Sie sind auf Seite 1von 9

Journal of Membrane Science 564 (2018) 682–690

Contents lists available at ScienceDirect

Journal of Membrane Science


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

Continuous production of drug nanocrystals by porous hollow fiber-based T


anti-solvent crystallization
Xinyi Zhoua,1, Xuan Zhua,1, Bing Wangb, Jingchao Lic, Qiuhong Liua, Xuemin Gaoa,
⁎ ⁎
Kamalesh K. Sirkard, , Dengyue Chena,
a
School of Pharmaceutical Sciences, Xiamen University, Xiamen, Fujian 361102, China
b
Key Laboratory of Urban Pollutant Conversion, Institute of Urban Environment, Chinese Academy of Sciences, Xiamen, Fujian 361021, China
c
College of Chemistry and Chemical Engineering, Xiamen University, Xiamen, Fujian 361005, China
d
New Jersey Institute of Technology, Otto York Department of Chemical, Biological and Pharmaceutical Engineering, Newark, NJ, USA

A R T I C LE I N FO A B S T R A C T

Keywords: Nanotechnology is being utilized to develop advanced concepts in drug delivery systems including production of
Drug nanocrystals nanodrugs such as, nanocrystals and nanosuspensions. Continuous crystallization methods are of increasing
Griseofulvin interest and porous hollow fiber membrane (HFM) based modules have been recently utilized as an anti-solvent
Anti-solvent crystallization crystallizer. The porous hollow fiber anti-solvent crystallizer (PHFAC) based method has been modified here to
Porous hollow fiber membrane
produce continuously nanocrystals of Griseofulvin (GF). In the PHFAC device, deionized water introduced from
the HFM bore into the shell side is used as the anti-solvent; acetone containing dissolved GF is introduced into
the shell side of the HFM module as the drug solution. Water, the anti-solvent, mixed vigorously with the drug
solution leading to drug crystallization and the drug crystals were separated by vacuum filtration and freeze-
dried. The experimental conditions were varied to control the particle size, size distribution and the appearance
of the nanoparticles. The properties of the drug nanocrystals were characterized via Transmission Electron
Microscopy (TEM), Scanning Electron Microscope (SEM), Dynamic Light Scattering (DLS), Raman Spectroscopy,
Energy Dispersive X-rays Spectroscopy (EDX), Differential Scanning Calorimetry (DSC), FT-IR Spectrometer, X-ray
Diffraction (XRD); drug dissolution tests were also implemented. Drug nanocrystals as small as 86.4 nm were
produced under modest pressure and temperature conditions in a controllable and continuous manner.

1. Introduction improved with drug nanoparticles for the purpose of protecting pep-
tides, vaccines and other drugs from clearance in the digestive tract [1].
Due to their small size, nanoparticles possess unique capabilities, Due to rapid development of nanoscience and nanotechnology, synth-
such as volume effect, quantum size effect, surface effect, macroscopic esis of drug nanocrystals has been widely applied to drug delivery
quantum tunneling effect and so on and are therefore widely used in system (DDS) [2]. Nanonization of the drug is a desirable method to
various fields. In pharmaceutical area, drug nanoparticles have at- improve bioavailability and efficacy in order to enhance their ther-
tracted increasing interest from researchers and industry for their dis- apeutic effect [3]. A number of size reduction techniques were devel-
tinct advantages. For instance, they can be passively targeted on liver, oped or optimized to produce stable drug nanocrystals [4].
spleen, bone marrow and other organs and tissues to cure the disease. Current approaches in production of drug nanoparticles involve
Nanoparticle drugs sized under 100 nm have biological permeability either top-down or bottom-up methods [5]. For top-down methods,
through tumor vascular walls and hence are able to achieve a curative drug particles are broken down to lower size particles by methods, such
effect. Moreover, nanoparticles can increase the efficacy of antibiotics as pearl milling, high pressure homogenization etc.; these have greater
and enhance the potency of antifungal and antiviral agents in the utility for drugs having a high degree of crystallinity [6–8]. For bottom-
treatment of intracellular bacterial infections. up methods of direct particle formation, a classical precipitation process
For drugs with poor water solubility and dissolution rate, the sta- of ‘via humida paratum’ (VHP) is developed where drug dissolved in a
bility and bioavailability of oral preparations can be drastically solvent is precipitated by mixing with a non-solvent. This method


Corresponding authors.
E-mail addresses: sirkar@njit.edu (K.K. Sirkar), dchen@xmu.edu.cn (D. Chen).
1
X. Zhou and Dr. X. Zhu contributed equally to this work and should be considered as co-first authors.

https://doi.org/10.1016/j.memsci.2018.07.082
Received 10 April 2018; Received in revised form 12 July 2018; Accepted 27 July 2018
Available online 29 July 2018
0376-7388/ © 2018 Elsevier B.V. All rights reserved.
X. Zhou et al. Journal of Membrane Science 564 (2018) 682–690

succeeds in producing crystalline drug nanoparticles while requiring hydrodynamic optimization of the module [28].
strict control of the process [9]. One of the common methods, super- To downsize the production of micron size GF drug crystals into
critical fluid (SCF) technology, solubilizes the solute in the supercritical nanocrystals of GF, we have redesigned the membrane module and
fluid first and subsequently precipitates it via supersaturation through investigated the hydrodynamic and crystallization conditions to
rapid expansion to obtain virtually contaminant-free particles [10,11]. achieve controlled and continuous synthesis of drug nanocrystals of GF
In another technique, the spray freeze drying (SFD), the method in- at ambient pressure and temperature. The GF nanocrystals produced
volves the steps of droplet generation, freezing and sublimation drying have been characterized extensively and the enhancement of their
[2,12,13]. However, high expenditure, intensive labor, demand of dissolution characteristics was also explored.
material and risks of dust explosion are disadvantages of these techni-
ques [14]. Confined liquid impinging jets (CLIJ) has been used to
produce particles in the submicron to nano range where precipitation 2. Experimental methods
takes place when a jet of drug solution impinges a jet of anti-solvent
from two opposing nozzles [15]. But secondary crystallization and 2.1. Materials
further growth may be induced by incomplete depletion of solute
during nucleation and crystal growth. A precipitation technique named Griseofulvin (purity 99.2%) was obtained from
as in situ micronization has been developed to produce the drug na- Yuanchenggongchuang Technology Co. Ltd (Wuhan, Hubei, China).
noparticles in a relatively simple and modest way [16]; however, it may Acetone (analytical purity) was purchased from Xilong Chemical Co.
result in larger particles and poor control of crystal size. The frequently Ltd (Guangzhou, Guangdong, China). Deionized water was produced by
utilized method of rapid expansion of supercritical solutions (RESS) multi-function ultrapure water system (Unique-R20, Ruisijie Water-
enables particle formation due to rapid supersaturation by applying purification Technology Co. Ltd, Xiamen, Fujian, China). All reagents
rapid phase change of supercritical fluids. Supercritical Anti-Solvent were of analytical grade and used as received.
(SAS) method [17–19] induces precipitation/crystallization of solutes
by saturating the polar liquid solvent with carbon dioxide under su-
percritical conditions [11]; such a process would require critical ex- 2.2. Apparatus and procedures
perimental conditions such as the generation of high pressure condition
for supercritical CO2. Moreover, this method is a batch process with low The experimental apparatus schematically shown in Fig. 1(b) was
output levels and questionable stability of the drug quality, which are designed for anti-solvent induced precipitation. The PHFAC system,
against the requirements for modern pharmaceutical production. Nano fabricated to perform the experiment, consisted of three sections: inlet
spray drying invented by Büchi® employs a vibration mesh spray system; crystallization system; vacuum filtration system. The inlet
technology for fine droplet generation and is ideal for heat-sensitive system had two parts, with the first part generating the drug solution
biopharmaceutical products [20]. stream and the other part generating the anti-solvent stream. The so-
Meanwhile, porous hollow fiber membrane-based techniques were lution containing GF was prepared by stirring a suspension of the as-
being developed to implement continuous phase-change processes. A received GF crystals in acetone in a conical flask on a magnetic stirring
continuously flowing solution of BaCl2 on the shell-side permeated into apparatus until the drug was fully dissolved in the solvent. Upon
a solution of K2SO4 flowing in the tube-side of a hollow fiber membrane complete dissolution, the GF solution was passed into the shell side of
device to achieve precipitation of BaSO4 particles via chemical reaction the crystallization system by a peristaltic pump (YZ1515x, Shenchen
[21–23]. This configuration is prone to the possibility of plugging of the Pump Industry Co. Ltd, Baoding, Hebei, China). Meanwhile, water, the
hollow fiber bore unless special designs are made and precautions are anti-solvent, was pumped into the lumen side of the PHFAC system by
taken. Instead of a chemical reaction-based precipitation, Zarkadas another peristaltic pump at a fixed flow rate at a slightly earlier time. In
[24], Zarkadas and Sirkar [25] utilized hollow fiber devices to imple- the PHFAC system, the left end of the tube side (shown in Fig. 1(b)) was
ment anti-solvent crystallization. They found that it was more useful to blocked by epoxy resin. This prevented water from flowing out from the
implement crystallization in the shell-side of the module so that chance tube side end and created the pressure difference between the tube and
of flow-passage blockage was much less due to many directional free- the shell side to push the drug solution out to the shell side bulk from
doms of flow. The specific configuration was adopted in which drug- the outside of the hollow fiber membrane surface. The PHFAC module
containing feed solution flowing in the hollow fiber bore permeated shell side was formed by a fluorinated ethylene propylene (FEP) tubing
into the shell side where the anti-solvent was flowing; thus crystal- with an outer diameter (OD) of 19 mm and an inner diameter (ID) of
lization took place in the shell-side. 17 mm. The effective tube length was 26 cm, which was also the length
Recently, Chen et al. [26] reported a convenient anti-solvent crys- of each of the 14 polyvinylidene fluoride (PVDF) porous hollow fiber
tallization technique to synthesize continuously polymer-coated drug membranes (pore size 0.1–0.15 µm; porosity, 0.75) having an OD of
particles using a porous hollow fiber membrane device. Interestingly, 2.5 mm and ID of 1.8 mm. As shown in Fig. 1(b), PHFAC module was
the first part of this study [26] described PHFAC-based continuous placed at an angle of 15° to the horizontal to maximize the contacting
crystallization of micron size GF crystals depending on various condi- area for the drug solution and anti-solvent in the shell side along the
tions that produced GF crystals with the median size from 1.61 µm to fiber length direction.
11.83 µm. They had used an alternate arrangement for shell-side crys- The perspective view of a single porous hollow fiber membrane
tallization: the anti-solvent water was introduced from the hollow fiber shown in Fig. 1(a) illustrates the precipitation process in the drug so-
bore into the shell-side having the flowing GF-in-acetone solution lution flowing on the shell-side induced by the anti-solvent. In the shell
(Fig. 1(a)). The introduction of the anti-solvent water through tiny side, the drug solution contacted uniformly with the anti-solvent which
membrane pores into the acetone-based drug solution in the shell-side penetrated out from the pores (0.1–0.15 µm) in the hollow fiber walls.
of the HFM module ensured a high-mixing zone in the microenviron- The drug crystals along with the excess solution flowed out to the outlet
ment, which further provided a suitable condition for the crystallization of the device continuously and subsequently to the vacuum filtration
of small drug particles. system (1000 mL, Tianjin Jinteng Experimental Equipment Co. Ltd,
This PHFAC technique and the HFM device were also used suc- Tianjin, China); this system allowed collection of the particles through
cessfully to produce polymer-coated submicron as well as nanoparticles 0.1 µm pore-sized membrane filters (VVLP04700, Merck Millipore Ltd,
of silica in a controllable way [27]. These successes suggested the County Cork, Ireland), which were freeze-dried prior to various char-
possibility of production of even smaller drug nanocrystals by con- acterization steps.
trolling the nucleation and growth condition of crystals as well as the

683
X. Zhou et al. Journal of Membrane Science 564 (2018) 682–690

Fig. 1. (a) Perspective view of single porous hollow fiber in the tube side. (b) Schematic of experimental installation of the porous hollow fiber anti-solvent
crystallizer (PHFAC).

2.3. Experimental operating conditions 2.4.5. Energy dispersive X-rays spectroscopy


EDX spectroscopy analysis (S-4800, Hitachi, Japan) was applied to
Since the feed drug solution exerted a significant influence on a confirm the composition of the elements for as-received drug powder
variety of results, a number of items were varied. The ratios of the flow and the crystallized drug nanocrystals.
rates of the anti-solvent and the drug solution contacting in the shell-
side mixing zone played a significant role in the procedure. The drug
2.4.6. XRD
concentration in the solution was also a critical parameter that had
X-ray diffraction (XRD) measurements were conducted on an Ultima
impact on the formation and size of the drug particles. An amount of
IV X-ray diffractometer (Rigaku, Matsubara-cho, Tokyo, Japan) with
0.8 g of as-received GF was placed in a flask containing 22 mL of the
a Cu-Kα radiation source (λ = 0.15418 nm) at room temperature.
solvent mixture. The solvent mixture was prepared with acetone and
The voltage and current were set at 40 kV and 30 mA. Placed in a
deionized water in the proportion of 10:1. Under continuous stirring on
glass sample holder, samples were scanned over the 2θ range of
the magnetic stirring apparatus, the drug solution changed from a
5–40° at a scanning rate of 15°/min with a step size of 0.02°.
suspension to a solution. In order to explore the formation of the finest
droplets, the flow rates of the anti-solvent and the drug solution were
set at different ratios and the drug concentrations were designed in a
2.4.7. DSC
gradient fashion.
Differential scanning calorimetric (DSC) analysis was performed
using a Synchronous TG-DSC Thermal Analyzer (STA 449 F5
2.4. Particle characterization Jupiter, Netzsch, Bavaria, Germany). The sample typically weighing
about 10 mg was placed in an alumina crucible and heated from
2.4.1. TEM 75 °C to 250 °C at a heating rate of 10 °C/min under nitrogen at-
The morphology and appearance of GF particles was characterized mosphere with a constant flow rate of 20 mL/min. An empty cru-
by transmission electron microscopy (TEM, Tecnai G, Spirit, FEI, Hong cible was utilized as a reference.
Kong). After sonication, the samples were well dispersed and subse-
quently deposited on carbon coated copper grids for characterization.
2.4.8. FT-IR
A FT-IR spectrometer (Bruker, Germany) was used to test whether
2.4.2. SEM the chemical structure of the obtained drug crystals had changed. The
The surface morphology of drug nanocrystals was observed using a mixture of the samples and potassium bromide were compressed into
scanning electron microscope (SEM, Supra 55, Zeiss, Oberkochen, thin pellets and scanned over the range of 400–4000 cm−1.
Baden-wurttemberg, Germany). The drug nanocrystals were first dis-
persed in deionized water; they were subsequently dropped onto the
2.4.9. In vitro dissolution testing
surface of a conducting resin and dried in air. Finally, the samples were
Dissolution tests were conducted in triplicate by the paddle method
sputtered with gold for 30 s under vacuum for observation.
using a RC806 dissolution tester (Tianjin Tianda Tianfa Technology Co.,
Ltd., Tianjin, China) to investigate whether there could be distinction
2.4.3. DLS between the dissolution rate of drug nanocrystals and as-received drug
The particle size distribution was measured by a dynamic light powder. The drug nanocrystals as well as the as-received GF powder
scattering (DLS) instrument (Nano-ZS, Malvern Instruments Ltd, were carefully weighed and prepared for tests both having a weight of
Worcestershire, UK). The samples were prepared by diluting in deio- 15 mg. Glass vessels with a maximum nominal volume of 1 L were used
nized water and analyzed at 25 °C, with a dispersant refractive index of with 500 mL pH 6.8 phosphate buffer (37 ± 0.5 °C) as the dissolution
1.33. medium; the paddles were rotated at a speed of 100 rpm. At 5, 10, 20,
30, 40, 50, 70, 90, 110, 130 min time intervals, 5 mL samples were
taken for the UV test and replaced with another 5 mL fresh medium.
2.4.4. Raman spectroscopy Passed through a 0.45 µm filter (Tianjin Jinteng Experimental
The molecular structure of the GF particles was determined with a Equipment Co. Ltd, Tianjin, China), the absorbance value of samples
confocal Raman system (Xplora, Horiba, Japan). The intensity of laser was determined by a UV-1750 ultraviolet spectrophotometer
power was set at 10 mW, with the incident laser wavelength at 780 nm. (Shimadzu Suzhou Instruments Mfg. Co., Ltd, Japan) at a wavelength of
Raman spectra were recorded over the range of 0–3500 cm−1. 292 nm.

684
X. Zhou et al. Journal of Membrane Science 564 (2018) 682–690

Fig. 2. TEM images of (a) synthesized drug nanocrystals by PHFAC under the condition of flow rate at 6.8 mL/min in tube side and 13.6 mL/min in shell side with
drug concentration at 0.040 g/mL, (b) as-received pure GF powder.

3. Results and discussion chlorine, which demonstrated that the successfully prepared particles
were nano-sized GF crystals.
3.1. Continuous synthesis of drug nanoparticles
3.1.2. Particle size characterization
Griseofulvin crystals were first precipitated in the PHFAC device by As shown in Fig. 5, the peak of particle size distribution (PSD) of the
anti-solvent induced crystallization in the shell side. During this pro- obtained nanocrystals was at 295.6 nm by DLS analysis (we will later
cess, DI water in the tube side penetrated through numerous tiny pores show that it can go as low as 86.4 nm), which was considerably smaller
(0.1–0.15 µm) on every hollow fiber wall; this caused rapid solubility than that of the as-received drug GF powder whose peak was at
reduction of the drug in every tiny mixing zone on the shell side which 3403.0 nm. The smaller particle size and PSD for synthesized drug
rapidly led to the formation of drug nanocrystals. particles could be due to the following reasons. Firstly, the well dis-
tributed submicron-sized pores (0.1–0.15 µm) on the hollow fiber
3.1.1. TEM, SEM and EDX membrane walls (with 75% porosity) created numerous tiny and uni-
TEM results of the produced drug nanocrystals are shown in form microenvironments in the shell-side when the penetrated anti-
Fig. 2(a). These drug crystals were produced using 4.33% (by wt.) so- solvent contacted the drug solution. The small and numerous mem-
lution of GF in acetone for the following flow rates of the acetone so- brane pores generated tiny mixing zones that provided the required
lution and the anti-solvent water: 13.6 mL/min and 6.8 mL/min. The conditions for nucleation and growth of drug nanocrystals. Moreover,
TEM image of Fig. 2(a) shows that relatively spherical drug particles the short effective length of the PHFAC and continuous flow in the
were synthesized successfully with a size of around 200 nm. The TEM device could decrease the residence time that further helped shorten the
image of Fig. 2(b) shows that the original particle size of the as-received crystal growth in the crystallization process, eventually leading to a
pure GF powder was about 3–4 µm, which is more than 10 times the small crystal size of the drug.
size of the synthesized drug nanoparticles obtained continuously.
SEM images of Fig. 3(a) and 3(b) identify the uniform distribution of 3.1.3. Composition analysis
drug particles with some crystalline structure. Besides, the morphology The results of Raman spectroscopy in Fig. 6 show that the main
of the GF powder shown in Fig. 3(c) seemed to have some agglom- peaks of the gained drug nanocrystals were at 2950.62, 1622.55,
eration. SEM-EDX spectroscopy analysis in Fig. 4(a) and 4(b) further 1343.43, 648.689 and 375.38 cm−1, respectively which are in agree-
identify the composition elements for the as-received sample and the ment with the characteristic peaks of the as-received crude drug. These
crystallized drug product, respectively. In both images, the composi- results indicated that the synthesized nanocrystals presented no trans-
tions of particles were identical vis-a-vis carbon, hydrogen, oxygen and formation in the peak position over the shift range from 0 to

Fig. 3. SEM images of (a) synthesized drug nanocrystals by PHFAC under the condition of flow rate at 6.8 mL/min in both tube side and shell side with drug
concentration at 0.040 g/mL on a small scale, (b) synthesized drug nanocrystals by PHFAC under the condition of flow rate at 6.8 mL/min in both tube side and shell
side with drug concentration at 0.040 g/mL on a large scale, and (c) as-received pure GF powder.

685
X. Zhou et al. Journal of Membrane Science 564 (2018) 682–690

Fig. 4. EDX spectroscopy analysis results of (a) obtained GF nanocrystals under the condition of flow rate at 6.8 mL/min in both tube side and shell side with drug
concentration at 0.040 g/mL, (b) as-received pure GF powder.

Fig. 5. Particle size distribution (PSD) of synthesized drug nanocrystals by PHFAC under the condition of flow rate at 6.8 mL/min in both tube side and shell side with
drug concentration at 0.040 g/mL and as-received pure GF powder.

3500 cm−1; the nanocrystals precipitated from the drug solution re- 3.1.4. Crystallinity analysis
tained the original structures of the molecular groups. XRD patterns for drug nanocrystals along with as-received drug
FT-IR analysis was utilized to examine the chemical structure of the powder are shown in Fig. 8. The sharp peaks in the patterns confirmed
obtained crystals as well. As shown in Fig. 7, the two characteristic the crystalline forms of both as-received drug powder and drug nano-
peaks at 1665 and 1707 cm−1 were derived from the stretching vibra- crystals. It is obvious that the crystalline peaks of drug nanocrystals
tion of the carbonyl groups (vC˭O). Meanwhile, the peaks at 1463, 1502, (86.4 nm) were in agreement with those of as-received drug powder.
1584, 1614 cm−1 were derived from the aromatic C˭C stretching vi- Meanwhile, the reduced intensity of crystalline peaks for drug nano-
bration (vC˭C) and C-H stretching vibration (vC-H) contributed to the crystals may be attributed to the reduction in the particle size.
peaks close to 2995 cm−1. It indicated that the chemical structures of DSC thermal curves for as-received drug powder and obtained drug
the gained nanocrystals were stable as all the characteristic peaks were nanocrystals are presented in Fig. 9. As can be seen, a melting tem-
relatively identical to those of the as-received powder. perature of drug nanocrystals (86.4 nm) was located at 219.2 °C,
slightly lower than that of as-received drug powder at 222.1 °C.

Fig. 6. Raman spectroscopy results for the obtained GF nanoparticles under the condition of flow rate at 6.8 mL/min in both tube side and shell side with drug
concentration at 0.040 g/mL and as-received GF powder.

686
X. Zhou et al. Journal of Membrane Science 564 (2018) 682–690

Fig. 7. FT-IR results of the obtained GF nanocrystals under the condition of flow rate at 6.8 mL/min in both tube side and shell side with drug concentration at
0.040 g/mL and as-received GF powder.

Fig. 8. XRD results of the obtained GF nanocrystals under the condition of flow Fig. 10. Dissolution profiles of as-received powder and synthesized drug na-
rate at 6.8 mL/min in both tube side and shell side with drug concentration at nocrystals under the condition of flow rate at 6.8 mL/min in both tube side and
0.030 g/mL and the as-received GF powder. shell side with drug concentration at 0.030 g/mL in PBS (phosphate buffer
saline) at pH 6.8.

dissolution rate of the nanocrystals (86.4 nm) increased to 54.0% in the


same period. The faster dissolution rate of drug nanoparticles was due
primarily to the increased specific surface area of obtained nano-
particles leading to much larger contact area with dissolution medium
compared with the crude micron-sized drug powders. Smaller nano-
crystals would get dissolved at even higher rates.

3.2. Variation of flow rate combinations

A series of flow rate combinations for the drug-containing solvent


and the anti-solvent (Table 1) were selected to study their effect on the
drug crystallization process and particle size in PHFAC. Shell-side
Reynolds number was calculated using the formula of Re = ρvd/μ. Here
Fig. 9. DSC results of the obtained GF nanocrystals under the condition of flow
ρ is the fluid density, v is the bulk velocity in the module, d is the
rate at 6.8 mL/min in both tube side and shell side with drug concentration at
characteristic dimension of the shell side channel and μ is the liquid
0.030 g/mL and the as-received GF powder.
viscosity. Reynolds number under the experimental conditions of F1, F2
and F3 was at 4.08, 7.00 and 10.10 respectively, showing a low Rey-
Regarding the drug nanocrystals, the endothermic peak did not notably nolds number laminar flow in the module. The results displayed in
alter, indicating high crystallinity consistent with that of as-received Table 1 and Fig. 11 show that the mean size of the attained drug na-
drug powder. noparticles was 190.3 nm under experimental condition F2, which was
also in accordance with the result shown in Fig. 12(b). Under condition
3.1.5. Dissolution testing F3, the nanoparticle size decreased to around 140 nm, as revealed in
Smaller particles have a higher surface area. Since particle dis- Table 1 and Fig. 12(c).
solution rate is proportional to the particle surface area, smaller parti- As the total flow rate of the drug solution in the shell side (sum of
cles should have a higher dissolution rate. Equilibrium solubility is also the flow rates of the drug solution on the shell-side and the tube-side
increased as the particle size is reduced. Therefore, dissolution rates of anti-solvent) increased, the particle size decreased due to the following
the nanoparticles formed should be higher than those of the larger reasons. The increasing total flow rate due to an increased drug solution
particles of as-received drug. The dissolution profiles for the as-received flow rate for a constant feed solution flow rate introduced an increasing
drug powder and the synthesized drug nanocrystals are shown in level of supersaturation for drug, which resulted in high rates of nu-
Fig. 10. The results suggest that over 12.4% of the as-received drug cleation and correspondingly less amount of material available for
(3–4 µm) got dissolved in the medium in around 130 min while the further growth. In addition, the shell-side residence time decreased

687
X. Zhou et al. Journal of Membrane Science 564 (2018) 682–690

Table 1
Particle size under variation of the ratio of the flow rates of the anti-solvent and the drug solution.
Experimental Tube-side flow rate (mL/min) Shell-side flow rate (mL/min) Residence time (s) Mean (nm) Standard Deviation (nm) d10 (nm) d90 (nm)

F1 6.8 6.8 14.9 295.6 31.6 263.0 333.0


F2 6.8 13.6 9.3 190.3 46.7 169.0 214.0
F3 6.8 20.4 5.0 144.6 24.0 127.0 163.0

6 mL/min in tube side and 11 mL/min in shell side for large module and
6.8 mL/min in tube side and 13.6 mL/min in shell side for small
module, the particle size could be significantly reduced as discussed
above. The membrane flux values were figured out with the formula of
J = V/(T × A). Here V is the volume of the anti-solvent in L, A is the
internal membrane surface area in m2, and T is the time in h. The flux
values of large pore hollow fiber membranes module and small pore
hollow fiber membranes module were respectively 415.19 L/m2 h and
79.29 L/m2 h. Fig. 13 demonstrates that the variation of pore size is an
effective way to reduce the particle size and thus to have some control
over crystallization for the precipitation of nanoparticles.
Fig. 11. Particle size distribution (PSD) under conditions of F1, F2, and F3.
3.4. Variation of drug concentration in feed solution
which reduced the opportunity for growth of the formed crystals.
In order to study the influence of drug concentration on the size of
A comparison of the particle size distribution and the TEM images
obtained drug particles, different concentrations of drug were used in
under various conditions show that the drug nanoparticles seemed to
the feed solution. As shown in Table 3, in experiment F1, the mean
remain crystalline with the same appearance and morphology.
particle size was 295.6 nm, which decreased to 141.8 nm when the drug
Consequently, it can be stated that the quality of drug nanoparticles
concentration was reduced under experimental condition of F4. Fur-
does not change when synthesized under different flow rate conditions.
ther, with the concentration decreasing to 0.030 g/mL in experiment
F5, the particle size dropped to 86.4 nm.
3.3. Variation of pore size of the hollow fiber membranes The reduced crystal size following a reduction in drug concentration
in the feed solution can be understood in the context of the processes of
Pore size in the hollow fiber membranes plays a critical role in the nucleation and crystal growth. First, owing to a lower concentration,
anti-solvent induced drug crystallization process since each area of tiny the drug dissolved in the solution was less sensitive to the anti-solvent
mixing zone was generated and controlled by water ejected from the resulting in a slower progress to supersaturation. In this situation, nu-
pore. The smaller the pore size on the hollow fiber wall are, the smaller cleation took place at a slow rate allowing only a reduced time for
the dimensions of the sub-micromixing zone will be generated when the growth if any. Further, there was less material available for growth of
penetrating water encounters the supersaturated drug solution in the the nuclei due to lower bulk concentration. Due to the local depletion of
shell side, resulting in a better control over the synthesized crystal size. concentration, the lack in material also contributed to reduction on
Compared with the previous study [23] in which micron-sized drug particle size. Consequently, the drug nanocrystals developed at a slow
crystals were successfully produced using the PHFAC method with pace and were swept out of the PHFAC crystallizer before much growth.
larger pore size membranes, the current method with smaller pore size
membranes was able to produce nano-sized drug crystals in a con- 3.5. Further considerations
tinuous and controllable manner. The specifications of two types of
hollow fiber membranes for the two studies are provided in Table 2. A few general comments on the membrane-based anti-solvent
The DLS results of the particle size distribution in Table 2 show that crystallization device are useful. First, such a device has exceptional
under optimized conditions with identical flow rate combination of fluid mixing capabilities via tiny anti-solvent fluid jets injected into the

Fig. 12. TEM images of nanocrystals gained under conditions of (a) F1, (b) F2, and (c) F3.

688
X. Zhou et al. Journal of Membrane Science 564 (2018) 682–690

Table 2
Particle size distributions for different hollow fiber membrane modules.
Material Membrane Pore size Fiber OD Fiber ID PSD Mean Standard Deviation PSD d10 PSD d90
(μm) (mm) (mm) (μm) (μm) (μm) (μm)

Large pore hollow fiber membranes Nylon-6 0.2–1.5 1.0 0.6 1.61 0.26 1.00 2.21
[26]
Small pore hollow fiber membranes PVDF 0.1–0.15 2.5 1.8 0.19 0.05 0.17 0.21
[this study]

Fig. 13. SEM images of crystals obtained from (a) hollow fibers having larger pores [23] and (b) hollow fibers having smaller pores.

shell-side axial flow [25]. Computational fluid dynamics-based studies 20.4 mL/min, the residence time could be further reduced. Accordingly,
of the mixing of two fluids in such a device [28] have demonstrated the size of the drug particles decreased with the decrease of the re-
such capabilities even though the Reynolds number of the shell-side sidence time since shorter residence time reduces the growth and nu-
flow is low and is in the laminar region. Consequently, supersaturation cleation time for newly produced GF crystals. With a decrease of drug
consumption is quite efficient in such a device. As the shell-side fluid/ concentration in the feed solution, the nanoparticle size could be re-
suspension moves, the solute concentration will decrease considerably duced to as low as 86.4 nm in mean size, while still retaining identical
along the axial direction. If the starting concentration of the solute in crystalline structure, appearance, and morphology. Moreover, the var-
the fluid entering the shell side is low, nanoparticles once formed will iation of pore size on the hollow fibers is a critical parameter to directly
hardly have any chance to grow due to rapid solute concentration de- influence the particle size and thus to create a more controllable crys-
pletion along the main flow axis. However, that does not prevent col- tallization environment for the production of nanoparticles. The smaller
lisions between nanoparticles in two separate streamlines since there is the pore size on the hollow fiber wall and narrower of the pore size
significant radial mixing due to anti-solvent injection through the distribution, the better is the control achieved over synthesized crystal
pores. There is an extraordinary opportunity of controlling various size. Under relatively similar conditions, the particle size of synthesized
parameters such as membrane pore size, incoming solute concentration, drug nanocrystals was reduced to 190.3 nm when using the small pore
length of device and the suspension velocity in such a configuration. size (0.1–0.15 µm) compared with the mean particle size of 1.61 µm
when applying large pore hollow fiber membrane (0.2–1.5 µm).
4. Concluding remarks In this technique, the porous hollow fibers crystallization device
enabled uniform and relatively controlled synthesis of drug nanocrys-
A porous hollow fiber crystallization device (PHFAC) was designed tals under ambient conditions. Moreover, the hollow fiber membrane
to successfully achieve continuous production of drug nanocrystals device can be easily scaled up by adjusting the numbers of the hollow
under room temperature and atmospheric pressure conditions. The as- fibers in the shell side while still not impacting the quality of produced
received micron-sized drug powder was first fully dissolved in an drug nanocrystals [28], which is promising to fit the needs of modern
acetone solution and pumped into the shell side of the module where industrial production. The achieved results in the study have demon-
the anti-solvent was already introduced through the lumen of the strated that the PHFAC technique for continuous manufacturing of drug
porous hollow fibers. The extensive mixing micro-environment gener- nanoparticles is feasible and realizable in a relatively straightforward
ated by the penetrating water streaming jets further provided con- way.
siderable supersaturation level that eventually led to nucleation and
crystallization of the drug nanoparticles. This device is flexible in its Acknowledgements
design and the interior parameters to allow relatively precise mon-
itoring and control of the drug nanocrystal production in a continuous This work was supported by the National Natural Science
manner. Foundation of China [grant number 81772278] and [grant number
By increasing the flow rates of shell side from 6.8 mL/min to 21706221] and the Natural Science Foundation of Fujian Province

Table 3
Mean particle size of drug crystals for different drug concentrations in the feed solution.
Experimental run Drug concentration (g/mL) Tube-side flow rate (mL/min) Shell-side flow rate (mL/min) Mean (nm) Standard Deviation (nm) d10 (nm) d90 (nm)

F1 0.040 6.8 6.8 295.6 31.6 263.0 333.0


F4 0.035 6.8 6.8 141.8 13.5 126.0 160.0
F5 0.030 6.8 6.8 86.4 14.1 73.5 101.0

689
X. Zhou et al. Journal of Membrane Science 564 (2018) 682–690

[grant number 2018J05143]. [14] K.R. Vandana, Y.P. Raju, V.H. Chowdary, M. Sushma, N.V. Kumar, An overview on
in situ micronization technique – an emerging novel concept in advanced drug
delivery, Saudi Pharm. J. 22 (2014) 283–289.
References [15] H.K. Chan, P.C.L. Kwok, Production methods for nanodrug particles using the
bottom-up approach, Adv. Drug. Deliv. Rev. 63 (2011) 406–416.
[1] F.D. Cui, X.Y. Long, Pharmacology, seventh ed., People's Medical Publishing House, [16] N. Rasenack, B.W. Muller, Dissolution rate enhancement by in situ micronization of
Beijing, 2011, p. 395. poorly water-soluble drugs, Pharm. Res. 19 (2002) 1894–1900.
[2] Y. Gao, J. Wang, Y. Wang, Q. Yin, B. Glennon, J. Zhong, J. Ouyang, X. Huang, [17] J.W. Tom, P.G. Debenedetti, Particle formation with supercritical fluids–a review, J.
H. Hao, Crystallization methods for preparation of nanocrystals for drug delivery Aerosol Sci. 22 (1991) 555–584.
system, Curr. Pharm. Des. 21 (2015) 3131–3139. [18] S.D. Yeo, E. Kiran, Formation of polymer particles with supercritical fluids: a re-
[3] P. Sheth, H. Sandhu, D. Singhal, W. Malick, N. Shah, M.S. Kislalioglu, Nanoparticles view, J. Supercrit. Fluids 34 (2005) 287–308.
in the pharmaceutical industry and the use of supercritical fluid technologies for [19] D.W. Matson, J.L. Fulton, R.C. Petersen, R.D. Smith, Rapid expansion of super-
nanoparticle production, Curr. Drug. Deliv. 9 (2012) 269–284. critical fluid solutions - solute formation of powders, thin-films, and fibers, Ind. Eng.
[4] W.E. Bawarski, E. Chidlowsky, D.J. Bharali, S.A. Mousa, Emerging nanopharma- Chem. Res. 26 (1987) 2298–2306.
ceuticals, Nanomed. Nanotechnol. Biol. Med. 4 (2008) 273–282. [20] S.H. Lee, D. Heng, W.K. Ng, H.K. Chan, R.B. Tan, Nano spray drying: a novel
[5] B. Sinha, R.H. Mueller, J.P. Moeschwitzer, Bottom-up approaches for preparing method for preparing protein nanoparticles for protein therapy, Int. J. Pharm. 403
drug nanocrystals: formulations and factors affecting particle size, Int. J. Pharm. (2011) 192–200.
453 (2013) 126–141. [21] G.G. Chen, G.S. Luo, J.H. Xu, J.D. Wang, Membrane dispersion precipitation
[6] H.B. Chen, C. Khemtong, X.L. Yang, X.L. Chang, J.M. Gao, Nanonization strategies method to prepare nanoparticles, Powder Technol. 139 (2004) 180–185.
for poorly water-soluble drugs, Drug. Discov. Today 16 (2011) 354–360. [22] R. Kieffer, D. Mangin, F. Puel, C. Charcosset, Precipitation of barium sulphate in a
[7] E. Merisko-Liversidge, G.G. Liversidge, Nanosizing for oral and parenteral drug hollow fiber membrane contactor: Part I Investigation of particulate fouling, Chem.
delivery: a perspective on formulating poorly-water soluble compounds using wet Eng. Sci. 64 (8) (2009) 1759–1767.
media milling technology, Adv. Drug. Deliv. Rev. 63 (2011) 427–440. [23] R. Kieffer, C. Charcosset, F. Puel, D. Mangin, Numerical simulation of mass transfer
[8] B. Van Eerdenbrugh, G. Van den Mooter, P. Augustijns, Top-down production of in a liquid-liquid membrane contactor for laminar flow conditions, Comput. Chem.
drug nanocrystals: nanosuspension stabilization, miniaturization and transforma- Eng. 32 (2008) 1333–1341.
tion into solid products, Int. J. Pharm. 364 (2008) 64–75. [24] D.M. Zarkadas, Crystallization Studies in Hollow Fiber Devices (Ph.d. thesis), New
[9] R. Shegokar, R.H. Müller, Nanocrystals: industrially feasible multifunctional for- Jersey Institute of Technology, New Jersey, 2004.
mulation technology for poorly soluble actives, Int. J. Pharm. 399 (2010) 129–139. [25] D.M. Zarkadas, K.K. Sirkar, Anti-solvent crystallization in porous hollow fiber de-
[10] B. Subramaniam, R.A. Rajewski, K. Snavely, Pharmaceutical processing with su- vices, Chem. Eng. Sci. 61 (2006) 5030–5048.
percritical carbon dioxide, J. Pharm. Sci. 86 (1997) 885–890. [26] D. Chen, D. Singh, K.K. Sirkar, R. Pfeffer, Continuous synthesis of polymer-coated
[11] J. Jung, M. Perrut, Particle design using supercritical fluids: literature and patent drug particles by porous hollow fiber membrane-based anti-solvent crystallization,
survey, J. Supercrit. Fluids 20 (2001) 179–219. Langmuir 31 (2015) 432–441.
[12] S.A. Desobry, F.M. Netto, T.P. Labuza, Comparison of spray-drying, drum-drying [27] D. Chen, D. Singh, K.K. Sirkar, R. Pfeffer, Porous hollow fiber membrane- based
and freeze-drying for beta-carotene encapsulation and preservation, J. Food Sci. 62 continuous technique of polymer coating on submicron and nanoparticles via anti-
(1997) 1158–1162. solvent crystallization, Ind. Eng. Chem. Res. 54 (19) (2015) 5237–5245.
[13] P. Teixeira, H. Castro, R. Kirby, Spray-drying as a method for preparing con- [28] D. Chen, B. Wang, K.K. Sirkar, Hydrodynamic modeling of porous hollow fiber anti-
centrated cultures of lactobacillus-bulgaricus, J. Appl. Bacteriol. 78 (1995) solvent crystallizer for continuous production of drug crystals, J. Membr. Sci. 556
456–462. (2018) 185–195.

690

Das könnte Ihnen auch gefallen