after its
between 0.4 and 0.9. While there was little variation in the size, CE, preparation as indicated by the minimal variations in its size and zeta
and yield at 0.4 ≤ MBSA/CUR ≤ 0.9, the payloads of the nanoplex prepared potential during its 24-h storage in ambient condition owed to the ab-
in this MBSA/CUR range were noticeably lower starting at MBSA/CUR = 0.7 sence of irreversible agglomeration (Fig. 5A). The size, which was
(Fig. 4B). This suggested that the optimal MBSA/CUR lay between 0.4 and equal to 87 ± 9 nm after preparation, decreased slightly to 74 ± 6 nm
0.6. Subsequently, the optimal MBSA/CUR was determined to be at 0.6 at- after 1 h, before it went back up to 82 ± 2 nm after 6 h and remained
tributed to the slightly higher zeta potential of the nanoplex produced at approximately that value after 24 h. Likewise, the zeta potential fluc-
at this condition (Fig. 4B). tuated within a narrow range of − 27 to − 31 mV with time, and it
To conclude, the optimal preparation condition was determined to dropped to below −25 mV only after 24 h (Fig. 5A).
be at pH 5.1 and MBSA/CUR = 0.6, resulting in CUR-BSA nanoplex in the
size range of 87 ± 9 nm and zeta potential of −27 ± 0.8 mV with pay- 3.3.2. Antimicrobial activity
load of 69 ± 2%, where the nanoplex was prepared at nearly 100% CE The CUR-BSA nanoplex possessed the same antimicrobial activity
(i.e. 99.8 ± 0.03%) and 80 ± 2% yield. On the significances of pH and against the clinically derived P. aeruginosa as the native CUR, where
30 H. Yu et al. / Materials Science and Engineering C 75 (2017) 25–33

both exhibited MICs equal to 150 μg/mL. This finding served as evidence continued to decrease to (2.8 ± 0.3) × CSat after 8 h before it eventually
that the complexation with BSA did not have any adverse effect on the bottomed out at nearly CSat after 24 h (Fig. 6B).
bioactivity of CUR. The inclusion of HPMC in the dissolution medium greatly reduced
the rate of decrease of the supersaturation level, resulting in a more
prolonged supersaturation level, while maintaining the same maximum
3.3.3. Dissolution rate supersaturation level (Fig. 6). Herein the presence of HPMC not only
The CUR-BSA nanoplex exhibited a rapid dissolution rate under a suppressed the precipitation of CUR from the supersaturated solution,
sink condition with a burst release pattern, where approximately 80% but also the Ostwald-ripening crystallization of the remaining solid
of the CUR in the nanoplex was recovered in the dissolution medium phase of the nanoplex undergoing dissolution. Specifically, the super-
in b1 min (Fig. 5B). The burst release pattern was not unexpected as it saturation level reached the maximum at ≈9 × CSat after 5 min, follow-
was attributed to the nanosize and amorphous state of the nanoplex. ed by its slow decrease to (7.8 ± 0.9) × CSat after 1 h (Fig. 6A). The
The % Dissolution reached the maximum at 86 ± 2% after 5 min, follow- supersaturation level then continued to decrease to (6.6 ± 0.1) × CSat
ed by its gradual decrease to 68 ± 1% after 30 min due to hydrolytic deg- and (2.3 ± 0.1) × CSat after 8 h and 24 h, respectively (Fig. 6B).
radation of the dissolved CUR in PBS. For comparison, the % Dissolution For comparison, the CUR-CHI nanoplex achieved a slightly lower
of the CUR-CHI nanoplex reached approximately 60% after 1 min and maximum supersaturation level at (7.5 ± 0.2) × CSat in the absence of
was at the maximum at 73 ± 5% after 5 min, hence denoting its slower HPMC (Fig. 6). The supersaturation level then rapidly decreased to
dissolution rate compared to the CUR-BSA nanoplex (Fig. 5B). (1.9 ± 0.1) × CSat after 1 h and bottomed out at CSat after just 2 h. The
inclusion of HPMC increased the maximum supersaturation level to
the level achieved by the CUR-BSA nanoplex, while also significantly
3.3.4. Supersaturation generation slowed down its rate of decrease as expected, where the supersatura-
The CUR-BSA nanoplex exhibited the typical “spring and parachute” tion level only decreased slightly to (7.3 ± 0.4) × CSat after 1 h and
supersaturation profile of amorphous drug formulations, which was did not reach CSat after 24 h. These results showed that while the max-
characterized by rapid generation of a supersaturation peak (“spring”), imum supersaturation level achieved by the CUR-BSA nanoplex and
followed by its gradual decrease to the thermodynamic saturation solu- CUR-CHI nanoplex was comparable, the rate of decrease of the supersat-
bility (“parachute”) [29]. In the absence of HPMC in the dissolution me- uration level was significantly faster in the latter, regardless of whether
dium, the maximum supersaturation level generated by the CUR-BSA HPMC was present or not in the dissolution medium.
nanoplex was equal to (9.0 ± 0.1) × CSat achieved in b 5 min (0.083 h)
after dissolution (Fig. 6A). The supersaturation level then gradually de-
creased to (4.7 ± 0.2) × CSat after 1 h due to the precipitation of CUR
from the supersaturated solution (Fig. 6A). The supersaturation level

Fig. 5. (A) Colloidal stability of the CUR-BSA nanoplex during ambient 24-h storage;
(B) dissolution timeprofile of the CUR-BSA nanoplex in comparison to that of the CUR- Fig. 6. Supersaturation time-profiles of the CUR-BSA nanoplex in the first (A) 1-h and
CHI nanoplex. (B) 24-h of dissolution in comparison with the CUR-CHI nanoplex.
H. Yu et al. / Materials Science and Engineering C 75 (2017) 25–33 31

3.4. Modulated CUR release respectively, to produce nanoplexes having similar payloads after the
thermal treatment at around 45% (Fig. 7A).
3.4.1. Effect of thermal treatment on the nanoplex's structural integrity Next, the size and zeta potential of the heat-treated nanoplexes were
The effect of heating temperature on the payload of the CUR-BSA characterized to investigate whether any structural changes had taken
nanoplex after the thermal treatment was presented in Fig. 7A. For the place upon heating. The size was found to only change slightly after
nanoplex prepared at MBSA/CUR = 0.6, the payload decreased from the thermal treatment, where the size of the nanoplex prepared at
69 ± 2% before heating to 46 ± 6% after heating at 65 °C as the heat- MBSA/CUR = 0.6, 1.0, and 2.5 changed from 108 ± 6 nm, 101 ± 2 nm,
induced unfolding of the BSA chains in the nanoplex caused the release and 90 ± 3 nm, respectively, before heating to 83 ± 2 nm, 121 ±
of CUR molecules that were originally bound to the hydrophobic cavi- 21 nm, and 74 ± 3 nm after heating (Fig. 7B), hence the size variations
ties of these chains. The payload continued to decrease with increasing were not statistically significant. The change in the zeta potential after
temperature to eventually result in payload b20% at temperatures the thermal treatment was also statistically insignificant, where the
above 75 °C, which denoted the onset of structural collapse of the zeta potentials remained in the narrow range of − 24 to − 30 mV
nanoplex. (Fig. 7B). The minimal changes in the size and zeta potential of the
Likewise, for the nanoplex prepared at MBSA/CUR = 1.0, the payload nanoplex signified the absence of aggregation of BSA in the nanoplex
decreased from 62 ± 5% before heating to 45 ± 3% after heating at that typically occurred upon protein denaturation. Hence, the thermal
65 °C. The after-treatment payload, nonetheless, remained at ≈ 45% treatments had not adversely affected the nanoplex's structural
for heating up to 80 °C. For the nanoplex prepared at MBSA/CUR = 2.5, integrity.
the after-treatment payload only started to decrease for heating above
80 °C, where it decreased from 48 ± 2% before heating to b 25% after 3.4.2. Effect of thermal treatment on the CUR release
heating. The onset temperatures for the structural collapse of the The dissolution time-profiles in the first 15 min of the 30-min disso-
nanoplexes prepared at MBSA/CUR = 0.6, 1.0, and 2.5 were thus equal lution period were shown in Fig. 8 to better distinguish the dissolution
to 75, 85, and 80 °C, respectively. The higher onset temperatures for profiles between the non-treated and heat-treated nanoplexes. The
the nanoplexes prepared at MBSA/CUR N 0.6 were attributed to their thermal treatment had a negligible effect on the % Dissolution for the
lower CUR payload, hence higher BSA content, such that a higher degree nanoplex prepared at MBSA/CUR = 0.6 (Fig. S2 of Supplementary mate-
of heating was needed to cause complete unfolding of the BSA chains in rials), whereas the heat-treated nanoplexes prepared at MBSA/CUR =
the nanoplex. Thermal treatments at 65 °C, 80 °C, and 80 °C were thus 1.0 and 2.5 exhibited faster dissolution rates than their non-heat-
performed on the nanoplexes prepared at MBSA/CUR = 0.6, 1.0, and 2.5, treated counterparts (Figs. 8A, B). On this note, the gradual decrease
in the % Dissolution observed after 5 min was due to the aforemen-
tioned hydrolytic degradation of CUR in PBS.
More specifically, for the heat-treated nanoplex prepared at MBSA/
CUR = 2.5, the % Dissolution reached the maximum at 93 ± 4% in
b1 min compared to 67 ± 6% achieved by the non-treated nanoplex
in the same period. A similar trend was observed for the nanoplex pre-
pared at MBSA/CUR = 1.0, however, the difference in the maximum % Dis-
solution between the heat-treated and non-treated nanoplexes was
smaller at 90 ± 3% and 83 ± 2%, respectively. The faster dissolution
rate after the thermal treatment was likely caused by the unfolding of
BSA chains after heating that in turn increased the exposure of the
bound CUR molecules to the dissolution medium. To conclude, the ef-
fects of the thermal treatment increased in significance with increasing
MBSA/CUR as higher MBSA/CUR resulted in a lower payload before the ther-
mal treatment, hence higher BSA contents.

4. Discussion

4.1. Effects of pH

Like most proteins, the charge heterogeneity of BSA, where patches


of positive and negative charges co-existed [20], enabled BSA to electro-
statically interact with both positively and negatively charged species
for a wide range of pH. In this study, both CUR and BSA were fully ion-
ized with opposite charges in their respective solutions (i.e. KOH and
AA, respectively) as we aimed to promote their electrostatic binding
upon mixing, which together with the hydrophobic binding led to the
formation of soluble CUR-BSA complex. As the mixed solution equili-
brated to the final pH, however, the degree of ionization of both CUR
and BSA in the complex varied and consequently influenced the
nanoplex formation process. While CUR-BSA nanoplexes having fairly
similar physical characteristics (i.e. size, zeta potential, and payload)
were successfully prepared independent of pH at MBSA/CUR = 1.2, the
preparation efficiency, on the contrary, exhibited a strong pH
dependence.
For final pH in the range of 5.6 to 4.4, which fell within or slightly
below the reported pI range of BSA, BSA essentially carried a zero net
Fig. 7. Effects of thermal treatment on the (A) payload; (B) size and zeta potential of the charge, while CUR also lost some of its charges as the pH fell below its
CUR-BSA nanoplex prepared at MBSA/CUR = 0.6, 1.0, and 2.5. pKa. Had the soluble CUR-BSA complex been predominantly held
32 H. Yu et al. / Materials Science and Engineering C 75 (2017) 25–33

preparation of CUR-CHI nanoplex in which the CE was lower at higher


mixing ratio of CHI to CUR (MCHI/CUR) [6]. The presence of an excess
amount of unbound BSA chains likely hindered the cooperative binding
of the free CUR molecules to the complex, resulting in the lower CE. The
relatively constant CE at MBSA/CUR b 1.0 indicated a critical MBSA/CUR
existed above which the cooperative binding became hindered. Impor-
tantly, the nearly 100% CE obtained at MBSA/CUR = 0.2 indicated that
only a small amount of BSA was needed to successfully form the CUR-
BSA nanoplex.
Likewise, the payload also exhibited a strong dependence on MBSA/
CUR, where the payload decreased with increasing MBSA/CUR. While the
decrease in the payload at MBSA/CUR ≥ 1.0 was due to the lower CE ob-
served at higher MBSA/CUR, the decrease in payload at MBSA/CUR b 1.0 oc-
curred despite the CE remained relatively constant. This suggested that
an increasing amount of BSA participated in the complexation for the
same amount of CUR transformed to the nanoplex when there was an
increased abundance of BSA. Thus, while only a small amount of BSA
was needed to achieve nearly 100% CE for CUR, the abundant presence
of BSA at higher MBSA/CUR caused more BSA to be transformed to the
nanoplex.

4.3. Comparisons of CUR-BSA nanoplex versus CUR-CHI nanoplex

In terms of both the physical characteristics and preparation effi-


ciency, the optimal formulation of the CUR-BSA nanoplex was found,
for the most part, to be superior to that of the CUR-CHI nanoplex pre-
sented earlier in Nguyen, Yu, Kiew and Hadinoto [6]. The only aspect
in which the CUR-CHI nanoplex was superior was in the payload,
where the CUR-CHI nanoplex exhibited payload of 84 ± 4% [6] versus
69 ± 2% for the CUR-BSA nanoplex.
First, the CUR-BSA nanoplex (87 ± 9 nm) was significantly smaller
than the CUR-CHI nanoplex (260 ± 9 nm) [6]. As shown earlier, the
smaller size of the CUR-BSA nanoplex led to its faster dissolution rate
Fig. 8. Dissolution time-profiles of the heat-treated and non-treated CUR-BSA nanoplexes owed to the larger specific surface area. The faster dissolution made
prepared at MBSA/CUR of (A) 1.0 and (B) 2.5. the CUR-BSA nanoplex less prone to the Ostwald-ripening recrystalliza-
tion of the amorphous solid phase while undergoing dissolution, which
together by the electrostatic binding, the loss of charges would have led was known to be prevalent in amorphous solid dispersion exhibiting a
to disengagement of the complex and consequently a lack of nanoplex slow dissolution rate [30]. The reduced crystallization propensity of
formation. The fact that the nanoplex remained formed at high CE the CUR-BSA nanoplex upon dissolution resulted in the increased su-
(N90%) indicated that significant disengagement of the complex did persaturation generation presented earlier, which explained for the
not take place. This signified the importance of the CUR-BSA hydropho- slower rate of decrease of its supersaturation level compared to the
bic binding in the CUR-BSA complex formation. CUR-CHI nanoplex.
For final pH of 11.6, which lay above the pI of BSA and pKa of CUR, Second, the CUR-BSA nanoplex exhibited higher colloidal stability
both BSA and CUR were negatively charged. The soluble complex, how- than the CUR-CHI nanoplex during ambient storage, despite them hav-
ever, remained formed as CUR theoretically could still electrostatically ing comparable zeta potentials of −27 ± 0.8 mV and 23 ± 3 mV [6], re-
bind to the positively charged patches of BSA, and the hydrophobic spectively. Specifically, the CUR-BSA nanoplex remained individually
binding still took place. The weak nanoplex formation, which was dispersed after 24 h compared to 6 h for the CUR-CHI nanoplex [6]
reflected by the low yield (b10%), however, indicated that aggregation hinting at the superior steric stabilizing role of BSA compared to CHI.
of the CUR-BSA complex did not take place as effectively, resulting in a The increased coverage of the nanoplex surface by BSA (as proteins
lack of critical mass for precipitation. are well-known to produce corona surrounding nanoparticles [12])
In this regard, the presence of the net negative charge of BSA in the was likely responsible for the improved colloidal stability.
complex was postulated to hinder the cooperative binding of CUR Third, while the preparation of both nanoplexes exhibited high CUR
from taking place. Herein the cooperative binding referred to the bind- utilization rates with CE N 90%, the yield of the CUR-BSA nanoplex prep-
ing of free CUR molecules to the CUR molecules already in the complex aration was significantly higher at 80 ± 2% compared to 47 ± 2% for the
via inter-CUR hydrophobic interactions. The net negative charge of BSA CUR-CHI nanoplex [6]. The higher yield was attributed to the lower
exerted repulsive forces to the similarly charged CUR molecules, hence mixing ratio required to achieve the high CE for the CUR-BSA nanoplex,
preventing the free CUR molecules from interacting with the complex. where the optimal MBSA/CUR was equal to 0.6 compared to the optimal
The lack of cooperative binding caused the aggregation of the complex MCHI/CUR equal to 1.2 [6]. The lower mixing ratio meant that the amount
to diminish, resulting in the weak nanoplex formation. of BSA needed to be supplied in the feed to transform nearly 100% of the
CUR into the nanoplex was less compared to that of CHI, resulting in the
4.2. Effects of MBSA/CUR higher yield.

In addition to its pH dependence, the preparation efficiency also ex- 5. Conclusions


hibited a strong dependence on MBSA/CUR at MBSA/CUR N 1.0, where the
increasing presence of excess BSA at higher MBSA/CUR resulted in lower A bioavailability enhancement strategy of CUR was successfully de-
CE and consequently lower yield. The same trend was observed in the veloped in the form of amorphous CUR-BSA nanoplex prepared via a
H. Yu et al. / Materials Science and Engineering C 75 (2017) 25–33 33

simple method based on drug-protein complexation. We have success- high payload and supersaturation generation, Eur. J. Pharm. Biopharm. 96 (2015)
1–10.
fully demonstrated for the first time the self-assembled formation of [7] A.O. Elzoghby, W.M. Samy, N.A. Elgindy, Albumin-based nanoparticles as potential
drug-protein nanoparticle complex having well-defined morphology controlled release drug delivery systems, J. Control. Release 157 (2012) 168–182.
and high drug payload. The effects of two key variables in drug- [8] J. Qi, P. Yao, F. He, C. Yu, C. Huang, Nanoparticles with dextran/chitosan shell and
BSA/chitosan core-doxorubicin loading and delivery, Int. J. Pharm. 393 (2010)
protein complexation (i.e. pH and MBSA/CUR) on the physical characteris- 176–184.
tics and preparation efficiency of the nanoplex produced were thor- [9] D. Zhao, X. Zhao, Y. Zu, J. Li, Y. Zhang, R. Jiang, Z. Zhang, Preparation, characterization,
oughly investigated. While the effects of the preparation pH and MBSA/ and in vitro targeted delivery of folate-decorated paclitaxel-loaded bovine serum al-
bumin nanoparticles, Int. J. Nanomedicine 5 (2010) 669–677.
CUR on the physical characteristics of the nanoplex were found in gener-
[10] S.H. Brewer, W.R. Glomm, M.C. Johnson, M.K. Knag, S. Franzen, Probing BSA binding
al to be insignificant (with the exception of the payload's dependence to citrate-coated gold nanoparticles and surfaces, Langmuir 21 (2005) 9303–9307.
on MBSA/CUR), the same could not be said about the preparation efficien- [11] Z. Ji, X. Jin, S. George, T. Xia, H. Meng, X. Wang, E. Suarez, H. Zhang, E.M.V. Hoek, H.
Godwin, A.E. Nel, J.I. Zink, Dispersion and stability optimization of TiO2 nanoparticles
cy. The CUR-BSA nanoplex formation was found to favor the acidic prep-
in cell culture media, Environ. Sci. Technol. 44 (2010) 7309–7314.
aration pH and MBSA/CUR below unity as evidenced by the consistently [12] A.O. Luby, E.K. Breitner, K.K. Comfort, Preliminary protein corona formation stabi-
high CE and yield observed under this condition. At the optimal prepa- lizes gold nanoparticles and improves deposition efficiency, Appl. Nanosci. 6
ration condition (i.e. pH = 5.1 and MBSA/CUR = 0.6), CUR-BSA nanoplex (2016) 827–836.
[13] P. Bourassa, C.D. Kanakis, P. Tarantilis, M.G. Pollissiou, H.A. Tajmir-Riahi, Resveratrol,
with size of ≈90 nm, zeta potential of ≈−27 mV, and payload of ≈70% genistein, and curcumin bind bovine serum albumin, J. Phys. Chem. B 114 (2010)
were produced at nearly 100% CE and ≈80% yield. 3348–3354.
The amorphous state stability after six-month storage, colloidal sta- [14] A. Papadopoulou, R.J. Green, R.A. Frazier, Interaction of flavonoids with bovine
serum albumin: a fluorescence quenching study, J. Agric. Food Chem. 53 (2005)
bility during 24-h ambient storage, and antimicrobial activity of the 158–163.
CUR-BSA nanoplex prepared at the optimal condition were successfully [15] M.H.M. Leung, T.W. Kee, Effective stabilization of curcumin by association to plasma
established. Upon dissolution, the CUR-BSA nanoplex produced a proteins: human serum albumin and fibrinogen, Langmuir 25 (2009) 5773–5777.
[16] M.L. Yang, Y. Wu, J.B. Li, H.B. Zhou, X.Y. Wang, Binding of curcumin with bovine
prolonged supersaturation level at ≈ 9 × of the saturation solubility serum albumin in the presence of iota-carrageenan and implications on the stability
that was maintained for N4 h in the presence of HPMC, hence denoting and antioxidant activity of curcumin, J. Agric. Food Chem. 61 (2013) 7150–7155.
its excellent supersaturation generation capability. We also successfully [17] Y.J. Wang, M.H. Pan, A.L. Cheng, L.I. Lin, Y.S. Ho, C.Y. Hsieh, J.K. Lin, Stability of
curcumin in buffer solutions and characterization of its degradation products, J.
demonstrated the modulated release functionality of the CUR-BSA Pharm. Biomed. Anal. 15 (1997) 1867–1876.
nanoplex achieved by thermal treatment of the nanoplex. The impact [18] A. Barik, K.I. Priyadarsini, H. Mohan, Photophysical studies on binding of curcumin to
of the thermal treatment on the modified CUR release was found to in- bovine serum albumin, Photochem. Photobiol. 77 (2003) 597–603.
[19] W.S. Cheow, K. Hadinoto, Self-assembled amorphous drug-polyelectrolyte nanopar-
crease in significance for nanoplex exhibiting high BSA contents before
ticle complex with enhanced dissolution rate and saturation solubility, J. Colloid In-
the thermal treatment. Most importantly, the CUR-BSA nanoplex was terface Sci. 367 (2012) 518–526.
found to be superior in terms of the size, supersaturation generation, [20] V. Ball, M. Winterhalter, P. Schwinte, P. Lavalle, J.C. Voegel, P. Schaaf, Complexation
colloidal stability, and yield when compared to the CUR-CHI nanoplex mechanism of bovine serum albumin and poly(allylamine hydrochloride), J. Phys.
Chem. B 106 (2002) 2357–2364.
prepared by drug-polysaccharide complexation. [21] Y. Moriyama, E. Watanabe, K. Kobayashi, H. Harano, E. Inui, K. Takeda, Secondary
structural change of bovine serum albumin in thermal denaturation up to 130 °C
Acknowledgement and protective effect of sodium dodecyl sulfate on the change, J. Phys. Chem. B
112 (2008) 16585–16589.
[22] R. Su, W. Qi, Z. He, Y. Zhang, F. Jin, Multilevel structural nature and interactions of
The authors would like to acknowledge the funding from bovine serum albumin during heat-induced aggregation process, Food Hydrocoll.
GlaxoSmithKline (Singapore) under its Sustainable Manufacturing Trust 22 (2008) 995–1005.
[23] M.H.M. Leung, H. Colangelo, T.W. Kee, Encapsulation of curcumin in cationic mi-
Fund 2013 (PI: Kunn Hadinoto Ong). celles suppresses alkaline hydrolysis, Langmuir 24 (2008) 5672–5675.
[24] L.R.S. Barbosa, M.G. Ortore, F. Spinozzi, P. Mariani, S. Bernstorff, R. Itri, The impor-
Appendix A. Supplementary data tance of protein-protein interactions on the pH-induced conformational changes
of bovine serum albumin: a small-angle X-ray scattering study, Biophys. J. 98
(2010) 147–157.
Supplementary data to this article can be found online at http://dx. [25] T. Estey, J. Kang, S.P. Schwendeman, J.F. Carpenter, BSA degradation under acidic
doi.org/10.1016/j.msec.2017.02.018. conditions: a model for protein instability during release from PLGA delivery sys-
tems, J. Pharm. Sci. 95 (2006) 1626–1639.
[26] D. Sun, X. Zhuang, X. Xiang, Y. Liu, S. Zhang, C. Liu, S. Barnes, W. Grizzle, D. Miller, H.-
References G. Zhang, A novel nanoparticle drug delivery system: the anti-inflammatory activity
of curcumin is enhanced when encapsulated in exosomes, Mol. Ther. 18 (2010)
[1] S. Prasad, S.C. Gupta, A.K. Tyagi, B.B. Aggarwal, Curcumin, a component of golden 1606–1614.
spice: from bedside to bench and back, Biotechnol. Adv. 32 (2014) 1053–1064. [27] CLSI, Clinical and Laboratory Standards Institute Guidelines on “Performance Stan-
[2] G. Kotan, F. Kardas, O.A. Yokus, O. Akyildirim, H. Saral, T. Eren, M.L. Yola, N. Atar, A dards for Antimicrobial Susceptibility Testing” (Document M100-S17), Wayne,
novel determination of curcumin via Ru@Au nanoparticle decorated nitrogen and USA, 2007.
sulfur-functionalized reduced graphene oxide nanomaterials, Anal. Methods 8 [28] P.R.K. Mohan, G. Sreelakshmi, C.V. Muraleedharan, R. Joseph, Water soluble com-
(2016) 401–408. plexes of curcumin with cyclodextrins: characterization by FT-Raman spectroscopy,
[3] P. Anand, A.B. Kunnumakkara, R.A. Newman, B.B. Aggarwal, Bioavailability of Vib. Spectrosc. 62 (2012) 77–84.
curcumin: problems and promises, Mol. Pharm. 4 (2007) 807–818. [29] D. Alonzo, G. Zhang, D. Zhou, Y. Gao, L. Taylor, Understanding the behavior of amor-
[4] O. Naksuriya, S. Okonogi, R.M. Schiffelers, W.E. Hennink, Curcumin phous pharmaceutical systems during dissolution, Pharm. Res. 27 (2010) 608–618.
nanoformulations: a review of pharmaceutical properties and preclinical studies [30] M.E. Matteucci, B.K. Brettmann, T.L. Rogers, E.J. Elder, R.O. Williams, K.P. Johnston,
and clinical data related to cancer treatment, Biomaterials 35 (2014) 3365–3383. Design of potent amorphous drug nanoparticles for rapid generation of highly su-
[5] M. Sun, X. Su, B.Y. Ding, X.L. He, X.J. Liu, A.H. Yu, H.X. Lou, G.X. Zhai, Advances in persaturated media, Mol. Pharm. 4 (2007) 782–793.
nanotechnology-based delivery systems for curcumin, Nanomedicine 7 (2012)
1085–1100.
[6] M.H. Nguyen, H. Yu, T.Y. Kiew, K. Hadinoto, Cost-effective alternative to nano-
encapsulation: amorphous curcumin-chitosan nanoparticle complex exhibiting

You might also like