EJPB 11919

No. of Pages 10, Model 5G

19 May 2015 European Journal of Pharmaceutics and Biopharmaceutics xxx (2015) xxx–xxx 1

Contents lists available at ScienceDirect

European Journal of Pharmaceutics and Biopharmaceutics journal homepage: www.elsevier.com/locate/ejpb

2

Research Paper

7 4 8

Optimization of controlled release nanoparticle formulation of verapamil hydrochloride using artificial neural networks with genetic algorithm and response surface methodology

5 6 9 10 11 12 13 15 14 16 1 2 8 3 19 20 21 22 23 24 25 26 27 28 29 30 31

Yongqiang Li a, Mohammadreza R. Abbaspour a, Paul V. Grootendorst a,b, Andrew M. Rauth c, Xiao Yu Wu a,⇑ a

Advanced Pharmaceutics and Drug Delivery Laboratory, Leslie Dan Faculty of Pharmacy, University of Toronto, Toronto, ON M5S 3M2, Canada Department of Economics, McMaster University, Hamilton, ON L8S 4M4, Canada c Faculty of Medicine, University of Toronto, ON M5G 2M9, Canada b

a r t i c l e

i n f o

Article history: Received 2 March 2015 Revised 17 April 2015 Accepted in revised form 27 April 2015 Available online xxxx Keywords: Optimization Artificial neural networks Continuous genetic algorithm Response surface methodology Polymer–lipid hybrid nanoparticles Verapamil hydrochloride

a b s t r a c t This study was performed to optimize the formulation of polymer–lipid hybrid nanoparticles (PLN) for the delivery of an ionic water-soluble drug, verapamil hydrochloride (VRP) and to investigate the roles of formulation factors. Modeling and optimization were conducted based on a spherical central composite design. Three formulation factors, i.e., weight ratio of drug to lipid (X1), and concentrations of Tween 80 (X2) and Pluronic F68 (X3), were chosen as independent variables. Drug loading efficiency (Y1) and mean particle size (Y2) of PLN were selected as dependent variables. The predictive performance of artificial neural networks (ANN) and the response surface methodology (RSM) were compared. As ANN was found to exhibit better recognition and generalization capability over RSM, multi-objective optimization of PLN was then conducted based upon the validated ANN models and continuous genetic algorithms (GA). The optimal PLN possess a high drug loading efficiency (92.4%, w/w) and a small mean particle size (100 nm). The predicted response variables matched well with the observed results. The three formulation factors exhibited different effects on the properties of PLN. ANN in coordination with continuous GA represent an effective and efficient approach to optimize the PLN formulation of VRP with desired properties. Ó 2015 Elsevier B.V. All rights reserved.

33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50

51 52

1. Introduction

53

Verapamil hydrochloride (VRP) is a calcium channel blocker commonly used to treat hypertension, cardiac arrhythmias and angina [5,16]. It has also been used as a P-glycoprotein (P-gp)-inhibitor to enhance chemotherapy of multidrug resistant cancer [7,27,29]. Although VRP is well absorbed through the gastrointestinal tract (P90%), its systemic bioavailability (20–35%) is

54 55 56 57 58

Abbreviations: PLN, polymer–lipid hybrid nanoparticles; VRP, verapamil hydrochloride; DS, dextran sulfate sodium; DA, dodecanoic acid; VRP-PLN, verapamil hydrochloride loaded polymer–lipid hybrid nanoparticles; ITC, isothermal titration calorimetry; TEM, transmission electron microscopy; CCD, central composite design; ANN, artificial neural networks; RSM, response surface methodology; GA, genetic algorithms; AIC, Akaike’s information criterion; DLE, drug loading efficiency; DLC, drug loading capacity; HLB, hydrophile–lipophile balance; DDI water, distilled and deionized water. ⇑ Corresponding author. Tel.: +1 (416) 978 5272; fax: +1 (416) 978 8511. E-mail address: [email protected] (X.Y. Wu).

poor and variable due to the extensive first-pass metabolism by cytochrome CYP450 3A4 [31]. Its short elimination half-life (2.8– 7.4 h) demands dosing of 3–4 times per day [17]. Hence a sustained release nanoparticle delivery system able to bypass liver metabolism would increase bioavailability and reduce dosing frequency. As lipids are absorbed mainly through the lymphatic system in the intestines, lipid-based particulate systems offer a promising approach to bypassing first-pass metabolism via lymphatic transport thus enhancing drug bioavailability [9,21,35,36,39]. In particular, solid lipid nanoparticles (SLN) [37] have been reported to improve oral bioavailability and optimize plasma profiles of loaded drugs [6,8,30,32,34,52]. Owing to the hydrophobic nature, SLN is not an ideal carrier for sustained release of a highly water-soluble drug like VRP with an aqueous solubility of 83 mg/mL. In order to deliver ionic water-soluble drugs while maintaining the advantages of SLN, our group exploited the characteristic physicochemical properties of polymer counterions and developed a novel polymer–lipid

http://dx.doi.org/10.1016/j.ejpb.2015.04.028 0939-6411/Ó 2015 Elsevier B.V. All rights reserved.

Please cite this article in press as: Y. Li et al., Optimization of controlled release nanoparticle formulation of verapamil hydrochloride using artificial neural networks with genetic algorithm and response surface methodology, Eur. J. Pharm. Biopharm. (2015), http://dx.doi.org/10.1016/j.ejpb.2015.04.028

59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76

EJPB 11919

No. of Pages 10, Model 5G

19 May 2015 2 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142

Y. Li et al. / European Journal of Pharmaceutics and Biopharmaceutics xxx (2015) xxx–xxx

hybrid nanoparticle (PLN) system [51]. This system exhibited sufficient drug loading, lowered initial burst release and subsequently sustained drug release. To further improve the performance of PLN, drug–polymer–lipid interactions, internal structure, drug loading and release mechanisms were investigated [25]. These efforts resulted in a rationally designed formulation with high drug loading capacity and sustained release kinetics [26]. Nevertheless, the variables of PLN formulation (e.g. drug to lipid ratio and surfactant concentrations) need to be further optimized in order to obtain relatively uniform particle size in the range of 100 nm for effective intestinal lymphatic uptake [12,46] while maintaining maximal drug loading levels. Our preliminary tests suggested that for PLN, the drug loading capacity and particle size depended on the processing conditions and formulation factors including the weight ratio of drug to lipid, and the amounts and ratio of surfactants (e.g. Pluronic F68 and Tween 80). Owing to multiple variables and their unclear cause-and-effect relationship, we applied factorial experiment design [3,41,42,53] and numerical modeling techniques to optimize the PLN formulation. The central composite design (CCD), a second-order polynomial model [38] was employed, as CCD is very efficient and able to provide much information on the effects of experiment variables and overall experimental error with a minimum number of required runs. Combining the factorial design with statistical modeling techniques provides a particularly useful way of understanding the underlying structure between the independent and dependent variables. Models, established by the response surface methodology (RSM) or artificial neural networks (ANN) [18], can then be used to predict the responses to the combinations of the independent factors that are not explicitly studied experimentally. Due to the simplicity and ease of deployment, RSM utilizing a quadratic equation to fit the experimental data has been widely used in pharmaceutical research [33,50]. Compared to RSM, ANN are less restrictive and hence suitable for recognizing more complex, multi-dimensional and non-linear patterns [4,19,20,40,45,48]. Since different modeling strategies are adopted in RSM and ANN, and each method has its own inherent advantages and limitations, we elected to compare their modeling performance in the optimization of PLN in this work. Once the model is established, contour plots are normally constructed to visualize the relationship between independent and dependent variables so as to search the optimal formulation manually. However, three independent variables and two dependent variables are involved in the optimization of PLN in this study. Identifying an optimal formulation of this system using the superimposing contour method is cumbersome. To overcome this difficulty, the equations for each of the dependent variables were incorporated into a single equation, i.e., a generalized distance function [22], which is called a ‘cost function’ or an objective function in the optimization process. Then a numerical optimizer can be used to locate the values of the independent variables that minimize the cost function. We employed the continuous genetic algorithms (GAs) [11,44,15] to minimize the cost function with respect to the three independent variables. GA is a robust, stochastic, adaptive heuristic searching algorithm. It can avoid the problem of being stuck in local minima which is common to the gradient-based algorithms. Moreover, continuous GA is better than binary GA as it utilizes original experimental data as input and thus can avoid the loss of information which takes place in the transformation of data. In summary, we used a three-factor spherical second-order CCD to map the underlying pattern between each of the dependent variables (particle size and drug loading of PLN) and independent variables (drug to lipid ratio and concentrations of Pluronic F68 and Tween 80). RSM and ANN were used to model and predict

the responses. Continuous GA was utilized to perform the search for a global optimum.

143

2. Materials and methods

145

2.1. Materials

146

Verapamil HCl (pKa = 8.6, solubility: 83 mg/mL), dodecanoic acid (DA, MP: 44–46 °C), dextran sulfate sodium (DS, MW: 5000 Da) and Tween 80 were purchased from Sigma–Aldrich Canada (Oakville, ON, Canada). Calcium chloride, potassium dihydrogen phosphate and di-sodium hydrogen orthophosphate anhydrous (dibasic) were purchased from Fisher Chemicals (Pittsburgh, PA, USA). Pluronic F68 was a gift from BASF (Mississauga, ON, Canada). AmiconÒ centrifugal filters (MWCO: 30,000 Da) were acquired from Millipore Inc. (Toronto, ON, Canada). Distilled and deionized (DDI) water was prepared with a Milli-Q water purifier (Milli-Pore, Etobicoke, ON, Canada).

147

2.2. Preparation of VRP-PLN

158

PLN were prepared by a modified micro-emulsion method followed by ultrasonication [51]. The process of PLN formation is schematically illustrated in Fig. 3. Briefly, the lipid, DA, was melted using a thermostated water bath at 50 °C. A particular amount of VRP (dependent on the desired drug loading capacity of PLN) was added to the molten lipid. Under constant magnetic stirring with a Corning stirring plate, an aqueous phase was prepared by dissolving specific amounts of Pluronic F68 and Tween 80 in DDI water and heating to the same temperature as the lipid phase. The hot aqueous phase was added to the lipid phase subsequently. A known amount of DS was added slowly to the mixture to obtain an ionic molar ratio of DS to VRP at 1.0. The coarse o/w micro-emulsion was left in the water bath for 20 min under vigorous stirring at 700 rpm and then sonicated for another 5 min at the same temperature. The PLN nanoparticles were obtained by injecting the final emulsion into a fixed volume of cold water (2–4 °C) under magnetic stirring.

159

2.3. Measurements of particle size and surface charge

176

The hydrodynamic diameters and surface charge of PLN were measured by a particle sizer equipped with the function for zeta potential measurements (Nicomp, Model 380 ZLS, Particle Sizing Systems Inc., Santa Barbara, CA, USA). All values were measured in DDI water at a fixed angle of 90° at 25 °C in 10 mm diameter cells with a He/Ne laser light source at 632.8 nm. Each experiment was performed in triplicate.

177

2.4. Determination of drug loading efficiency

184

The drug loading efficiency of the system was determined by measuring the concentration of free drug remained in the dispersion medium, compared to the initial drug added. Briefly, 3 mL of undiluted sample of PLN was placed in the sample chamber of a Millipore filter consisted of a membrane with MWCO of 30,000 Da. The unit was centrifuged at 4000 rpm (relative centrifugal force: 3150g for 45 min). The PLN containing encapsulated drug–polymer complex remained in the sample chamber and the aqueous phase passed into the recovery chamber through the filter membrane. The aqueous filtrate (i.e., first filtrate from centrifugation) was collected, the recovery chamber was washed three times using DDI water to remove uncapsulated drug, 1 mL of the 3 diluted filtrate was incubated with the same volume of 0.3 M CaCl2 solution for 24 h at room temperature. The concentrated

185

Please cite this article in press as: Y. Li et al., Optimization of controlled release nanoparticle formulation of verapamil hydrochloride using artificial neural networks with genetic algorithm and response surface methodology, Eur. J. Pharm. Biopharm. (2015), http://dx.doi.org/10.1016/j.ejpb.2015.04.028

144

148 149 150 151 152 153 154 155 156 157

160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175

178 179 180 181 182 183

186 187 188 189 190 191 192 193 194 195 196 197 198

EJPB 11919

No. of Pages 10, Model 5G

19 May 2015 3

Y. Li et al. / European Journal of Pharmaceutics and Biopharmaceutics xxx (2015) xxx–xxx 199 200 201 202 203 204 205 206

207 209

CaCl2 solution was used to displace all ionically-bound VRP in the VRP-DS complex; divalent Ca2+ is more effective than monovalent Na+ as it can exchange two VRP+ ions out of the VRP-DS complex [43,1,24,26,28,51]. The VRP content in the aqueous filtrate was analyzed with a UV–vis spectrophotometer (Hewlett–Packard 8452A, Palo Alto, CA) at 278 nm. Drug loading efficiency (DLE) and drug loading capacity (DLC) were calculated using the following equations:

DLE ð%Þ ¼

W initial drug added  W free drug  100 W initial drug added

ð1Þ

DLC ð%Þ ¼

W drug in PLN  100 W lipid

ð2Þ

210 212

216

where Winitial drug added refers to the amount of the drug initially added, Wfree drug is the amount of drug that has not been incorporated into PLN, Wdrug in PLN is the amount of drug incorporated into the PLN, and Wlipid is the weight of lipid phase.

217

2.5. Transmission electron microscopy

218

223

The morphology of PLN particulates was examined by transmission electron microscopy (TEM) (Hitachi 7000, Tokyo, Japan). The samples were stained with 2% (w/v) negatively charged phosphotungstic acid (PTA) for 2 min before TEM imaging. 12.4% drug loading level of PLN (w/w, with respect to the lipid) was used in this study.

224

2.6. Experimental design

225

Three formulation factors, i.e., weight ratio of drug to lipid (X1), concentrations of Tween 80 (X2) and Pluronic F68 (X3), were chosen as independent variables. Drug loading efficiency (Y1) and mean particle size (Y2) of PLN were selected as dependent variables. A spherical central composite design was constructed to perform 15 experiments as shown in Fig. 1 and Table 1. This design is also called the three-factor spherical second-order composite experimental design [23,48]. Of the 15 runs, the first eight

213 214 215

219 220 221 222

226 227 228 229 230 231 232

Table 1 Spherical center composite design (CCD) for three factors with three levels. Formulation no.

Code of drug to lipid ratio (w/w)

Code of conc. of Tween 80

Code of conc. of Pluronic F68

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

1 1 1 1 1 1 1 1 1.73 1.73 0 0 0 0 0

1 1 1 1 1 1 1 1 0 0 1.73 1.73 0 0 0

1 1 1 1 1 1 1 1 0 0 0 0 1.73 1.73 0

experiments (formulations # 1–8) constituted a factorial design of three factors (X1, X2, and X3), each evaluated at two levels (1, 1), i.e., 23 experiments. In the other three levels, 0 represents a middle level between the factorial points, 1 and 1, and the axial points, a = ±31/2 = ±1.73 represent the extreme levels. With the addition of the center point and the axial points, the experiments include five levels for each factor (see Table 2). Note that the formulation #15 is a repetition of the center experiment (n = 6) and is expressed as the mean ± S.D. (standard deviation). The unique feature of this three-factor spherical second-order composite design is that each experimental point has the same distance to the center experiment, so that all point estimates are estimated with the same degree of precision (see Fig. 2).

233

2.7. ANN modeling and structure

246

Our ANN model expresses Y1 and Y2 as the following function of X1, X2 and X3:

247

Anionic polymer

Drug-polymer complex

hq ¼

+

+

ehij þe 1 þ ehij

j ¼ 1; 2

ð3Þ

where a is the threshold (or bias) value of each logit (neuron) to fire, k is the number of logits being averaged, /ij is the weight attached to the ith logit, hij is a linear function of exploratory variables X1, X2, X3 (as defined in Eq. (4)), and e is the ‘error term’, i.e. the effect on Yj of all other factors, besides those that are explicitly modeled. Therefore, Yj is modeled as a weighted sum of several sigmoids (logit curves), each of which is a function of the independent variables, X1, X2, X3. Specifically, the sum of input (hq) to the nodes in the hidden layer (i.e., the layer between the input and output layers) is expressed as

Cationic drug

Pluronic F68 Tween 80

Melted lipid

X

236 237 238 239 240 241 242 243 244 245

248

251 252 253 254 255 256 257 258 259 260 261

262

wpq xq

ð4Þ

264

where wpq is the weight connection between the node q in the present layer (hidden layer or unknown parameters) and the node p in the previous layer (input layer or independent variables) and xq is

265

Table 2 Levels of causal factors in physical form. Factor

Nanoparticles Fig. 1. Schematic illustration of preparation of polymer–lipid hybrid nanoparticles (PLN) and the processes for drug loading.

235

249 k X Yj ¼ a þ /ij i¼1

+

234

Ratio of VRP to DA (%, w/w) Conc. of Tween 80 (%, w/v) Conc. of Pluronic F68 (%, w/v)

Factor level in coded form 1.73

1

0

1

1.73

5.00 1.50 1.50

12.38 1.92 1.92

22.50 2.50 2.50

32.62 3.08 3.08

40.00 3.50 3.50

Please cite this article in press as: Y. Li et al., Optimization of controlled release nanoparticle formulation of verapamil hydrochloride using artificial neural networks with genetic algorithm and response surface methodology, Eur. J. Pharm. Biopharm. (2015), http://dx.doi.org/10.1016/j.ejpb.2015.04.028

266 267

EJPB 11919

No. of Pages 10, Model 5G

19 May 2015 4

Y. Li et al. / European Journal of Pharmaceutics and Biopharmaceutics xxx (2015) xxx–xxx #14

#4

#8

#2

#6

#12 #9

#10 #11

X3

#15 #3

X2

#7

X1

#1

#5

#13

Simulations using different Nhidden values demonstrated that for drug loading efficiency (Y1), an ANN model with 2 or 3 hidden nodes had close AIC values, i.e., 47.54 and 40.31, respectively. In comparison, the ANN models for mean particle size (Y2) with 2 or 3 hidden nodes exhibited distinct AIC values of 2.74 and 25.06, respectively. Therefore, as an optimal ANN structure, the ANN model with 2 nodes in the hidden layer was selected for the modeling of drug loading efficiency (Y1) and a model with 3 nodes in the hidden layer for mean particle size (Y2) in the subsequent analysis (see Fig. 3).

314

2.8. Response surface methodology

324

Classic response surface methodology (RSM) employs a second-order polynomial regression equation to predict the response variables [14]. A general quadratic model with three variables is defined as follows:

325

Fig. 2. Diagram of three-factor second-order composite experimental design.

FðXÞ ¼ a0 þ 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286

the output value from the previous layer (input layer or independent variables). In this study, three formulation factors corresponding to the weight ratio of drug to lipid (X1), the concentration of Tween 80 (X2) and the concentration of Pluronic F68 (X3) constituted the units of the input layer. The output layer was composed of the response variables of drug loading efficiency (Y1) and mean particle size (Y2) of each formulation. ANN models for the optimization of PLN were developed with a simulator that combines an ANN module for modeling and prediction and a continuous GA module for multi-objective optimization. A partitioned ANN structure proposed to model dependent variables individually was applied in this study to address the problem of the limited availability of experimental training pairs, while still achieving reasonably good generalization capability of ANN models [49]. The number of hidden nodes in the hidden layer, determined by the number of training pairs, was estimated according to the equation introduced by Carpenter and Hoffman [10] as follows:

287 289

Nhidden ¼

N sample =b  Noutput Ninput þ Noutput þ 1

ð5Þ

305

where Nhidden is the number of hidden nodes (i.e., different logits); Nsample is the number of training data pairs; Ninput is the number of input nodes and Noutput is the number of output nodes (i.e., output variables being modeled). b refers to the degree of determination of the network and its value is determined by the number of the inputs, the outputs and the size of data available for training. b has three values: b < 1, under-determined; b = 1, exactly determined and b > 1, over-determined. To prevent the occurrence of overfitting, b P 1 is preferred in modeling practice. In this study, Nsample = 15, Ninput = 3 and Noutput = 1 for ANN models approximating each response variables separately. Thus Nhidden was calculated from Eq. (5) to be 3 for b = 0.94 and 2 for b = 1.36, respectively. The optimality of ANN models with different Nhidden values was evaluated using Akaike’s information criterion (AIC) [2,13] as follows:

306 308

AIC ¼ nS  lnðSSÞ þ 2  nW

290 291 292 293 294 295 296 297 298 299 300 301 302 303 304

309 310 311 312 313

3 X

3 X

i¼1

i¼1

bi X i þ

cii X 2i þ

þ eijk X i X j X k þ e;

XX

316 317 318 319 320 321 322 323

326 327 328

329

dij X i X j

i 0.025). The optimized PLN possessed a zeta potential of 20.57 ± 0.32 mV (n = 3), close to that of the solid lipid nanoparticles without the negatively charged polymer DS, suggesting the relative stability of this formulation and the complete neutralization of sulfate groups on DS.

479

3.6. Effect of weight ratio of drug to lipid

480

Fig. 7a and b illustrates the effect of drug to lipid weight ratio (X1) on the drug loading efficiency of PLN (Y1) when the

428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462

466 467 468 469 470 471 472 473 474 475 476 477

481

7

concentration of Tween 80 (X2) or Pluronic F68 (X3) was fixed at 1.92% (w/v, coded form: 1). The predicted results indicate that, within the studied range of surfactant concentrations, i.e., from 1.50% to 3.50% (w/v), the drug loading efficiency of PLN essentially remains high and steady as the drug to lipid ratio increases from 5.00% (code form: -1.73) to 32.6% (code form: 1), except at a high drug to lipid ratio (40.0%, code form: 1.73) and high surfactant concentrations. The drug loading efficiency of PLN decreases sharply to 75.8% in a formulation consisting of 40% VRP, 1.92% (w/v) Tween 80 and 3.5% (w/v) Pluronic F68 (coded forms: 1.73, 1, 1.73) (Fig. 7a), and to 55.9% with 3.5% (w/v) Tween 80 and 1.92% (w/v) Pluronic F68 (coded forms: 1.73, 1.73, 1) (Fig. 7b). The relatively high drug loading efficiency of PLN at a wide range of weight ratio of drug to lipid (X1) is consistent with the compatibility between VRP-DS complex and DA found in previous work [25,26]. The dramatic drop in the loading efficiency at a drug to lipid ratio higher than 32.6% and high surfactant levels may be due to the exhaustion of encapsulation capacity of the lipid phase and increased solubilization power of the surfactants. Unlike the loading efficiency, the mean particle size of PLN (Y2) is sensitive to the weight ratio of drug to lipid (X1) and shows a complex pattern (Fig. 8). The mean particle size of PLN (Y2) decreases with increasing drug to lipid ratio initially up to 22.5% (coded forms: 0), and then increases till a peak value at 32.6% (code form: 1) before decreasing again.

Table 5 Predicted and experimental response variables for the optimal formulation.

a

Response

Predicted

Experimentala

Drug loading efficiency (%) Mean particle size (nm)

92.39 100.58

92.51 ± 0.32 105.42 ± 18.30

The mean ± S.D. for three measurements.

Fig. 7. Predicted response surface of drug loading efficiency (Y1) of PLN as a function of drug to lipid weight ratio (X1) and (a) the concentration of Pluronic F68 (X3) with the concentration of Tween 80 (X2) being fixed at 1.92% (w/v) (code form: 1); or (b) Tween 80 (X2) with the concentration of Pluronic F68 (X3) being fixed at 1.92% (w/v) (code form: 1).

Please cite this article in press as: Y. Li et al., Optimization of controlled release nanoparticle formulation of verapamil hydrochloride using artificial neural networks with genetic algorithm and response surface methodology, Eur. J. Pharm. Biopharm. (2015), http://dx.doi.org/10.1016/j.ejpb.2015.04.028

482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506

EJPB 11919

No. of Pages 10, Model 5G

19 May 2015 8

Y. Li et al. / European Journal of Pharmaceutics and Biopharmaceutics xxx (2015) xxx–xxx

Fig. 8. Predicted response surface of mean particle size (Y2) of PLN (a) as a function of weight ratio of drug to lipid (X1) and the concentrations of Pluronic F68 (X3) with a fixed concentration of Tween 80 (X2) at 1.92% (w/v); and (b) as a function of weight ratio of drug to lipid and the concentration of Tween 80 (X2) with a fixed concentration of Pluronic F68 (X3) at 1.92% (w/v).

507

3.7. Effect of Pluronic F68 and Tween 80 concentration

508

The effect of Pluronic F68 and Tween 80 on the drug loading efficiency of PLN (Y1) is illustrated by the three-dimensional surface plots of the drug loading efficiency of PLN (Y1) against the concentration of Pluronic F68 (X3) and weight ratio of drug to lipid (X1) (Fig. 7a) or Tween 80 (X2) and weight ratio of drug to lipid (X1) (Fig. 7b). It is seen that when the concentration of Pluronic F68 increases at a fixed Tween 80 concentration of 1.92% (w/w), the drug loading efficiency remains high and relatively steady until it reaches a high level at high drug to lipid ratios. A similar trend is also observed in case of Tween 80 (Fig. 7b). However, the drug loading efficiency decreases more dramatically with increasing concentration of Tween 80; it dropped to 55.9% (w/w) as compared to 75.8% (w/w) for Pluronic F68 at the same concentration of 3.08%. Fig. 9a shows that the concentration of surfactant plays an important role in the drug loading efficiency of PLN. At a fixed drug to lipid weight ratio (X1) of 17.44% (w/w, coded form: 0.5), when the concentration of Tween 80 or Pluronic F68 exceeds a certain threshold, the drug loading efficiency (Y1) of PLN drops drastically to zero and the formulation becomes flocculated. This phenomenon is also seen in Fig. 9b at high drug to lipid ratios when the concentration of Pluronic F68 remains constant at 2.21% (w/v, coded form: 0.5). The concentration of surfactants also affects the mean particle size of the final product (Y2), as shown in Fig. 10a and b. The degree of the influence depends on the weight ratio of drug to lipid (X1) and the concentration of the second surfactant. When the drug to lipid weight ratio (X1) is given, increasing the concentration of Tween 80 (X2) leads to larger mean particle size (Y2) of PLN. However, the mean particle size and the change pattern at a drug to lipid ratio of 17.4% (coded form: 0.5) differ from those at 32.6% (coded form: 1). The particle size is much larger for PLN at a drug to

509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538

Fig. 9. Predicted response surface of drug loading efficiency (Y1) of PLN (a) as a function of the concentrations of Tween 80 (X2) and Pluronic F68 (X3) with the weight ratio of drug to lipid (X1) being fixed at 17.4% (w/w); and (b) as a function of the concentration of Tween 80 (X2) and the weight ratio of drug to lipid (X1) with the concentration of Pluronic F68 (X3) being fixed at 2.21% (w/v).

lipid ratio of 32.6% than at 17.4%; nonetheless, the sharp increase in the particle size occurs in the high surfactant concentration region for both (Fig. 10a and b). The interaction between the concentration of the two surfactants or the surfactant concentration and the drug to lipid ratio is evident in Fig. 8a and b and Eq. (10). Although the interaction terms are less significant than the individual terms, their values propagate rapidly as the values of X1, X2, and X3 all become big.

539

3.8. Functions of Pluronic 68 and Tween 80

547

Tween 80, a non-ionic surfactant with a hydrophilic–lipophilic balance value (HLB) of 15, was used as a solubilizer in this study. It mainly lowers the surface tension at the oil–water interface to form the oil in water (o/w) emulsion. Pluronic F68 (HLB = 17) is a non-ionic, di-functional block copolymer surfactant composed of poly(ethylene oxide) (PEO) and poly(propylene oxide) (PPO). It was used as a co-emulsifier and a suspension stabilizer in the present study. It is mainly adsorbed onto the surface of PLN particles through the hydrophobic interaction, preventing them from aggregation via steric repulsion due to its high percentage of hydrophilic groups of ethylene oxide (EO). The number of EO units for a single Tween 80 molecule is 20, while that for Pluronic F68 is 75. As compared to Pluronic F68, Tween 80 is a smaller molecule with a molecular weight (Mn) of 1310 Da and a shorter hydrophobic chain. Pluronic F68 is a tri-block co-polymer with a molecular weight of 8500 Da. In addition to the longer PEO chains, Pluronic F68 has a long hydrophobic chain containing 30 propylene oxide (PO) units. This difference in

548

Please cite this article in press as: Y. Li et al., Optimization of controlled release nanoparticle formulation of verapamil hydrochloride using artificial neural networks with genetic algorithm and response surface methodology, Eur. J. Pharm. Biopharm. (2015), http://dx.doi.org/10.1016/j.ejpb.2015.04.028

540 541 542 543 544 545 546

549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565

EJPB 11919

No. of Pages 10, Model 5G

19 May 2015 Y. Li et al. / European Journal of Pharmaceutics and Biopharmaceutics xxx (2015) xxx–xxx

Fig. 10. Predicted response surface of mean particle size (Y2) of PLN as a function of the concentration of Tween 80 (X2) and the concentration of Pluronic F68 (X3). (a) The weight ratio of drug to lipid (X1) was fixed at 17.4% (w/w), and (b) the weight ratio of drug to lipid (X1) was fixed at 32.6% (w/w).

566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592

593 595

the molecular structure between Pluronic F68 and Tween 80 imparts them different roles in the preparation and formulation stability of the PLN. The major function of Tween 80 in the preparation of PLN is as a solubilizer, while Pluronic F68 as a stabilizer. Although both Tween 80 and Pluronic F68, with HLB > 10, are suitable for stabilizing oil-in-water (O/W) emulsions, Tween 80 (HLB 15) can make more stable emulsions than Pluronic F68 (HLB17) [38]. This may explain why drug loading efficiency dropped dramatically at high Tween 80 concentrations when the Pluronic F68 concentration was fixed (Figs. 7 and 9). Tween 80, at high concentrations, could solubilize both melted lipid (oil) and drug–polymer complex, separate them from merging, and thus reduce drug loading efficiency. When the concentrations of Tween 80 (X2) or Pluronic F68 (X3) are fixed, the drug loading efficiency (Y1) decreased sharply (Figs. 7 and 9) while particle size increased at high weight ratio of drug to lipid (X1) (Fig. 10) might be attributed to the limitation of the solubilization capacity of the surfactants at this composition and the change in the HLB value of the oil phase with changing drug to lipid ratio. As the ionic molar ratio of drug to polymer was fixed at 1:1, an increase in the drug to lipid ratio led to a higher concentration of the drug and the hydrophilic polymer. As a consequence, the HLB of the oil phase increases and thus a different HLB value of surfactants is required to form stable emulsions [38]. When two surfactants are applied in the preparation, the combined HLB value can be estimated from the weighted average as follows:

p  HLB1 þ ð1  pÞ  HLB2 ¼ HLBcombined

ð11Þ

9

where p is the proportion of surfactant 1; HLB1 and HLB2 are respectively the HLB value of surfactant 1 and surfactant 2. As the HLB value of Pluronic F68 (HLB = 17) is quite different from that of Tween 80 (HLB = 15), the HLB value of their mixture is sensitive to their ratio. The correlation between the surfactant concentrations and the drug loading efficiency and particle size has been established from ANN models. With these correlations and understanding of the functions of the surfactants, one may develop more formulations for desirable PLN properties.

596

4. Conclusion

605

We successfully optimized the formulation of PLN of VRP using the spherical central composite design and the ANN methodology. The resultant PLN exhibited high drug loading efficiency (92%) and small mean particle size (100 nm) which is desirable for lymphatic transport via oral administration. We found that ANN fit the experimental data better than RSM in this typical study attributable to ANN’ capacity to accommodate more complex and non-linear functional relationships. The combined application of factorial design, ANN and continuous GA in modeling and multi-objective optimization can facilitate the development of PLN with desired properties.

606

Acknowledgments

617

This work was partially supported by the Canadian Institutes of Health Research. The authors would like to thank Dr. T.V. Chalikian for kind permission on use of ITC and valuable discussion, and Mr. B. Calvieri for TEM imaging. In addition, financial support from the Natural Sciences and Engineering Research Council (NSERC) postgraduate scholarship (doctoral level) and University of Toronto Top-up award to Y. Li is also gratefully acknowledged.

618

References

625

[1] M.J. Abdekhodaie, X.Y. Wu, Drug release from ion-exchange microspheres: mathematical modeling and experimental verification, Biomaterials 29 (2008) 1654–1663. [2] H. Akaike, New look at statistical-model identification, IEEE Trans. Autom. Control 19 (1974) 716–723. [3] N.A. Armstrong, Pharmaceutical Experimental Design and Interpretation, CRC Press, London, 2006. pp. 83–129. [4] P. Barmpalexis, F.I. Kanaze, K. Kachrimanis, E. Georgarakis, Artificial neural networks in the optimization of a nimodipine controlled release tablet formulation, Eur. J. Pharm. Biopharm. 74 (2) (2010) 316–323. [5] J. Basile, The role of existing and newer calcium channel blockers in the treatment of hypertension, J. Clin. Hyperten. 6 (2004) 621–629. [6] T. Bekerman, J. Golenser, A. Domb, Cyclosporin nanoparticulate lipospheres for oral administration, J. Pharm. Sci. 93 (2004) 1264–1270. [7] D. Belpomme, S. Gauthier, E. Pujade-Lauraine, T. Facchini, M.J. Goudier, I. Krakowski, G. Netter-Pinon, M. Frenay, C. Gousset, F.N. Marie, M. Benmiloud, F. Sturtz, Verapamil increases the survival of patients with anthracyclineresistant metastatic breast carcinoma, Ann. Oncol. 11 (2000) 1471–1476. [8] P.M. Bummer, Physical chemical considerations of lipid-based oral drug delivery–solid lipid nanoparticles, Crit. Rev. Ther. Drug Carrier Syst. 21 (2004) 1–20. [9] S.M. Caliph, W.N. Charman, C.J. Porter, Effect of short-, medium-, and longchain fatty acid-based vehicles on the absolute oral bioavailability and intestinal lymphatic transport of halofantrine and assessment of mass balance in lymph-cannulated and non-cannulated rats, J. Pharm. Sci. 89 (2000) 1073–1084. [10] W.C. Carpenter, M.E. Hoffman, Understanding Neural Network Approximations and Polynomial Approximations Helps Neural Network Performance, AI Expert, 1995, pp. 31–33. [11] K. Deb, A. Anand, D. Joshi, A computationally efficient evolutionary algorithm for real-parameter optimization, Evol. Comput. 10 (2002) 371–395. [12] M.P. Desai, V. Labhasetwar, G.L. Amidon, R.J. Levy, Gastrointestinal uptake of biodegradable microparticles: effect of particle size, Pharm. Res. 13 (1996) 1838–1845. [13] D.B. Fogel, An information criterion for optimal neural network selection, IEEE Trans. Neur. Net. 2 (1991) 490–497. [14] D.E. Fonner, J.R. Buck, G.S. Banker, Mathematical optimization techniques in drug product design and process analysis, J. Pharm. Sci. 59 (1970) 1587–1596.

626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663

Please cite this article in press as: Y. Li et al., Optimization of controlled release nanoparticle formulation of verapamil hydrochloride using artificial neural networks with genetic algorithm and response surface methodology, Eur. J. Pharm. Biopharm. (2015), http://dx.doi.org/10.1016/j.ejpb.2015.04.028

597 598 599 600 601 602 603 604

607 608 609 610 611 612 613 614 615 616

619 620 621 622 623 624

EJPB 11919

No. of Pages 10, Model 5G

19 May 2015 10 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717

Y. Li et al. / European Journal of Pharmaceutics and Biopharmaceutics xxx (2015) xxx–xxx

[15] A. Ghaffari, H. Abdollahi, M.R. Khoshayand, I.S. Bozchalooi, A. Dadgar, M. Rafiee-Tehrani, Performance comparison of neural network training algorithms in modeling of bimodal drug release, Int. J. Pharm. 327 (2006) 126–138. [16] E. Grossman, F.H. Messerli, Calcium antagonists, Prog. Cardiovasc. Dis. 47 (2004) 34–57. [17] S.R. Hamann, R.A. Blouin, R.G. McAllister Jr., Clinical pharmacokinetics of verapamil, Clin. Pharmacokinet. 9 (1984) 26–41. [18] A.S. Hussain, X.Q. Yu, R.D. Johnson, Application of neural computing in pharmaceutical product development, Pharm. Res. 8 (1991) 1248–1252. [19] S. Ibric´, J. Djuriš, J. Parojcˇic´, Z. Djuric´, Artificial neural networks in evaluation and optimization of modified release solid dosage forms, Pharmaceutics 4 (2012) 531–550. [20] K. Kachrimanis, V. Karamyan, S. Malamataris, Artificial neural networks (ANNs) and modeling of powder flow, Int. J. Pharm. 250 (2003) 13–23. [21] A.A. Khan, J. Mudassir, N. Mohtar, Y. Darwis, Advanced drug delivery to the lymphatic system: lipid-based nanoformulations, Int. J. Nanomed. 8 (2013) 2733–2744. [22] A.I. Khuri, M. Conlon, Simultaneous optimization of multiple responses represented by polynomial regression functions, Technometrics 23 (1981) 363–375. [23] A.I. Khuri, J.A. Cornell, Response Surface. Design and Analysis, Marcel Dekker, New York, 1978. pp. 116–140. [24] Y. Li, A.M. Rauth, X.Y. Wu, Prediction of kinetics of doxorubicin release from sulfopropyl dextran ion-exchange microspheres using artificial neural networks, Eur. J. Pharm. Sci. 24 (2005) 401–410. [25] Y. Li, N. Taulier, A.M. Rauth, X.Y. Wu, Screening of lipid carriers and characterization of drug–polymer–lipid interactions for the rational design of polymer–lipid hybrid nanoparticles (PLN), Pharm. Res. 23 (2006) 1877– 1887. [26] Y. Li, H.L. Wong, A.J. Shuhendler, A.M. Rauth, X.Y. Wu, Molecular interactions, internal structure and drug release kinetics of rationally developed polymer– lipid hybrid nanoparticles, J. Control. Rel. 128 (2008) 60–70. [27] Z. Liu, X.Y. Wu, R. Bendayan, In vitro investigation of ionic polysaccharide microspheres for simultaneous delivery of chemosensitizer and antineoplastic agent to multidrug-resistant cells, J. Pharm. Sci. 88 (1999) 412–418. [28] Z. Liu, R. Cheung, X.Y. Wu, J.R. Ballinger, R. Bendayan, A.M. Rauth, A study of doxorubicin loading onto and release from sulfopropyl dextran ion-exchange microspheres, J. Control. Rel. 77 (2001) 213–224. [29] Z. Liu, J. Ballinger, M. Rauth, R. Bendayan, X.Y. Wu, Delivery of an anticancer drug and a chemosensitizer to murine breast sarcoma by intratumoral injection of sulfopropyl dextran microspheres, J. Pharm. Pharmacol. 55 (2003) 1063–1073. [30] K. Manjunath, J.S. Reddy, V. Venkateswarlu, Solid lipid nanoparticles as drug delivery systems, Methods Find Exp. Clin. Pharmacol. 27 (2005) 127–144. [31] D. McTavish, E.M. Sorkin, Verapamil. An updated review of its pharmacodynamic and pharmacokinetic properties, and therapeutic use in hypertension, Drugs 38 (1989) 19–76. [32] W. Mehnert, K. Mader, Solid lipid nanoparticles: production, characterization and applications, Adv. Drug Deliv. Rev. 47 (2001) 165–196. [33] Y. Miyamoto, Y. Obata, M. Miyajima, M. Matsui, H. Sato, K. Takayama, T. Nagai, An application of the computer optimization technique to wet granulation process involving explosive growth of particles, Int. J. Pharm. 149 (1997) 25– 36.

[34] R.H. Muller, K. Mader, S. Gohla, Solid lipid nanoparticles (SLN) for controlled drug delivery – a review of the state of the art, Eur. J. Pharm. Biopharm. 50 (2000) 161–177. [35] Y. Nishioka, H. Yoshino, Lymphatic targeting with nanoparticulate system, Adv. Drug Deliv. Rev. 47 (2001) 55–64. [36] C.M. O’Driscoll, Lipid-based formulations for intestinal lymphatic delivery, Eur. J. Pharm. Sci. 15 (2002) 405–415. [37] J. Pardeike, A. Hommoss, R.H. Muller, Lipid nanoparticles (SLN, NLC) in cosmetic and pharmaceutical dermal products, Int. J. Pharm. 366 (2009) 170– 184. [38] R. Phan-Tan-Luu, D. Mathieu, Experimental Design in Emulsion and Suspension Formulations: Theoretical Aspects, Marcel Dekker Inc., New York, 2000. pp. 465–534. [39] C.J. Porter, Drug delivery to the lymphatic system, Crit. Rev. Ther. Drug Carrier Syst. 14 (1997) 333–393. [40] D.E. Rumelhart, G.E. Hinton, R.J. Williams, Learning representations by backpropagating errors, Nature 323 (1986) 533–536. [41] A. Sabir, B. Evans, S. Jain, Formulation and process optimization to eliminate picking from market image tablets, Int. J. Pharm. 215 (2001) 123–135. [42] B. Singh, R. Kumar, N. Ahuja, Optimizing drug delivery systems using systematic ‘‘design of experiments’’. Part I: fundamental aspects, Crit. Rev. Ther. Drug Carrier Syst. 22 (2005) 27–105. [43] A.J. Shuhendler, R.Y. Cheung, J. Manias, A. Connor, A.M. Rauth, X.Y. Wu, A novel doxorubicin-mitomycin C co-encapsulated nanoparticle formulation exhibits anti-cancer synergy in multidrug resistant human breast cancer cells, Breast Can Res. Treat. 119 (2010) 255–269. [44] B.H. Sumida, A.I. Houston, J.M. McNamara, W.D. Hamilton, Genetic algorithms and evolution, J. Theor. Biol. 147 (1990) 59–84. [45] Y. Sun, Y. Peng, Y. Chen, A.J. Shukla, Application of artificial neural networks in the design of controlled release drug delivery systems, Adv. Drug Deliv. Rev. 55 (2003) 1201–1215. [46] M.A. Swartz, The physiology of the lymphatic system, Adv. Drug Deliv. Rev. 50 (2001) 3–20. [47] K. Takayama, M. Fujikawa, T. Nagai, Artificial neural network as a novel method to optimize pharmaceutical formulations, Pharm. Res. 16 (1999) 1–6. [48] K. Takayama, M. Fujikawa, Y. Obata, M. Morishita, Neural network based optimization of drug formulations, Adv. Drug Deliv. Rev. 55 (2003) 1217–1231. [49] K. Takayama, A. Morva, M. Fujikawa, Y. Hattori, Y. Obata, T. Nagai, Formula optimization of theophylline controlled-release tablet based on artificial neural networks, J. Control. Rel. 68 (2000) 175–186. [50] S. Vaithiyalingam, M.A. Khan, Optimization and characterization of controlled release multi-particulate beads formulated with a customized cellulose acetate butyrate dispersion, Int. J. Pharm. 234 (2002) 179–193. [51] H.L. Wong, R. Bendayan, A.M. Rauth, X.Y. Wu, Development of solid lipid nanoparticles containing lonically complexed chemotherapeutic drugs and chemosensitizers, J. Pharm. Sci. 93 (2004) 1993–2008. [52] G.P. Zara, A. Bargoni, R. Cavalli, A. Fundaro, D. Vighetto, M.R. Gasco, Pharmacokinetics and tissue distribution of idarubicin-loaded solid lipid nanoparticles after duodenal administration to rats, J. Pharm. Sci. 91 (2002) 1324–1333. [53] A.S. Zidan, O.A. Sammour, M.A. Hammad, N.A. Megrab, M.J. Habib, M.A. Khan, Quality by design: understanding the formulation variables of a cyclosporine A self-nanoemulsified drug delivery systems by Box–Behnken design and desirability function, Int. J. Pharm. 332 (2007) 55–63.

718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772

Please cite this article in press as: Y. Li et al., Optimization of controlled release nanoparticle formulation of verapamil hydrochloride using artificial neural networks with genetic algorithm and response surface methodology, Eur. J. Pharm. Biopharm. (2015), http://dx.doi.org/10.1016/j.ejpb.2015.04.028

Optimization of controlled release nanoparticle formulation of verapamil hydrochloride using artificial neural networks with genetic algorithm and response surface methodology.

This study was performed to optimize the formulation of polymer-lipid hybrid nanoparticles (PLN) for the delivery of an ionic water-soluble drug, vera...
3MB Sizes 0 Downloads 7 Views