Acessibilidade / Reportar erro

Development and validation of a reversed-phase HPLC method for quantification of 1’-acetoxychavicol acetate content in a nanostructured lipid carrier formulation

Abstract

1’-acetoxychavicol acetate (ACA)-loaded nanostructured lipid carriers (NLCs) were formulated for prostate cancer therapy and to determine the optimal therapeutic dose, we developed a rapid, specific, and accurate reversed-phase high-performance liquid chromatography (RP-HPLC) method to quantify the ACA content in NLCs. The method was validated according to International Council for Harmonisation of Technical Requirements for Pharmaceuticals for Human Use (ICH) guidelines. Chromatographic separation of ACA from the lipid components was performed with an Agilent 1220 Infinity LC system and ultraviolet detector using an Agilent Poroshell C18 column (4.6 x 250.0 mm). The mobile phase consisted of acetonitrile and water (80:20 [v/v]) with a flow rate of 0.8 mL/min in isocratic mode. Linearity of the standard curve was assessed at an ACA concentration range of 5-200 µg/mL, and a 1/x weighted linear regression was adopted for the calibration curve. The calculated limits of detection and quantification were 0.59 µg/mL and 1.79 µg/mL, respectively. The mean percent recovery of ACA was 100.02% (relative SD, 2%), and the coefficients of variation for intraday and interday assays were within the values required by the ICH. We also demonstrated robustness of the method by altering the mobile phase ratio and flow rate. Furthermore, we proved specificity of the method for ACA by comparing chromatograms of the blank NLC and ACA-NLC. Hence, we effectively used this validated method to determine the drug-loading capacity and entrapment efficiency of the NLCs.

Keywords:
High-performance liquid chromatography; Validation; 1’-acetoxychavicol acetate; Nanostructured lipid carrier; Chemotherapy

INTRODUCTION

1’-acetoxychavicol acetate (ACA) (Figure 1) is a phytoconstituent isolated from the rhizomes of the Southeast Asian ethnomedicinal plants Alpinia conchigera Griff. and Alpinia galanga (L.) Willd. (Zingiberaceae) (Janssen, Scheffer, 1985Janssen AM, Scheffer JJ. Acetoxychavicol acetate, an antifungal component of Alpinia galangal. Planta Med.1985;51(6):507-11.; Barik, Kundu, Dey, 1987Barik BR, Kundu AB, Dey AK. Two phenolic constituents from Alpinia galanga rhizomes. Phytochemistry. 1987;26(7):2126-27.; Kondo et al., 1993Kondo A, Ohigashi H, Murakami A, Jiwajinda S, Koshimizu K. A potent inhibitor of tumor promoter-induced Epstein-Barr virus activation, 1’-acetoxychavicol acetate from Languas galanga, a traditional Thai condiment. Biosci Biotechnol Biochem.1993;57(8):1344-45.; Yang, Eilerman, 1999Yang X, Eilerman RG. Pungent principal of Alpinia galangal (L.) swartz and its applications. J Agric Food Chem.1999;47(4):1657-62.). ACA induces apoptosis-mediated cell death in many cancer cell lines with minimal toxicity in normal cells (In et al., 2012In LLA, Arshad NM, Ibrahim H, Azmi MN, Awang K, Nagoor NH. 1’-Acetoxychavicol acetate inhibits growth of human oral carcinoma xenograft in mice and potentiates cisplatin effect via proinflammatory microenvironment alterations. BMC Complement Altern Med. 2012;12:179.). Studies demonstrated that ACA prevents Ehrlich ascites carcinoma, skin tumor, and adenocarcinoma formation (Tanaka, Kawabata et al., 1997Tanaka T, Makita H, Kawamori T, Kawabata K, Mori H, Murakami A, et al. A xanthine oxidase inhibitor 1’-acetoxychavicol acetate inhibits azoxymethane-induced colonic aberrant crypt foci in rats. Carcinogenesis. 1997;18(5):1113-8.; Tanaka, et al., 1997Tanaka T, Kawabata K, Kakumoto M, Makita H, Matsunaga K, Mori H, et al. Chemoprevention of azoxymethane-induced rat colon carcinogenesis by a xanthine oxidase inhibitor, 1’-acetoxychavicol acetate. Jpn JCancer Res .1997;88(9):821-30.; Nakamura et al., 1998Nakamura Y, Murakami A, Ohto Y, Torikai K, Tanaka T, Ohigashi H. Suppression of tumor promoter-induced oxidative stress and inflammatory responses in mouse skin by a superoxide generation inhibitor 1’-acetoxychavicol acetate. Cancer Res. 1998;58(21):4832-9.; Narukawa et al., 2010Narukawa M, Koizumi K, Iwasaki Y, Kubota K, Watanabe T. Galangal pungent component, 1’-acetoxychavicol acetate, activates TRPA1. Biosci Biotechnol Biochem . 2010;74(8):1694-6.). Moreover, ACA induces apoptosis of myeloid leukemic cells via mitochondrial and Fas-mediated mechanisms (Ito et al., 2004Ito K, Nakazato T, Murakami A, Yamato K, Miyakawa Y, Yamada T, et al. Induction of apoptosis in human myeloid leukemic cells by 1′-acetoxychavicol acetate through a mitochondrial-and Fas-mediated dual mechanism. Clin Cancer Res. 2004;10(6):2120-30.). In vitro studies also showed that ACA induces dose- and time-dependent cytotoxicity in tumor cells, potentially induces cell-cycle arrest at G0/G1 phase and suppresses the proliferation and migration rates for oral squamous cell carcinoma (Awang et al., 2010Awang K, Azmi MN, Aun LI, Aziz AN, Ibrahim H, Nagoor NH. The apoptotic effect of 1’-S-1’-acetoxychavicol acetate from Alpinia conchigera on human cancer cells. Molecules. 2010;15(11):8048-8059.). In short, ACA’s potential as an effective antitumor agent renders it useful for testing in vivo.

FIGURE 1
Chemical structure of ACA.

However, in animal studies of ACA in its free form, investigators have encountered several drawbacks, such as poor in vivo solubility, resulting in a decline in biological activity. ACA is a hydrophobic compound, which makes delivering it to cancer cells within an aqueous environment challenging. In addition, passive targeting of tumor cells causes inefficient use of free ACA (Arshad et al., 2015Arshad NM, In LLA, Soh TL, Mohamad NA, Ibrahim H, Awang K, et al. Recombinant human alpha fetoprotein synergistically potentiates the anti-cancer effects of 1′-S-1′-acetoxychavicol acetate when used as a complex against human tumours harbouring AFP-receptors. Oncotarget. 2015;6(18):16151-67.). To address these problems, researchers have loaded ACA into nanostructured lipid carriers (NLCs) for delivery to cancer cells. Previously, the cosmetics industry extensively employed NLCs for drug delivery (Müller, Radtke, Wissing, 2002Müller RH, Radtke M, Wissing SA. Solid lipid nanoparticles (SLN) and nanostructured lipid carriers (NLC) in cosmetic and dermatological preparations. Adv Drug Deliv Rev. 2002;54(Suppl 1):S131-55.; Beloqui et al., 2016Beloqui A, Solinis MA, Rodriguez-Gascon A, Almeida AJ, Preat V. Nanostructured lipid carriers: Promising drug delivery systems for future clinics. Nanomedicine. 2016;12(1):143-61.). In the present study, we used a modified NLC to encapsulate ACA for targeted parenteral delivery to cancer cells. Then, we selected the optimized nanoparticle based on its favorable properties of small size, protection against drug degradation, high drug-loading capacity and entrapment efficiency, enhanced dispersion, sustained drug release, and prolonged stability (Pathak, 2009Pathak Y. Recent developments in nanoparticulate drug delivery systems. In: Pathak Y, Thassu D, editors. Drug Delivery Nanoparticles Formulation and Characterization (1st ed.). Boca Raton, FL: CRC Press; 2009. p. 1-7.).

To determine the drug-loading capacity and entrapment efficiency of the nanoparticles, a validated methodology for quantitative determination of the drug content is essential. Researchers have previously quantified ACA as a free drug using high-performance liquid chromatography (HPLC) under different chromatographic conditions (Batra et al., 2012Batra V, Syed Z, Gill JN, Coburn MA, Adegboyega P, DiGiovanni J, et al. Effects of the tropical ginger compound, 1’-acetoxychavicol acetate, against tumor promotion in K5. Stat3C transgenic mice. J Exp Clin Cancer Res. 2012;31(1):57.; Haque et al., 2017Haque AKMM, Leong KH, Lo YL, Awang K, Nagoor NH. In vitro inhibitory mechanisms and molecular docking of 1’-S-1’-acetoxychavicol acetate on human cytochrome P450 enzymes. Phytomedicine. 2017;31:1-9.). In addition, some authors have reported on HPLC methods for isolation and purification of ACA from A. galanga extracts without specific ACA quantification techniques (Kaur et al., 2010Kaur A, Singh R, Dey CS, Sharma SS, Bhutani KK, Singh IP. Antileishmanial phenylpropanoids from Alpinia galanga (Linn.) Willd. Indian J Exp Biol. 2010;48(3):314-7.; Baradwaj, Rao, Kumar, 2017Baradwaj RG, Rao MV, Kumar TS. Novel purification of 1’-S-1’-Acetoxychavicol acetate from Alpinia galanga and its cytotoxic plus antiproliferative activity in colorectal adenocarcinoma cell line SW480. Biomed Pharmacother. 2017;91:485-493.). However, quantification of ACA in nanoparticle formulations has neither been validated nor reported. An ACA-loaded NLC is a novel formulation; thus, accurate quantification of ACA in NLCs requires a reliable, validated analytical method. The drug content in nanoparticle-based formulations is commonly determined using ultraviolet (UV) spectrophotometry or HPLC, but HPLC is preferred for its greater sensitivity. Also, interference of contaminants during determination of drug content can be prevented with the component separation technique employed in HPLC. Therefore, we undertook this study to develop and validate a reversed-phase HPLC method to quantify ACA and determine the drug-loading capacity and entrapment efficiency of the NLCs.

MATERIAL AND METHODS

Chemicals and reagents

D, L-1’-acetoxychavicol acetate (ACA) (98.8%) was purchased from LKT Laboratories Inc. (St. Paul, MN, USA). Cocoa butter (White Naturals, Cape Coral, FL, USA), isopropyl myristate (Thermo Fisher Scientific, Waltham, MA, USA), Span 40 (Sigma-Aldrich, St. Louis, MO, USA), and Tween 80 (Merck & Co., Inc., Kenilworth, NJ, USA) were also procured. HPLC-grade water and acetonitrile were purchased from Merck & Co., Inc. All other chemicals and solvents obtained commercially were of analytical or HPLC grade. Before instrumentation, the mobile phase solvents were filtered through a 0.22-µm Millipore membrane filter (Merck & Co., Inc.) and degassed using a vacuum pump.

Instrument

An Agilent 1220 Infinity LC system (Agilent Technologies, Inc., Santa Clara, CA, USA) equipped with column and sample compartment with temperature control, UV-visible diode-array detector, binary pump, autosampler, and variable wavelength detector was used in the study. All analyses were conducted using an Agilent Poroshell C18 column with a 4-µm particle size, 4.6-mm internal diameter, and 250-mm length. The column was equilibrated for 1 h before the analysis. OpenLab CDS ChemStation Edition (Agilent Technologies, Inc.) was used for data acquisition, analysis, and reporting.

Chromatographic conditions

The chromatographic conditions were standardized by evaluating the peak symmetry of ACA under different mobile phase conditions, flow rates, and UV wavelengths. Methanol or acetonitrile was used as the organic solvent at 10:90, 20:80, and 30:70 (v/v) water:organic solvent proportions. The flow rate was tested at 0.8-1.0 mL/ min. In addition, UV wavelengths of 216, 226, and 254 nm were investigated. After testing all parameters and analyzing the ACA peak symmetry, the chromatographic conditions were standardized as follows: the mobile phase consisted of A-water and B-acetonitrile at a proportion of 20:80 (v/v) in isocratic mode, the flow rate was set at 0.8 mL/min, the column temperature was set at 25 °C, the sample injection volume was 10 µL, and the UV wavelength was set at 216 nm. The method run time was 5 min. To quantify ACA concentration, the peak area of UV absorbance was recorded.

Preparation of samples and standard solutions

ACA-NLCs were prepared in triplicate via melt and a high-shear homogenization method (Severino, Santana, Souto, 2012Severino P, Santana MH, Souto EB. Optimizing SLN and NLC by 2(2) full factorial design: Effect of homogenization technique. Mater Sci Eng C Mater Biol Appl. 2012;32(6):1375-9.). The lipid phase containing cocoa butter, isopropyl myristate, and ACA was heated to 45 °C before being dispersed together with the aqueous phase containing deionized water, Tween 80, and Span 40 heated to the same temperature. The mixture was then homogenized using a high-shear homogenizer (Heidolph SilentCrusher, Sigma-Aldrich, St. Louis, MO, USA) for 20 min. The resulting sample was immediately kept in a 4 °C refrigerator overnight to allow for the formation of drug-loaded nanoparticles. The samples used to evaluate drug load and entrapment efficiency were obtained as described further in the “Method applicability” section below. As for the standards for calibration curve, ACA stock solution (1 mg/mL) was dissolved in acetonitrile to provide concentrations ranging from 5 to 200 µg/mL. All samples and standard solutions were filtered through Whatman 0.2-µm-pore nylon membrane syringe filters (Whatman, Maidstone, UK) before analysis.

System suitability and method validation

Six replicates of standard ACA solutions (100 µg/ mL) were analyzed to determine the system suitability. Factors such as the number of theoretical plates (N) and capacity factor (k’) were examined. International Council for Harmonisation of Technical Requirements for Pharmaceuticals for Human Use (ICH) guidelines were followed to validate this analytical method (ICH, 2005ICH. Validation of analytical procedures: text and methodology Q2 (R1). In: International Conference on Harmonisation of Technical Requirements For Registration of Pharmaceuticals for Human Use. Geneva, Switzerland. 2005;11-12.) with the following essential parameters: specificity, linearity, limit of detection (LOD), limit of quantification (LOQ), precision, accuracy, and robustness.

Specificity

Specificity of the method is the capacity to identify the analyte of interest unequivocally, even in the presence of foreign compounds. This was ensured in the study by comparing chromatograms of blank NLCs and ACA-NLCs, using standard ACA solution as the reference. One milliliter each of the blank NLC and ACA-NLC suspension was subjected to centrifugation (3500 rpm, 1 h, 25 °C; KUBOTA 2800, Tokyo, Japan) in a Vivaspin 6 concentrator (10000 Da MWCO; Sartorius, Göttingen, Germany). The obtained precipitate was resuspended in 1 mL of acetonitrile, diluted 1:10 in acetonitrile, vortexed for 1 min, and filtered through 0.2-µm-pore nylon membrane syringe filters before analysis. The HPLC chromatograms were assessed for interfering peaks at the same retention time as the ACA standard solution. Ideally, the ACA-NLC but not the blank NLC should show a peak at the retention time corresponding to ACA.

Linearity

Linearity of the calibration curve was assessed using linear regression. Eight standard solutions of ACA at varying concentrations (5, 10, 20, 40, 60, 80, 100, and 200 µg/mL; n = 9) were prepared from the stock solution (1 mg/mL) via dilution with acetonitrile. Samples were then injected at a volume of 10 µL into the HPLC system. The peak area at each concentration was recorded and used for linear regression analysis of the slope from the plot of the mean peak area versus the analyte concentration. Because the correlation coefficient (r2) from the association between the x and y variables is insufficient to accept the linear regression model, a lack-of-fit test was conducted, and weighted linear regression with 1/x and 1/x2 weighting factors was used for the calibration curve (Sonawane et al., 2019Sonawane SS, Chhajed SS, Attar SS, Kshirsagar SJ. An approach to select linear regression model in bioanalytical method validation. J Anal Sci Technol. 2019;10:1-7.). Data analysis was conducted using Excel 2016 software (Microsoft Corporation, Redmond, WA, USA).

The weighting factor was selected by calculating the percent relative error (%RE) using equation 1, where CFOUND is the back-calculated concentration and CNOM is the nominal concentration that represents the standard.

% RE = [ ( C FOUND - C NOM ) / C NOM ] x 100 (1)

For the F-test, the experimental F-value (FEXP) was determined using equation 2, where MSLF represents the lack-of-fit mean square and MSPE represents the pure error mean square from analysis of variance. The FEXP values were then compared with the tabulated F-value (FTAB) (F6, 64) from the F-statistic distribution table at a 95% CI using c - 2 numerator and denominator degrees of freedom, where c represents 8 levels of concentrations and n represents 72 total data points.

F EXP = MSLF / MSPE (2)

The weighting factor with the lowest total %RE and FEXP < FTAB was selected for the standard calibration curve.

LOD and LOQ

The LOD is the smallest amount of analyte in a sample that can be detected but not required to be quantified, whereas the LOQ is the lowest amount of analyte that can be quantified with predefined accuracy and precision. The LOD and LOQ of the method were calculated based on the slope (S) of the calibration curve and least SD obtained from the response according to equations 3 and 4. This was done in accordance with ICH guidelines.

LOD = 3 . 3 x SD / S (3)

LOQ = 10 x SD / S (4)

Precision

Precision is the degree of agreement among multiple measurements of the same homogenous samples under standardized conditions. This can be done by evaluating the repeatability and intermediate precision of the method. Repeatability (intraday precision) was analyzed using standard ACA solutions of low (5 µg/mL), medium (40 µg/mL), and high (100 µg/mL) concentrations on the same day in triplicate. Intermediate precision (interday precision) was similarly investigated on three different days by quantifying the three concentration levels of standard ACA solutions in triplicate. The results were recorded as percent recovery and relative SD (RSD).

Accuracy

The accuracy of an analytical method is the closeness of agreement between the theoretical value and the experimental value. In the present study, the accuracy of the method was tested by spiking blank NLC suspensions with standard ACA solutions at concentrations of 5, 40, and 100 µg/mL. Samples were then mixed thoroughly, diluted 1:10 in acetonitrile, and analyzed. Analyses were performed in triplicate. The results were recorded as percent recovery and RSD.

Robustness

Robustness refers to the ability of an analytical method to remain unchanged by slight alterations in chromatographic parameters. This ensures reliability during normal applications. Robustness of the method was assessed by making minimal changes in experimental conditions, such as the mobile phase proportion (acetonitrile:water ratio, 81:19 and 79:21 [v/v]) and flow rate (0.75 mL/min and 0.85 mL/min). The percent recovery and RSD of standard ACA solutions at 5, 40, and 100 µg/mL were evaluated in triplicate.

Method applicability: drug-loading capacity and entrapment efficiency

ACA-NLCs with various ACA concentrations (0.1-1.0 mg/mL) were formulated in triplicate as described above. After overnight storage at 4 °C, 1 mL of an ACA-NLC suspension was subjected to centrifugation (3500 rpm, 1 h, 25 °C; KUBOTA 2800) in a Vivaspin 6 concentrator. The obtained precipitate was resuspended in 1 mL of acetonitrile, diluted 1:10 in acetonitrile, vortexed for 1 min, filtered through a 0.2-µm-pore filter, and run in the HPLC system to quantify the ACA content. The drug-loading capacity and entrapment efficiency of ACA-NLCs were then measured using equations 5 and 6, respectively, where MD refers to the mass of the drug in the nanoparticles, ML refers to the mass of the total lipid, CP refers to the concentration of the drug in the precipitate, and CI refers to the initial concentration of the drug added during the formulation of ACA-NLC.

DL ( % ) = M D / M L x 100 (5)

EE ( % ) = C P / C I × 100 (6)

RESULTS AND DISCUSSION

Method development

Preliminary runs consisted of optimizing the mobile phase composition in isocratic mode at different flow rates. In cases with erratic shaping and tailing of the ACA peak, we rejected the runs. The mixture of acetonitrile and water at the ratio of 80:20 (v/v) with a flow rate of 0.8 mL/min provided the most symmetrical ACA peak. We sought to determine the optimal wavelength for absorbance of ACA among wavelengths of 216, 226, and 254 nm based on previous studies (Niyomkam et al., 2010Niyomkam P, Kaewbumrung S, Kaewnpparat S, Panichayupakaranant P. Antibacterial activity of Thai herbal extracts on acne involved microorganism. Pharm Biol. 2010;48(4):375-80.; Haque et al., 2017Haque AKMM, Leong KH, Lo YL, Awang K, Nagoor NH. In vitro inhibitory mechanisms and molecular docking of 1’-S-1’-acetoxychavicol acetate on human cytochrome P450 enzymes. Phytomedicine. 2017;31:1-9.) and the wavelengths recommended by the ACA supplier. The highest absorbance was at 216 nm with a peak retention time of 4.06 min (Figure 2). The number of theoretical plates (N= 49,672) and capacity factor (k’ = 2.16) were in accordance with the specified limits (N>2000 and 2<k’<10).

FIGURE 2
HPLC chromatogram of 100 μg/mL standard ACA solution at UV detection wavelengths of a) 254 nm, b) 226 nm and c) 216 nm. Conditions: mobile phase acetonitrile:water ratio of 80:20 (v/v), flow rate of 0.8 mL/min, column temperature of 25 °C, sample temperature of 25 °C, and injection volume of 10 μL.

Method validation

Specificity

The specificity of the method was assessed by running blank NLCs and ACA-NLCs. ACA-NLCs had the characteristic ACA peak at 4.06 min, whereas the blank NLCs did not have a peak at the same retention time (Figure 3). This shows that the method is specific for ACA without the interference of other constituents from the nanoparticles. This is consistent with previous drug formulation studies that validated the specificity of their analytical methods. In all those tests, other components from the nanoparticle formulation did not interfere at the retention time for the analyte of interest (Lopes et al., 2017Lopes CE, Langoski G, Klein T, Ferrari PC, Farago PV. A simple HPLC method for the determination of halcinonide in lipid nanoparticles: Development, validation, encapsulation efficiency, and in vitro drug permeation. Braz J Pharm Sci. 2017;53(2):e15250.; Savadkouhi et al., 2017Savadkouhi MB, Vahidi H, Ayatollahi AM, Hooshfar S, Kobarfard F. RP-HPLC method development and validation for determination of eptifibatide acetate in bulk drug substance and pharmaceutical dosage forms. Iran J Pharm Res. 2017;16(2):490-497.).

FIGURE 3
HPLC chromatograms of a) precipitate from blank NLCs and b) precipitate from ACA-NLCs. Conditions: mobile phase acetonitrile:water ratio of 80:20 (v/v), flow rate of 0.8 mL/min, column temperature of 25 °C, sample temperature of 25

Linearity

We injected standard ACA solutions at eight concentrations ranging from 5 to 200 µg/mL into the HPLC system on 3 random days. The regression equation obtained for the unweighted calibration curve was y = 25.975x + 14.818, and the resulting correlation coefficient (r2) was 0.9995 (Figure 4). However, the total %RE calculated was 132.4%, and the FEXP for the linearity test was 120.7, which was significantly higher than the FTAB of 2.24 (Table I). This suggests a nonequal variance distribution of the standards range and indicates heteroscedasticity of the data. Hence, inaccurate results of subsequent analyses are possible if unweighted regression is used even though the r2 value is almost 1 (Boulanger et al., 2003Boulanger B, Chiap P, Dewe W, Crommen J, Hubert PH. An analysis of the SFSTP guide on validation of chromatographic bioanalytical methods: progresses and limitations. J Pharm Biomed Anal. 2003;32(4-5):753-65.).

Alternatively, we applied the 1/x and 1/x2 weighted regressions using the same data set of the ACA standard solutions to select the model with the lowest %RE and FEXP. The results are shown in Table I. The %RE for each weighted regression was lower than that for the unweighted model, but only the 1/x weighting factor exhibited FEXP < FTAB. Because the 1/x weighted calibration curve was more suited to homogenizing the variance of the residuals, as demonstrated in a previous study quantifying a different analyte (Sonawane et al., 2019Sonawane SS, Chhajed SS, Attar SS, Kshirsagar SJ. An approach to select linear regression model in bioanalytical method validation. J Anal Sci Technol. 2019;10:1-7.), we selected it for use in subsequent analyses. Therefore, we constructed the calibration curve with ACA standards at the concentration range of 5 to 200 µg/mL using a 1/x weighting factor.

FIGURE 4
Mean calibration curve of standard ACA solutions at concentrations ranging from 5 to 200 μg/mL (n=3)

TABLE I
Parameters for weighted and unweighted regression models for linearity

LOD and LOQ

We determined LOD and LOQ for the method based on the slope of the calibration curve and least SD obtained from the response. This was necessary to ensure that the LOD and LOQ can be reliably detected or quantified to prevent erroneous results when applying the method for future quantification (Armbruster, Pry, 2008Armbruster DA, Pry T. Limit of blank, limit of detection and limit of quantitation. Clin Biochem Rev. 2008;29(Suppl 1):S49-S52.). Using equations 3 and 4 in “Material and Methods”, the LOD was 0.59 µg/mL, and the LOQ was 1.79 µg/ mL. To confirm this, we subjected standard solutions of ACA at these concentrations to HPLC analysis. As a result, we detected 0.6 µg/mL ACA and could quantify 1.8 µg/mL ACA with a percent recovery of 100.1% and RSD of 1.85%. Similar to a previous study quantifying halcinonide in lipid nanoparticles, (Lopes et al., 2017Lopes CE, Langoski G, Klein T, Ferrari PC, Farago PV. A simple HPLC method for the determination of halcinonide in lipid nanoparticles: Development, validation, encapsulation efficiency, and in vitro drug permeation. Braz J Pharm Sci. 2017;53(2):e15250.), the LOD and LOQ were much smaller than the lowest standard ACA concentration used (5 µg/mL). This showed that the HPLC method is satisfactory for detecting and quantifying ACA within the concentration range of 5 to 200 µg/mL.

Precision

Precision is used to evaluate the degree of agreement between different test results when the method is used repeatedly with multiple samplings. We assessed intraday (repeatability) and interday (intermediate precision) runs of 5, 40, and 100 µg/mL (n = 3) standard ACA solutions for precision analysis (Tables II and III). The maximum RSD values were 1.91% for the repeatability test and 1.81% for the intermediate precision test. In addition, the percent recovery of ACA for all concentrations ranged from 98% to 104%, indicating that the relative error of the method was low even with repeated analysis. This agrees with previous studies that validated their method precision for quantification of different analyte concentrations according to ICH guidelines (Pecchio et al., 2014Pecchio M, Salman H, Irache JM, Renedo MJ, Dios-Viéitez MC. Development and validation of a HPLC method for the determination of cyclosporine a in new bioadhesive nanoparticles for oral administration. Indian J Pharm Sci. 2014;76(2):132-137.; Martins, Mainardes, 2017Martins LG, Mainardes RM. Application of a validated HPLC-PDA method for the determination of melatonin content and its release from poly (lactic acid) nanoparticles. J Pharm Anal. 2017;7(6):388-93.).

TABLE II
Repeatability for different concentrations of standard ACA solutions
TABLE III
Intermediate precision for different concentrations of standard ACA solutions

Accuracy

Accuracy is defined as the proximity of the results of an experiment to the true value. In this study, we tested the accuracy of the method by spiking blank NLC suspensions with 5, 40, and 100 µg/mL standard ACA solution. The percent recovery for all concentrations ranged from 97% to 101%, and the maximum RSD value was 0.32% (Table IV). The mean percent recovery for all concentrations was 100.02% (RSD, 2%). Hence, the HPLC method was ascertained to have low variability between the theoretical and experimental values of standard ACA solution concentrations at different levels. This is in accordance with the ICH guidelines for accuracy of an analytical method and agrees with the results of previous analytical method validation studies (ICH, 2005ICH. Validation of analytical procedures: text and methodology Q2 (R1). In: International Conference on Harmonisation of Technical Requirements For Registration of Pharmaceuticals for Human Use. Geneva, Switzerland. 2005;11-12.; Lopes et al., 2017Lopes CE, Langoski G, Klein T, Ferrari PC, Farago PV. A simple HPLC method for the determination of halcinonide in lipid nanoparticles: Development, validation, encapsulation efficiency, and in vitro drug permeation. Braz J Pharm Sci. 2017;53(2):e15250.).

TABLE IV
Accuracy measurement at different concentrations of standard ACA solutions

Robustness

The robustness of an analytical method is its capacity to be insignificantly affected by deliberate variations in chromatographic conditions. In this study, we evaluated the robustness of the method by making small changes to the mobile phase (acetonitrile:water) proportion and flow rate. The maximum RSD was 1.98% when we changed the mobile phase proportion to 79:21 for a 5 µg/mL sample of ACA standard solution. The percent recovery for all concentrations under different conditions ranged from 95% to 107%. None of the changes in the chromatographic conditions affected the RSD significantly (<2%) (Table V), which was consistent with findings of a previous study (Lopes et al., 2017Lopes CE, Langoski G, Klein T, Ferrari PC, Farago PV. A simple HPLC method for the determination of halcinonide in lipid nanoparticles: Development, validation, encapsulation efficiency, and in vitro drug permeation. Braz J Pharm Sci. 2017;53(2):e15250.). This demonstrated that the HPLC method is robust and reliable for use in future quantification of ACA content (Fontana, Bastos, Beck, 2010Fontana MC, Bastos MO, Beck RCR. Development and validation of a fast RP-HPLC method for the determination of clobetasol propionate in topical nanocapsule suspensions. J Chromatogr Sci. 2010;48(8):637-40).

TABLE V
Robustness results according to changes in mobile phase and flow rate

Method applicability

Determining the drug load and entrapment efficiency is essential to ensure that the drug-loading capacity of NLCs is maximized without excessively altering the entrapment of drugs during the formulation process. As shown in Table VI, all of the NLC formulations we used had greater than 87% entrapment efficiency. We did not observe a significant trend in entrapment efficiency with increasing ACA content. As for the drug-loading capacity of the NLC, increasing the drug load was feasible with our NLC formulation, as ACA exhibited a high affinity towards the lipid matrix. Previous studies demonstrated that the entrapment efficiency of their NLC system can be maximized by increasing the drug content (Negi, Jaggi, Talegaonkar, 2013Negi LM, Jaggi M, Talegaonkar SA. Logical approach to optimize the nanostructured lipid carrier system of irinotecan: Efficient hybrid design methodology. Nanotechnology. 2013;24(1): 015104.; Ferreira et al., 2015Ferreira M, Chaves LL, Lima SA, Reis S. Optimization of nanostructured lipid carriers loaded with methotrexate: A tool for inflammatory and cancer therapy. Int J Pharm. 2015;492(1-2):65-72.). Although a 5% drug load was the target of the present study, future studies can be conducted to determine the maximum drug-loading capacity of the NLCs without severely affecting their entrapment efficiency.

In addition, we carried out the drug-loading capacity and entrapment efficiency studies by separating the supernatant from the ACA-NLC nanosuspension and measuring the ACA content in the precipitate. This is also referred to as the direct method of quantification, in which the drug content in the lipid phase of the precipitate is measured. The nanoparticle precipitate is solubilized in acetonitrile, vortexed to disrupt the lipid matrix, filtered, and subjected to HPLC for component separation (Gaikwad et al., 2019Gaikwad VL, Choudhari PB, Bhatia NM, Bhatia MS. Characterization of pharmaceutical nanocarriers: in vitro and in vivo studies. In: Grumezescu AM, editors. Nanomaterials for Drug Delivery and Therapy. Norwich, NY: William Andrew Publishing; 2019. p. 33-58.). Contrary to previous studies that usually used the supernatant to measure drug content in NLCs (Lopes et al., 2017Lopes CE, Langoski G, Klein T, Ferrari PC, Farago PV. A simple HPLC method for the determination of halcinonide in lipid nanoparticles: Development, validation, encapsulation efficiency, and in vitro drug permeation. Braz J Pharm Sci. 2017;53(2):e15250.; Martins, Mainardes, 2017Martins LG, Mainardes RM. Application of a validated HPLC-PDA method for the determination of melatonin content and its release from poly (lactic acid) nanoparticles. J Pharm Anal. 2017;7(6):388-93.), this method prevented technical errors due to the use of highly diluted supernatant samples. It proved to be effective because other components of the lipid phase did not interfere with drug quantification as observed in the specificity chromatograms (Figure 3). Furthermore, no other time-consuming extraction methods were required to isolate ACA from the NLC suspensions because the percent analyte recovery in the accuracy test was high. This means that ACA can be separated from the lipid components of the NLC suspensions during chromatographic runs and does not have to be extracted beforehand.

TABLE VI
Drug-loading capacity and entrapment efficiency of ACA-NLCs

CONCLUSION

This report describes our development of a consistent and viable reversed-phase HPLC method, which is essential for determination of the ACA content in NLCs. Although researchers have isolated and quantified free ACA using HPLC under different chromatographic conditions, this is the first documented study of ACA quantification in nanoparticles. The presence of foreign substances in nanoparticle formulations makes it essential that we perform an accurate separation technique with the NLC sample before quantification of the drug. In the present study, we developed a sensitive and quick method of qualitative and quantitative determination of ACA content in lipid nanoparticles without any drug extraction method. Because a short retention time is preferred in analysis of pharmaceutical compounds, our HPLC method is suitable for entrapment efficiency studies involving numerous samples. Overall, we successfully validated this method according to all ICH guidelines and showed that it can be used in future studies involving ACA-loaded NLCs.

ACKNOWLEDGMENTS

The authors acknowledge the Centre for Research in Biotechnology for Agriculture (CEBAR) for supporting this paper through their Research University grant TU002F-2018. The assistance of the scientific editor Donald R. Norwood in Editing Services, Research Medical Library at MD Anderson Cancer Center is also gratefully acknowledged.

REFERENCES

  • Armbruster DA, Pry T. Limit of blank, limit of detection and limit of quantitation. Clin Biochem Rev. 2008;29(Suppl 1):S49-S52.
  • Arshad NM, In LLA, Soh TL, Mohamad NA, Ibrahim H, Awang K, et al. Recombinant human alpha fetoprotein synergistically potentiates the anti-cancer effects of 1′-S-1′-acetoxychavicol acetate when used as a complex against human tumours harbouring AFP-receptors. Oncotarget. 2015;6(18):16151-67.
  • Awang K, Azmi MN, Aun LI, Aziz AN, Ibrahim H, Nagoor NH. The apoptotic effect of 1’-S-1’-acetoxychavicol acetate from Alpinia conchigera on human cancer cells. Molecules. 2010;15(11):8048-8059.
  • Baradwaj RG, Rao MV, Kumar TS. Novel purification of 1’-S-1’-Acetoxychavicol acetate from Alpinia galanga and its cytotoxic plus antiproliferative activity in colorectal adenocarcinoma cell line SW480. Biomed Pharmacother. 2017;91:485-493.
  • Barik BR, Kundu AB, Dey AK. Two phenolic constituents from Alpinia galanga rhizomes. Phytochemistry. 1987;26(7):2126-27.
  • Batra V, Syed Z, Gill JN, Coburn MA, Adegboyega P, DiGiovanni J, et al. Effects of the tropical ginger compound, 1’-acetoxychavicol acetate, against tumor promotion in K5. Stat3C transgenic mice. J Exp Clin Cancer Res. 2012;31(1):57.
  • Beloqui A, Solinis MA, Rodriguez-Gascon A, Almeida AJ, Preat V. Nanostructured lipid carriers: Promising drug delivery systems for future clinics. Nanomedicine. 2016;12(1):143-61.
  • Boulanger B, Chiap P, Dewe W, Crommen J, Hubert PH. An analysis of the SFSTP guide on validation of chromatographic bioanalytical methods: progresses and limitations. J Pharm Biomed Anal. 2003;32(4-5):753-65.
  • Ferreira M, Chaves LL, Lima SA, Reis S. Optimization of nanostructured lipid carriers loaded with methotrexate: A tool for inflammatory and cancer therapy. Int J Pharm. 2015;492(1-2):65-72.
  • Fontana MC, Bastos MO, Beck RCR. Development and validation of a fast RP-HPLC method for the determination of clobetasol propionate in topical nanocapsule suspensions. J Chromatogr Sci. 2010;48(8):637-40
  • Gaikwad VL, Choudhari PB, Bhatia NM, Bhatia MS. Characterization of pharmaceutical nanocarriers: in vitro and in vivo studies. In: Grumezescu AM, editors. Nanomaterials for Drug Delivery and Therapy. Norwich, NY: William Andrew Publishing; 2019. p. 33-58.
  • Haque AKMM, Leong KH, Lo YL, Awang K, Nagoor NH. In vitro inhibitory mechanisms and molecular docking of 1’-S-1’-acetoxychavicol acetate on human cytochrome P450 enzymes. Phytomedicine. 2017;31:1-9.
  • ICH. Validation of analytical procedures: text and methodology Q2 (R1). In: International Conference on Harmonisation of Technical Requirements For Registration of Pharmaceuticals for Human Use. Geneva, Switzerland. 2005;11-12.
  • In LLA, Arshad NM, Ibrahim H, Azmi MN, Awang K, Nagoor NH. 1’-Acetoxychavicol acetate inhibits growth of human oral carcinoma xenograft in mice and potentiates cisplatin effect via proinflammatory microenvironment alterations. BMC Complement Altern Med. 2012;12:179.
  • Ito K, Nakazato T, Murakami A, Yamato K, Miyakawa Y, Yamada T, et al. Induction of apoptosis in human myeloid leukemic cells by 1′-acetoxychavicol acetate through a mitochondrial-and Fas-mediated dual mechanism. Clin Cancer Res. 2004;10(6):2120-30.
  • Janssen AM, Scheffer JJ. Acetoxychavicol acetate, an antifungal component of Alpinia galangal. Planta Med.1985;51(6):507-11.
  • Kaur A, Singh R, Dey CS, Sharma SS, Bhutani KK, Singh IP. Antileishmanial phenylpropanoids from Alpinia galanga (Linn.) Willd. Indian J Exp Biol. 2010;48(3):314-7.
  • Kondo A, Ohigashi H, Murakami A, Jiwajinda S, Koshimizu K. A potent inhibitor of tumor promoter-induced Epstein-Barr virus activation, 1’-acetoxychavicol acetate from Languas galanga, a traditional Thai condiment. Biosci Biotechnol Biochem.1993;57(8):1344-45.
  • Lopes CE, Langoski G, Klein T, Ferrari PC, Farago PV. A simple HPLC method for the determination of halcinonide in lipid nanoparticles: Development, validation, encapsulation efficiency, and in vitro drug permeation. Braz J Pharm Sci. 2017;53(2):e15250.
  • Martins LG, Mainardes RM. Application of a validated HPLC-PDA method for the determination of melatonin content and its release from poly (lactic acid) nanoparticles. J Pharm Anal. 2017;7(6):388-93.
  • Müller RH, Radtke M, Wissing SA. Solid lipid nanoparticles (SLN) and nanostructured lipid carriers (NLC) in cosmetic and dermatological preparations. Adv Drug Deliv Rev. 2002;54(Suppl 1):S131-55.
  • Nakamura Y, Murakami A, Ohto Y, Torikai K, Tanaka T, Ohigashi H. Suppression of tumor promoter-induced oxidative stress and inflammatory responses in mouse skin by a superoxide generation inhibitor 1’-acetoxychavicol acetate. Cancer Res. 1998;58(21):4832-9.
  • Narukawa M, Koizumi K, Iwasaki Y, Kubota K, Watanabe T. Galangal pungent component, 1’-acetoxychavicol acetate, activates TRPA1. Biosci Biotechnol Biochem . 2010;74(8):1694-6.
  • Negi LM, Jaggi M, Talegaonkar SA. Logical approach to optimize the nanostructured lipid carrier system of irinotecan: Efficient hybrid design methodology. Nanotechnology. 2013;24(1): 015104.
  • Niyomkam P, Kaewbumrung S, Kaewnpparat S, Panichayupakaranant P. Antibacterial activity of Thai herbal extracts on acne involved microorganism. Pharm Biol. 2010;48(4):375-80.
  • Pathak Y. Recent developments in nanoparticulate drug delivery systems. In: Pathak Y, Thassu D, editors. Drug Delivery Nanoparticles Formulation and Characterization (1st ed.). Boca Raton, FL: CRC Press; 2009. p. 1-7.
  • Pecchio M, Salman H, Irache JM, Renedo MJ, Dios-Viéitez MC. Development and validation of a HPLC method for the determination of cyclosporine a in new bioadhesive nanoparticles for oral administration. Indian J Pharm Sci. 2014;76(2):132-137.
  • Savadkouhi MB, Vahidi H, Ayatollahi AM, Hooshfar S, Kobarfard F. RP-HPLC method development and validation for determination of eptifibatide acetate in bulk drug substance and pharmaceutical dosage forms. Iran J Pharm Res. 2017;16(2):490-497.
  • Severino P, Santana MH, Souto EB. Optimizing SLN and NLC by 2(2) full factorial design: Effect of homogenization technique. Mater Sci Eng C Mater Biol Appl. 2012;32(6):1375-9.
  • Sonawane SS, Chhajed SS, Attar SS, Kshirsagar SJ. An approach to select linear regression model in bioanalytical method validation. J Anal Sci Technol. 2019;10:1-7.
  • Tanaka T, Kawabata K, Kakumoto M, Makita H, Matsunaga K, Mori H, et al. Chemoprevention of azoxymethane-induced rat colon carcinogenesis by a xanthine oxidase inhibitor, 1’-acetoxychavicol acetate. Jpn JCancer Res .1997;88(9):821-30.
  • Tanaka T, Makita H, Kawamori T, Kawabata K, Mori H, Murakami A, et al. A xanthine oxidase inhibitor 1’-acetoxychavicol acetate inhibits azoxymethane-induced colonic aberrant crypt foci in rats. Carcinogenesis. 1997;18(5):1113-8.
  • Yang X, Eilerman RG. Pungent principal of Alpinia galangal (L.) swartz and its applications. J Agric Food Chem.1999;47(4):1657-62.

Publication Dates

  • Publication in this collection
    16 Jan 2023
  • Date of issue
    2022

History

  • Received
    27 July 2020
  • Accepted
    19 Jan 2021
Universidade de São Paulo, Faculdade de Ciências Farmacêuticas Av. Prof. Lineu Prestes, n. 580, 05508-000 S. Paulo/SP Brasil, Tel.: (55 11) 3091-3824 - São Paulo - SP - Brazil
E-mail: bjps@usp.br