You are on page 1of 52

Advanced Drug Delivery Reviews 58 (2006) 1274 – 1325

www.elsevier.com/locate/addr

Mathematical modeling and simulation of drug release from


microspheres: Implications to drug delivery systems ☆
Davis Yohanes Arifin a , Lai Yeng Lee a , Chi-Hwa Wang a,b,⁎
a
Molecular Engineering of Biological and Chemical Systems Program, Singapore–MIT Alliance, 4 Engineering Drive 3,
Singapore 117576, Singapore
b
Department of Chemical and Biomolecular Engineering, National University of Singapore, 4 Engineering Drive 4,
Singapore 117576, Singapore
Received 13 July 2006; accepted 4 September 2006
Available online 26 September 2006

Abstract

This article aims to provide a comprehensive review of existing mathematical models and simulations of drug release from
polymeric microspheres and of drug transport in adjacent tissues. In drug delivery systems, mathematical modeling plays an
important role in elucidating the important drug release mechanisms, thus facilitating the development of new pharmaceutical
products by a systematic, rather than trial-and-error, approach. The mathematical models correspond to the known release
mechanisms, which are classified as diffusion-, swelling-, and erosion-controlled systems. Various practical applications of
these models which explain experimental data are illustrated. The effect of γ-irradiation sterilization on drug release mechanism
from erosion-controlled systems will be discussed. The application of existing models to nanoscale drug delivery systems
specifically for hydrophobic and hydrophilic molecules is evaluated. The current development of drug transport modeling in
tissues utilizing computational fluid dynamics (CFD) will also be described.
© 2006 Elsevier B.V. All rights reserved.

Keywords: Release mechanism; Polymeric system; Diffusion; Swelling; Erosion; Irradiation; Transport; Tissue; Brain; Computational fluid
dynamics

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1275
2. Mathematical models for diffusion-controlled systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1276
2.1. Reservoir systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1276


This review is part of the Advanced Drug Delivery Reviews theme issue on “Computational Drug Delivery", Vol. 58/12-13, 2006.
⁎ Corresponding author. Department of Chemical and Biomolecular Engineering, National University of Singapore, 4 Engineering Drive 4,
Singapore 117576, Singapore. Tel.: +65 6516 5079; fax: +65 6779 1936.
E-mail address: chewch@nus.edu.sg (C.-H. Wang).

0169-409X/$ - see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.addr.2006.09.007
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1275

2.2. Matrix systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1277


2.2.1. Dissolved drug systems (C0 b Cs ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1277
2.2.2. Dispersed drug systems (C0 N Cs ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1278
3. Mathematical models for swelling-controlled systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1282
4. Mathematical models for erosion-controlled systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1290
4.1. Empirical models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1292
4.2. Mechanistic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1293
4.2.1. Diffusion-and-reaction models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1293
4.2.2. Cellular-automata models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1301
5. Irradiation effect on drug release profile from polymeric microspheres . . . . . . . . . . . . . . . . . . . . . 1305
6. Going to nanoscale: the implication on drug release mechanism . . . . . . . . . . . . . . . . . . . . . . . . . 1307
6.1. Drug targeting and biodistribution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1307
6.2. Drug release from nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1308
6.2.1. Hydrophobic drugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1308
6.2.2. Hydrophilic drugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1310
7. Simulation of drug delivery in tissue: linking drug release profile and tissue elimination kinetics to predict
temporal and spatial drug transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1311
7.1. Simulation of drug delivery in brain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1313
7.2. Simulation of drug delivery in liver . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1317
7.3. Simulation of drug delivery in bone. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1318
7.4. Challenges ahead the simulation of drug delivery in tissue . . . . . . . . . . . . . . . . . . . . . . . . 1319
8. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1319
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1320
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1320

1. Introduction In all three systems, diffusion is always involved.


For a non-biodegradable polymer matrix, drug release
For decades, polymeric systems have been used for is due to the concentration gradient by either diffusion
pharmaceutical applications, especially to provide or matrix swelling (enhanced diffusion). For biode-
controlled release of drugs. Drug–polymer systems gradable polymer matrix, release is normally con-
may also be useful in protecting the drug from biolog- trolled by the hydrolytic cleavage of polymer chains
ical degradation prior to its release. The development that lead to matrix erosion, even though diffusion may
of these device starts with the use of non-biodegrad- be still dominant when the erosion is slow. This
able polymers, which rely on the diffusion process, categorization allows mathematical models to be
and subsequently progresses to the use of biodegrad- developed in different ways for each type of system.
able polymers, in which swelling and erosion take Mathematical modeling of drug release provides
place. insights concerning mass transport and chemical
Based on the physical or chemical characteristics of processes involved in drug delivery system as well
polymer, drug release mechanism from a polymer as the effect of design parameters, such as the device
matrix can be categorized in accordance to three main geometry and drug loading, on drug release mecha-
processes (systems) [1], which are: nism. Thus, the optimized device design for a
required drug release profile can be predicted using
1. Drug diffusion from the non-degraded polymer a systematic approach with a minimum number of
(diffusion-controlled system). experimental studies.
2. Enhanced drug diffusion due to polymer swelling This review presents the concepts contained in
(swelling-controlled system). important and readily available mathematical models
3. Drug release due to polymer degradation and for controlled release primarily from microspheres.
erosion (erosion-controlled system). Mathematical models for cylindrical geometry, especially
1276 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

for swelling- and erosion-controlled systems, will also 3. The surrounding system is a well-stirred finite
be discussed as real systems are readily available in volume. This implies that the concentration of the
this geometry. It also discusses recent improvements surrounding system changes with time. The surface
and the major advantages and limitations of each resistance may or may not be negligible.
model. The original notations are not retained in this
review; instead a common notation is used to Based on the matrix region where the drug diffusion
facilitate understanding and comparison between the primarily takes place, the diffusion-controlled system
models. The penultimate section of this review can be further categorized to reservoir and matrix
discusses the implication of phenomena at the systems. The reservoir system consists of a drug reser-
nanoscale and the last section focuses on coupling to voir surrounded by the polymer matrix shell. On the
transport in tissue. other hand, in matrix system, the drug is incorporated in
the polymer matrix in either dissolved or dispersed
2. Mathematical models for diffusion-controlled condition. The schematic illustrations of drug loading
systems distribution for these systems are shown in Fig. 1.

For diffusion-controlled microspheres, drug re- 2.1. Reservoir systems


lease profile is obtained by solving Fick's second
law of diffusion subject to appropriate boundary The reservoir model is the simplest model of a
conditions. For one-dimensional drug release from a solute of drug released from a sphere [3]. It assumes
microsphere, the second Fick's law of diffusion is that drug is confined by a spherical shell of outer radius
given by: R and inner radius Ri; thus, the drug must diffuse
  through a layer of thickness (R − Ri).
∂C 1 ∂ ∂C Fick's second law of diffusion is solved to obtain the
¼ 2 Dr2 ð1Þ
∂t r ∂r ∂r drug distribution within the shell with boundary
conditions imposed as in a hollow sphere (Ri ≤ r ≤ R),
where D and C are the diffusion coefficient and where the surface r = Ri is kept at a constant reservoir
drug concentration in the polymer matrix. The concentration Cr =KrCi, where Cr is the reservoir drug
boundary conditions are influenced by the mass concentration and Kr is the drug partition coefficient
transfer process at the surface and the volume of the between the reservoir and polymer matrix. On the other
surrounding system. Based on these conditions, hand, at the shell surface (r = R), drug concentration is
there are three main cases which are commonly assumed to be zero when there is no mass transfer
considered [2]: limitation and the surrounding volume is large.
This unsteady diffusion equation is solved to give
1. The mass transfer resistance at the surface is the concentration profile and the cumulative amount of
negligible and the surrounding release medium is drug released (Mt) as functions of time, as follows:
infinitely large (perfect sink condition), implying 
that the concentration on the surface of the Ri ðR−rÞ 2 X l
Ci Ri r−Ri
Cðr; tÞ ¼ Ci − sin np
matrix (Cs) is constant (Cs = KCb = constant at r ðR−Ri Þ p n¼1 n R−Ri
r = R). Here, Cb is the drug concentration in !
n2 p2
surrounding bulk medium and K is the drug exp − Dt ð2Þ
partition coefficient between the matrix and bulk ðR−Ri Þ2
medium.
2. The mass transfer resistance at the surface is finite Mt Dt 1
¼ −
and the surrounding volume is in perfect sink condi- 4pRi RðR−Ri ÞCi ðR−Ri Þ 6
2
!
tion, implying that the concentration of the sur- 2X l
ð−1Þn n2 p2
rounding system is constant, but the convective mass − 2 exp − Dt
p n¼1 n2 ðR−Ri Þ2
transfer coefficient
  (h) will determine the surface

concentration −D ∂C ¼ hðCjr¼R −KCb Þ . ð3Þ
∂r r¼R
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1277

Fig. 1. Schematic illustration of cross-section of drug-loaded spheres of (a) reservoir system, (b) dissolved drug system, and (c) dispersed drug
system. In reservoir system, drug is confined by a spherical shell of outer radius R and inner radius Ri; therefore, the drug must diffuse through a
polymer layer of thickness (R − Ri). In dissolved drug system, drug is dissolved uniformly at loading concentration C0 in the polymeric matrix. In
dispersed drug system, the radius of inner interface between “core” (non-diffusing) and matrix (diffusing) regions, r′(t), shrinks with time. The
“core” region is assumed to be at drug loading concentration C0.

When the observation time is long enough (t → ∞), A very popular biocompatible, but non-biodegrad-
the series term approaches zero and there is a constant able, polymer that is representative of the matrix system
diffusion rate of drug leaving the reservoir sphere; is poly(ethylene-co-vinyl acetate) (EVAc). EVAc is a
thus, the amount of drug released equation can be hydrophobic polymer that swells less than 0.8% in
simplified as follows: water. It is commercially available with composi-
tion range of 10–40%-w/w vinyl acetate.
Mt RRi
¼ 4p DCi ð4Þ
t ðR−Ri Þ 2.2.1. Dissolved drug systems (C0 b Cs )
However, the use of both solutions has to be taken Dissolved drug system prevails when initial drug
carefully since it seems to suggest that Mt increases loading concentration (C0) is below the drug saturation
without limit as time increases. One must consider that concentration in the polymer matrix (Cs); therefore, the
the drug content in sphere is finite and the diffusional drug is dissolved uniformly in the polymer matrix.
release is only valid for a certain time range that may When the surface resistance to mass transfer at the
differ for different systems. surface is negligible, the drug concentration profile and
fractional amount of drug released can be expressed as
follows [3]:
2.2. Matrix systems
2R X ð−1Þn npr
l
C−C0
¼1þ sin ð5Þ
Mathematical models for matrix systems are often KCb −C0 pr n¼1 n R
valid for drug devices developed based on non- 
−Dn2 p2 t
biodegradable polymers. In these models, the drug exp
is commonly assumed to be uniformly distributed R2
inside the non-biodegradable polymer matrix. There

are two possible cases, which are (i) the initial drug Mt 6Xl
1 −Dn2 p2 t
loading is lower than the solubility of the drug inside ¼ 1− 2 exp ð6Þ
Ml p n¼1 n2 R2
the polymer matrix (C0 b Cs), which implies a dis-
solved drug system, and (ii) the initial drug loading is where M∞ is the cumulative drug released at infinite
higher than the solubility of the drug inside the time.
polymer matrix (C0 N Cs), which implies a dispersed In contrast, if finite
 convective mass transfer pre-
drug system. vails, where −D ∂C ∂r r¼R ¼ hðCjr¼R −KCb Þ, the drug
1278 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

concentration profile and fractional amount of drug drug contained in the reservoir region does not
released can be expressed as follows [2]: incorporate polymer matrix. On the other hand, the
“core” region is able to shrink as drug released via
C−KCb 2ShR X
l
1 sinbn Rr diffusion in the diffusion region. The necessity to
¼ ð7Þ
C0 −KCb r n¼1 b2n þ Sh2 −Sh sinbn involve the moving-boundary problem makes the
 2 governing equation of the mathematical model for
b
exp − n2 Dt dispersed drug system difficult to solve analytically. A
R
schematic illustration of concentration profiles for
dispersed drug system is shown in Fig. 2.
Xl  2
Mt 6Sh2 b The diffusion-controlled mathematical model for
¼ 1− exp − n2 Dt ð8Þ dispersed drug system (C0 N Cs) in a planar sheet and a
Ml n¼1 b ðb
2 2
n n þ Sh 2
−ShÞ R
sphere was initiated by Higuchi [6] with the assump-
where Sh is Sherwood number defined as Sh = hR D , and tion that diffusion is in pseudo-steady state. In the
the βns are roots of the equation βncotβn = 1 − Sh. Eqs. planar system, this assumption leads to a linear con-
(7) and (8) are identical to Eqs. (5) and (6) when centration profile of the drug in the diffusing region,
Sh ≫ 1. It can be seen that the finite mass transfer which is between the dissolution interface and the
solution depends on Sh, which is the ratio between initial matrix surface. Given this assumption, the
mass transfer resistance at the surface and diffusional simplest and most popular version of Higuchi's
resistance throughout the polymer matrix. equation for planar system is readily obtained:
For Sh ≫ 1, Baker and Lonsdale [4,5] also pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
proposed two simplified solutions by using early- Mt ¼ S ð2C0 −Cs ÞCs Dt ð11Þ
time and late-time approximations to explain the drug
release from dissolved matrix system of a sphere Here, S is the available surface area for drug release
during early and late-time periods respectively. The to surrounding. For a spherical system, the pseudo-
approximation results are expressed as follows: steady concentration profile in the diffusing region is
C r V ðR−rÞ
 1=2 ¼ . Integration of the mass flux equation
Cs r ðR−r VÞ
Mt Dt 3Dt Mt
¼6 2
− 2 for b0:4 ð9Þ gives the relationship between the moving-boundary
Ml pr r Ml interface position (r′) and time (t) as follows:
 2 6DCs Rt ¼ C0 ðR
3
þ 2r V3 −3Rr V2 Þ
Mt 6 p Dt
¼ 1− 2 exp − 2 for
Mt
ð10Þ
p
N0:6 R
Ml r Ml þ Cs 4r V2 R þ R3 ln −R3 −R2 r V−2r V3
rV
2.2.2. Dispersed drug systems (C0 N Cs ) ð12Þ
Similar to the reservoir system, the polymer matrix
in the dispersed drug system can be primarily divided In the case of C0 N Cs, this solution can be further
into two regions, which are (i) the “core” (non- simplified to:
diffusing) region, in which the undissolved solute  2  3
6DCs rV rV
exists at concentration C0, and (ii) the dissolved t ¼ 1−3 þ2 ð13Þ
(diffusing) region, where all solute is dissolved and C0 R2 R R
diffusion occurs. The distinct separation for these two
The fractional cumulative release can be expressed
regions is valid when C0 N Cs, therefore the limited
as follows:
application of these mathematical models must be
 3
carefully taken into account when C0/Cs is not Mt rV
significantly large. Here, the “core” region is different ¼ 1− ð14Þ
Ml R
from the reservoir region in the reservoir system since
the former is a constant thickness polymer region with Koizumi and Panomsuk [7] applied Higuchi's
certain drug loading concentration (C0), whereas the pseudo-steady approach to obtain an approximation
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1279

ancy. Here, a1, a2 and a3 are constants for the quadratic


radial-dependent concentration profile to be satisfied
by the perfect sink boundary conditions. When the
boundary conditions are satisfied, it results in the form
as follows:
2 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 3
 2 ffi 
C 4 C0 C 0 5 R−r
¼ 1þ − −1 ð16Þ
Cs Cs Cs R−r V
2 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 3
 2 ffi 
4C 0 C 0 5 R−r 2
− − −1
Cs Cs R−r V

In addition, the relationship between the moving front


r′ and t can be given by,
2 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 3
Fig. 2. A schematic diagram illustrating concentration profile in a   2 ffi 
Dt 1 4 C0 C0 5 R−r V 2
sphere of dispersed drug system with the presence of a boundary ¼ 5 −4 þ −1
layer in a perfect sink medium. The drug concentration at the matrix R2 12 Cs Cs R
surface (C′)s is related to drug concentration at the boundary layer
 
interface with the matrix (Ca) with the partition coefficient K1 (C′s = 1 C0 R−r V 3
K1Ca). If boundary layer or coating does not exist, the drug − −1 ð17Þ
concentration at the matrix surface (C′)
3 Cs R
s is directly related to drug
concentration in the bulk medium (Cb) with the partition coefficient
K (C′s = KCb) (reprinted from [15] with permission from Elsevier). Finally, the resulting fractional release can be written
as follows:
"  # 
solution for drug release from a sphere. With several Mt R−r V 3 Cs R−r V Cs
series simplification, a simple and explicit expression ¼ 1− 1− 1− þ3
Ml R C0 R C0
can be obtained for the amount of drug released as h   i 
function of time for a perfect sink condition as follows: a2 a3 a1 a2 a3 R−r V
 a1 þ þ − þ þ
   2 3 2 3 4 R
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 4Cs Cs
Mt ¼ 4pR 2
2ðC0 −Cs ÞDt þ Dt ð18Þ
9R ð2C0 −Cs Þ
ð15Þ Even though Lee's solution is explicit and easy to
apply, several models that provide exact solutions have
However, it must be noted that the solution does not also been developed to remove the discrepancy of
suggest Mt increases to infinity as time increases as approximate solutions. Paul and McSpadden [10]
only finite amount of drug is available in the sphere. obtained the exact solution for drug release from a
The Higuchi approximation (Eqs. (12)–(14)) pro- planar system into a perfect sink. It turns out that Lee's
vides a good correlation of drug release profile when approximate solution agrees well with it. Abdekhodaie
C0 N Cs; but, in the limit of C0 → Cs, its result does not and Cheng [11,12] try to develop an exact solution for
precisely match the classical diffusion solution for the diffusion from a spherical matrix into both infinite and
dissolved drug system (C0 b Cs) at the limiting case finite external medium by using a combination of
C0 → Cs. Cohen and Erneux [8] suggested a boundary variables technique. This technique allows the reduc-
layer analysis to accommodate the discrepancy that led tion of Fick's second law of diffusion into an ordinary
to a simple analytical solution for the release rate with differential equation. The finite external medium is
the effect of time-dependent solubility. Lee [9] applied assumed to have concentration of Cb. It is related to the
the approximate  quadratic
 concentration profile, concentration at the matrix surface (r = R) by the
ðCs −CÞ R−r R−r 2
C
¼ a1 þ a2
R−r V
þ a3
R−r V
, to reduce the discrep- partition coefficient K (Cs = KCb), implying that the
s
1280 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

magnitude of mass transfer at the surface is finite as it neglected at sufficiently small τ. Here, Γ and τ are
depends on the bulk concentration. The effect of finite related implicitly as follows:
volume and the partition coefficient is lumped into an " # 
pffiffiffi ð1−CÞ ð1−CÞ2 1−C 1
additional variable termed the effective volume ratio p pffiffiffi exp erf pffiffiffi ¼   ð24Þ
3Vb 2 s 4s 2 s
(λ), which is defined as k ¼ . The solution is C0
−1
Cs
4KpR3
obtained as follows:
    Wu and Zhou [14,15] developed a numerical solution
pffiffiffi 1−f 1−C
h ¼ A1 þ A2 p erf pffiffiffi −erf pffiffiffi ð19Þ for a finite external medium system (Vb) with an
2 s 2 s additional diffusion boundary layer incorporated on the
where surface of the matrix. Therefore, there are two diffusing
zones, in which the drug has different diffusivities. The
r Dt rV r C first diffusing zone is the polymeric matrix region
f¼ ; s¼ 2; C¼ ; h¼ ; and
R R R R Cs between r′ and R in which diffusivity is D, whereas the
ð20Þ
C0 second diffusing zone is a constant thickness (δ)
j¼ −1 boundary layer in which diffusivity is Da. In addition,
Cs
the model accounts for the number of microspheres (N);
However, the combination variables technique fails to therefore, the total drug released becomes Mt = Nm,
reduce the boundary conditions to obtain the appropri- where m is the cumulative amount of drug released from
ate integration constants. Therefore, even though the a single microsphere. Despite the model improvement to
solution is aimed to predict drug release in a finite account for the presence of a boundary layer, the
medium, the solution is basically valid for the cases of pseudo-steady-state approximation of Higuchi is as-
perfect sink condition with pseudo-steady assumption sumed for both diffusing regions; therefore, the con-
which is similar to Higuchi's approach. As a rule of centration profile for the diffusing region in the matrix is
thumb, the pseudo-steady approximation is considered expressed as follows:
to be valid when CC0s N 3. The numerical solutions using a
finite element method for slab, cylinder, and sphere Rðr−r VÞ
C ¼ Cs −ðCs −Cs VÞ ð25Þ
matrices are available in other literatures [13]. Fig. 3a rðR0 −r VÞ
shows the concentration profile at the moving-boundary
for drug release into perfect sink medium, whereas Fig. where Cs′ is the concentration at the moving inter-
3b depicts the influence of λ on the total amount of drug face between the dissolved and boundary layer regions
released into finite volume medium; both are shown for (C′s = K1Ca) and Ca is the drug concentration at the
the case of CC0s = 2. Other simpler numerical methods can boundary layer interface with the matrix. K1 is the
be employed to solve this problem by satisfying the partition coefficient between the matrix surface and the
following boundary conditions: boundary layer. Here, it is assumed that there is no drug
accumulation in the boundary layer region so the rate of
hðC; sÞ ¼ 1 ð21Þ drug diffusion in the matrix region of the microspheres is
the same as that through the boundary layer. The
Z 1
C0 schematic illustration of the drug concentration in this
ð1−CÞ ¼ hk þ hdf at f ¼ 1 ð22Þ system is shown in Fig. 2. The cumulative drug release
Cs C
from a single microsphere (m) is given by:
The variation of moving-boundary (Γ ) with time is
defined as the mass balance at the interface r′ as follows:
Z R
  4pC0 3 3
C0 dC 1 ∂h m¼ ðR −r V Þ− 4pr2 Cdr
−1 ¼ −1 ð23Þ 3 rV
Cs ds C ∂f
where initial condition (Γ0,τ0) can be adopted from the
¼
4pC0 3 3
3
ðR −r V Þ þ
2p
3
½
ð2r V3 −Rr V2 −R2 r VÞCs ð26Þ

−ð2R −Rr V −R r VÞC V


Paul and McSpadden solution [10] for small value of τ 3 2 2
s
as the curvature effect in the spherical geometry can be
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1281

Fig. 3. (a) A moving-boundary concentration profile for a sphere of dispersed drug system in perfect sink condition with R = 1 × 10− 3 m,
D = 1 × 10− 10 m2/s, and C0 / Cs = 2; and (b) drug release profiles into finite volume medium with various effective volume ratios (λ) with
R = 1 × 10− 3 m, D = 1 × 10− 10 m2/s, and C0 / Cs = 2. The 100% drug release is obtained by the plateau region of release curve for λ = ∞. Both panels
are obtained by solving the moving-boundary problem of dispersed matrix system using finite element method (reprinted from [13] with
permission from Wiley-Liss, Inc., a subsidiary of John Wiley & Sons, Inc. Copyright © 1999).

The total amount of drug released (Mt) for a number When this relationship is simplified to the special
of microspheres (N) and drug concentration in the case of C0 N Cs, the amount of dissolved drug in the
finite external medium (Cb) are expressed respectively matrix can be ignored and the following explicit solution
as follows: can be obtained:
Mt ¼ mN ð27Þ

Cb ¼
Mt
¼
3m Vs
ð28Þ
tðr VÞ ¼
−Vb R2
6K1 Vs DB1 f
"
pffiffiffi
lnðB2 Þ þ 2 3tan−1 ðB3 Þ
 3 ! #

g
Vb 4pR3 Vb 
DdK1 rV Vs C0
−2B 1− ln 1−K2 1−
where Vb is the volume of the finite external medium Da R R Vb Cs
and Vs = W / ρ is the total volume of the microspheres.
W and ρ are the mass and bulk density of microspheres, ð30Þ
respectively.
Upon differentiation of the accumulative mass  3 #"
release with time and equating to the drug release rV
Mt ¼ Nm ¼ Vs C0 1− ð31Þ
from the microspheres, the relationship between the R
moving-boundary r′ and time t is obtained as follows:
Z where B1, B2, and B3 are complex variables given by the
rV
½−4pC0 r2 þ gðr VÞðR−rÞ following expressions,
tðr VÞ ¼ dr ð29Þ
R 4pDRrðCs −Cs VÞ

where g(r′) is a derivative function of the amount of Cs Vb 1=3
B1 ¼ 1−
drug that is still diffusing in the matrix diffusing zone C0 Vs K1
and is defined as follows: 
1 þ B1 þ B21 r V=R−B1 2
B2 ¼ ð32Þ
d 2p ðr V=RÞ2 þ B1 r V=R þ B21 1−B1
gðr VÞ ¼ ½ð2r V3 −Rr V2 −R2 r VÞCs
dr V 3 pffiffiffi
−ð2R3 −Rr V2 −R2 r VÞCs V g B3 ¼ 2
2 3B1 ðr V=R−1Þ
3B1 þ ð2r V=R þ B1 Þð2 þ B1 Þ
1282 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

3. Mathematical models for swelling-controlled leads to matrix swelling that results in the rubbery (gel
systems layer) region, in which there is an “enhanced diffusion”
where drug mobility increases. The polymer will also
The idea of using a swelling polymer is to provide dissolve at the interface when the entanglement is weak
more control over the release of drug, especially when since polymer concentration is very low. Thus, in this
its diffusivity in polymer is very low. For this purpose, a system, the deviation from Fickian model is observed
swellable device is commonly made using a hydrophilic when the drug release is not only controlled by the
polymer so that water is able to imbibe into the polymer diffusion of the drug inside the matrix, but also by the
matrix and cause polymer disentanglement. The level of polymer matrix disentanglement and dissolution
polymer disentanglement as a function of polymer process.
concentration is illustrated in Fig. 4. The imbibing water For swelling-controlled system, the hydrophilic
into the polymer matrix decreases the polymer polymer is susceptible to swelling as water tends to
concentration and changes the level of polymer disen- penetrate and relax the polymer matrix. In this case, the
tanglement. The polymer matrix disentanglement also composition of the hydrophilic polymer will determine

Fig. 4. A schematic illustration of level of polymer matrix disentanglement as a function of the polymer concentration in a swelling-controlled
drug system. The disentanglement is due to the imbibing of water into the system that results in polymer concentration changes in the matrix. In
the dry glassy core, polymer concentration is very high. In the swollen glassy layer, solvent diffusion creates a more mobile network, but with very
strong chain entanglement. In the gel layer, polymer and solvent concentrations are comparable. Finally, in the high water-rich layer of diffusion
layer, chain entanglement becomes weak so that the polymer can disentangle and dissolve at the interface (reprinted from [28] with permission
from Wiley-Liss, Inc., a subsidiary of John Wiley & Sons, Inc. Copyright © 1995).
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1283

the extent of swelling. Typical hydrophilic polymers layer (rubbery), exist in the polymer matrix. Here, the
used for drug delivery application that exhibit swelling concept of two moving fronts, namely the glassy–
behavior include hydroxypropyl methylcellulose rubbery front (R) and the rubbery–solvent front (S), is
(HPMC), poly(hydroxyethyl methacrylate) or poly introduced. Initially during swelling, front R moves
(HEMA), and poly(vinyl alcohol) (PVA). Since inwards, whereas front S moves outward. When the
swelling-controlled drug delivery devices are often polymer at interface S reaches its thermodynamic
made as tablets and the overall drug release mechanism equilibrium with the surrounding medium, interface S
strongly depends on the design (composition and will start dissolving and, therefore, front S moves
geometry) of the device, corresponding mathematical inwards. Both fronts will move inwards until the front
models for this system are commonly developed for a R diminishes as the glassy core disappears. Subse-
cylindrical geometry, while extension to other geom- quently at later time, only the rubbery region is present
etries is readily available with proper coordinate and dissolution at interface S eventually controls the
transformation. In addition, several studies addressing shrinking process.
the modeling of swelling-controlled systems are also Upon contact with water, the drug dissolves due to a
available in literatures [16–18]. concentration difference at interface R and diffuses out
A model that describes the mechanism of polymer through interface S due to a concentration gradient
matrix swelling is shown in Fig. 5. This model is between interface of two states. If water penetration is
developed by Lee and Peppas [19] by taking into ac- negligible, polymer relaxation does not occur and drug
count swelling moving fronts. The swelling occurs to release is controlled by Fickian diffusion through the
achieve thermodynamics equilibrium when water pen- glassy polymer matrix. On the other hand, when the
etrates crosslinked region inside the polymer matrix water mobility is dominant in penetrating the polymer
due to a water concentration gradient. As water pene- matrix, a “non-Fickian Case-II transport” of drug re-
trates and swelling takes place, a transformation of lease takes place [20]. It is characterized by the exis-
polymer from a glassy to rubbery state occurs and the tence of a sharp interface advancing at constant
dimension of the matrix increases. This change of state velocity and the drug release is zero-order since it is
basically creates a gel layer of rubbery region for the controlled by the polymer dissolution process at the
drug to diffuse, in which the drug diffusivity in- moving interface. Between these two extreme cases,
creases substantially. Therefore, during the swelling, the “anomalous transport”, which has intermediate
two different states, namely the glassy core and gel characteristics, is defined. The “anomalous” transport

Fig. 5. Schematic illustration of one-dimensional swelling process due to solvent diffusion and polymer dissolution as proposed by Lee [19]: (a)
initial thickness of the carrier, (b) early-time swelling when there are increasing position of the rubbery/solvent interface (S) and decreasing
position of the glassy/rubbery interface (R), (c) late-time swelling when there are decreases of both interface S and R positions, and (d) final
dissolution process when the slab only comprises rubbery region with the decrease of interface S (reprinted from [19] with permission from
Elsevier).
1284 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

of drug release is often identified in swelling- Modification of the Korsmeyer's semi-empirical


controlled systems since both diffusion and dissolution equation by incorporating the concept of lag time was
occur altogether and they are quite indistinguishable. introduced by Kim and Fassihi [25] and Ford et al.
Here, the drug transport is commonly modeled as the [26]. Having accounted for the burst effect, the
diffusion in which the diffusion coefficient depends modified equation is given by:
strongly on polymer concentration since polymer
swelling and relaxation enhance the drug mobility, Mt
whereas the polymer dissolution is modeled to follow ¼ kðt−tlag Þn þ b ð34Þ
Ml
first-order kinetics at the interface with surrounding
medium.
where k is a constant and b is the total fractional drug
In the simplest manner, Korsmeyer et al. [21,22]
released from the burst effect.
developed a semi-empirical equation based on a
Another modification to the semi-empirical model
power-law expression to describe the drug release
was performed by Peppas and Sahlin [27] by de-
from swelling-controlled systems as follows:
coupling diffusion and “Case-II transport” with the
Mt following expression:
¼ kt n ð33Þ
Ml
Mt
¼ k1 t m þ k2 t 2m ð35Þ
Here, k is a constant and n is the diffusional Ml
exponent. The power-law equation can be observed
as the superposition of two processes of Fickian where k1, k2, and m are constants. The first term of
diffusion and “Case-II transport”. It has a feature to the right-hand side is the diffusional contribution,
identify the relative importance between Fickian whereas the second term is the “Case-II transport”
diffusion (n = 0.5) and “Case-II transport” (n = 1). contribution.
Aforementioned, in the latter case, the drug disso- Ju et al. [28,29] developed scaling laws for
lution at the moving front due to water imbibition is predicting polymer and drug release profiles from
the dominant step so it is characterized by a linear hydrophilic polymer matrices. The system used is the
time dependence of drug release. In between these HPMC matrix tablet in water. An anisotropic expan-
two processes, the “anomalous transport” takes place sion model is used to account for the anisotropic ex-
where phenomena of both processes are coupled pansion of the matrix when anisotropic swelling
(0.5 b n b 1). It is important to note that this rule of occurs. The model provides a qualitative relationship
thumb is particularly valid for slab geometry. between the polymer disentanglement concentration
Different values of n for cylindrical and spherical (ρp,dis) and the equivalent molecular weight of the
geometries are available in the literature [23,24]. For polymer matrix (Meq) in a HPMC/water system. The
spheres, the n values are 0.43 and 0.85 for diffusion polymer disentanglement concentration (ρp,dis) is
and “Case-II transport” drug release, respectively. defined as the concentration below which polymer
Beside the dependence on the geometry, the value of chains start dissolving or detaching from the gel
n may also be influenced by particle size distribution matrix, whereas the equivalent molecular weight (Meq)
as observed by Ritger [23,24]. For a hypothetical is the weight-averaged molecular weight of the
mixture of 20% 20-μm, 60% 100-μm, and 20% 500- polymer matrix. Results of these studies are summa-
μm spheres, the n values for Fickian diffusion and rized in two scaling laws for the fractional polymer and
the “Case-II transport” become 0.30 and 0.45, drug releases respectively, namely MMp;l
p;t −1:05
∝Meq and MMt ∝Meq−0:24.
l
M M
respectively. In comparison to the release profile In these scaling laws, M and M are the cumulative
p;t
p;l
t
l
from monodispersed particle, this mixture shows an fraction release of polymer and drug, respectively. The
acceleration of drug transport at early times due to scaling laws are developed based on radial mass transfer
the portion of particles smaller than the mean size model since the expansion for tablet matrix systems is
and retardation at longer times due to the portion of mostly in the axial direction and, therefore, the
particles larger than the mean size. dissolution surface is available mostly in the radial
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1285

direction. These results imply that drug release is less Harland et al. [32] improved Lee's moving front
sensitive to the polymer molecular weight as compared model to explain the mechanism of drug release from
to the polymer release. This is attributed to the assump- swelling tablets by adding a third component, which is
tion of infinite drug convective mass transfer coefficient the drug. Since drug is additionally incorporated in this
at the gel–diffusion layer interface. Therefore, the model, three components are now taken into account,
outward drug release by diffusion is less influenced by namely water (component 1), polymer (component 2),
the polymer molecular weight change than the polymer and drug (component d). In this model, the transport of
release that is controlled by the disentanglement and drug and water components is assumed to be by Fickian
dissolution. diffusion and expressed in volume-fraction form. At
Since empirical models are not able to provide the interface R, the polymer (υ*2 ) and drug (υ*d ) volume
sufficient physical understanding on how the swell- fraction reach a thermodynamic equilibrium, whereas,
ing process affects drug release, several mechanistic at the interface S, polymer disentanglement occurs and
models have been developed to capture the swell- the water (υ1,eq) and drug (υd,eq) volume fraction are in
ing mechanism. The development is focused on the the equilibrium with the surrounding medium. There
modeling of moving fronts of glassy and rubbery are two periods of swelling observed from the
regions during the swelling, as shown in Fig. 5, and the experimental results, namely the early swelling and
feature of concentration-dependent diffusivity to the front synchronization period. The early swelling
account for the increase of drug mobility in the gel period refers to the inward moving of interface R due to
layer region. Here, the evolving polymer and drug polymer and drug dissolution and outward moving
concentrations at the moving fronts can be predicted interface S due to polymer swelling. Here, the gel layer
by classical thermodynamics theories of polymer– thickness (S − R) increases with time. Interestingly, it is
solvent mixture [30–32,34]. Aforementioned, the first also observed from experimental results that there is a
model of swelling moving fronts was developed by synchronization of the two fronts when (S − R)
Lee et al. [19] for a one-dimensional swellable poly- becomes independent of time as shown in Fig. 6a.
mer system without any loaded drug (two-component The front synchronization period takes place after the
system). swelling has stopped, in which both interfaces move
For the swelling-controlled system, Colombo et al. inward at approximately the same rate. However, the
[31] suggested from experimental studies a new mechanism of front synchronization phenomenon has
finding that the gel layer consists of two regions of not been explained in detail. It is possible that the
dissolved and undissolved drug gel layer thickness, in synchronization period for swelling-controlled matrix
which the distance of dissolved gel layer thickness is systems only occurs in several special cases, which is
the most important parameter that influences the drug probably dependent on the matrix composition,
release. This is observed in several swelling systems, geometry, and the fabrication technique. This is
i.e. diclofenac, diprophylline, and cimetidine in supported by experimental observations, where front
HPMC, PVA, and carboxymethyl–cellulose sodium synchronization is observed for several polymer
(CMC) polymer matrix. This is expected since matrices of PVA with three drugs of different solubility
dissolution process does not limit drug release from (diclofenac, diprophylline, and cimetidine) [32], but it
the glassy region in the undissolved region and, does not prevail for polymer matrices of HPMC with
therefore, it is able to provide any amount of drug the same loaded drugs [33]. In this model, υd,eq and
available at the interface R for restricted diffusion in drug mass transfer coefficient at interface S (kS) are
the dissolved rubbery region. Hence, the diffusion in the fitted to the experimental fractional drug release curve.
dissolved rubbery layer is the controlling step for the The fitting to experimental data is obtained for the
drug release process and its distance becomes important release of sodium diclofenac, diprophylline, and ci-
to determine the drug release. This case is also true metidine-HCl from PVA–mannitol tablets. Fig. 6b
when the undissolved region is barely observed as the shows the cumulative release profile of sodium
dissolution at the interface R is relatively fast. Here, the diclofenac from PVA–mannitol system. The applica-
gel layer thickness becomes the important distance for tion of kS in this model may not be appropriate since
drug release. drug release at the interface is essentially due to the
1286 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

Fig. 6. The time variation of (a) normalized gel layer thickness, δgel = (S − R) / l, where l is the half of slab thickness or the initial position of both
interfaces R and S, with the line referring to the model prediction; and (b) cumulative fraction of sodium diclofenac release. The tablet consists of
30% PVA, 20% mannitol, and 50% sodium diclofenac. The release is observed at 37 °C in intestinal simulating fluid at 100 rpm. The line refers to
the dissolution model proposed by Harland et al. (reprinted from [32] with kind permission from Springer Science and Business Media).

polymer dissolution and controlled by drug diffusion Here, at the interface R, the polymer and drug volume
through the gel layer, but it is seldom limited by drug fractions are given by the free-volume theory [35–37],
mass transfer at the interface. whereas, at the interface S, the constant volume frac-
Narasimhan and Peppas [34] also used Lee's tions of polymer and drug are given by the Flory–
moving front concept to improve Harland's model Rehner equation [38].
for the swelling-controlled drug device. Similar to the The moving-boundary condition at the interface R
previous model, three components of water (compo- is defined by applying the following mass balance
nent 1), polymer (component 2), and drug (component equation:
d) are taken into account. The water and drug species 
⁎ ⁎ dR ∂t1 ∂td
balances are formulated in volume-fraction form based ðt1 þ td Þ ¼ − D1 þ Dd ð37Þ
on Fick's second law with initial conditions (t = 0) dt ∂x ∂x
υ1 = 0 and υd = υd,0. The boundary conditions are im-
The interface S initially swells outward due to polymer
posed as previously, where υ2 = υ*2 and υd = υ*d at
chain disentanglement and eventually moves inward
interface R and υ1 = υ1,eq and υd = υd,eq at interface S. In
when polymer disentanglement is limited as surface
this improved model, at interface S, the disentangle-
polymer concentration has reached the equilibrium
ment (dissolution) rate of the polymer chains (kd) is
state. The way of the movement of interface S is
used instead of the drug mass transfer coefficient (kS).
expressed as follows:
It is predicted to be the ratio of polymer gyration radius
(rg) to its reptation time (trept), as follows: dS ∂t1 ∂td
ðt1;eq þ td;eq Þ ¼ D1 t1;eq þ Dd td;eq
rg dt ∂x ∂x ð38Þ
kd ¼ ð36Þ ∂t2
trept −D2
∂x
The reptation time is the time when the solvent pene-
tration begins to disentangle the polymer chains until it The first two terms in the right-hand side account
is released into the surrounding medium. trept can be for polymer swelling due to water penetration and
determined by rheological experiments, whereas rg drug diffusion, respectively, whereas the third term
can be determined by light scattering experiments. accounts for polymer shrinking due to the limited
This model also improves the boundary conditions chain disentanglement. It is important to note that
estimation of volume fraction at the interfaces R and S. the mass balance at interfaces R and S (Eqs. (37)
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1287

and (38)) does not conserve each species separately. where 2l is the initial slab thickness and α = AB,
When densities of polymer and drug are not similar,
 
the flux of either one could be overestimated or t1;eq 1
A ¼ D1 ðt1;eq 1−t⁎ Þ þ
underestimated based on the density difference. t1;eq þ td;eq t⁎ ⁎
1 þ td
Here, it is assumed that the polymer chain transport  
td;eq 1
to dissolve at the interface S through a diffusion þDd ðtd;eq −t⁎ Þ þ ð43Þ
boundary layer of thickness δb is given by:
d
t1;eq þ td;eq t⁎1 þ td

 B¼
kd
∂t2 ∂ ∂t2 dS ∂td t1;eq þ td;eq
¼ D2 − ð39Þ
∂t ∂x ∂x dt ∂x
It is again implied that the model accounts for the
The initial and boundary conditions at the boundary superposition of Fickian diffusion and dissolution
layer interface for the thin boundary layer region are as process. The first term on the right-hand side of Eq.
follows: (42) is the diffusion-controlled release term, which has a
square-root dependency on time. In contrast, the second
t¼0 8x t2 ¼ 0
term is the dissolution term due to the chain disentan-
tN0 x ¼ SðtÞ þ db t2 ¼ 0
∂t2 glement, as expressed by a first-order dependency on
0btbtrept x ¼ S þ ðtÞ −D2 ¼0 ð40Þ time. The magnitude of each process is determined by
∂x
∂t2 the value of α, which is the dimensionless ratio of
tNtrept x ¼ S þ ðtÞ −D2 ¼ kd polymer dissolution at interface S to polymer and drug
∂x
þ
tNtrept x ¼ S þ ðtÞ t2 ¼ t2;eq diffusion in the rubbery region. As α approaches zero,
“Case-II transport” prevails and a linear release profile is
There are three stages to define the boundary condition observed. On the other hand, higher values of α indicate
at interface S in the thin boundary layer δb. When that the drug release is Fickian. In the former case, the
0 b t b trept, the disentanglement rate is negligible
 and no
 gel layer thickness (S − R) becomes relatively flat as drug
water is able to penetrate the interface S −D2 ∂t ∂x ¼ 0 .
2
is released in a zero-order manner due to the dissolution
The second stage is observed when t N trept as the rate of process at the interface S.
diffusion is sufficiently  high and the disentanglement The model is able to predict the normalized gel
rate is controlling −D2 ∂t ∂x
2
¼ k d . The last stage layer thickness variation with time for a tablet
indicates that the equilibrium state has been reached containing cimetidine hydrochloride (50%-wt), PVA
(υ2 = υ2,eq) when polymer is concentrated in the (10%-wt), and mannitol (40%-wt) as shown in Fig. 7a.
boundary layer since the polymer diffusion rate in the In addition, it can also explain the diprophylline
boundary layer becomes insufficient to transport the release from PVA–mannitol tablets (Fig. 7b). In the
chain to the surrounding medium. latter case, it is important to note that the release
Having imposed the pseudo-steady approximation prediction is independent of the experimental data (not
that results in linear concentration profiles of water and a fitting). More importantly, the synchronization
drug in the rubbery region, the variation of gel layer period is not observed, suggesting that it is likely to
thickness variation with time is also obtained mathe- be a special case. Even though this concept is
matically as follows: developed for one-dimensional system, it can later be
  adopted and extended to a model with any geometry of
ðS−RÞ interest.
−ðS−RÞ−aln 1− ¼ Bt ð41Þ
a Siepmann et al. [39] developed a model for drug
release from HPMC matrix by combining diffusion,
The fraction of drug release from the swelling-con- swelling, and dissolution mechanisms into Fujita-type,
trolled system is finally given by: exponential, concentration-dependent diffusivities of
pffiffiffi the solute and solvent. Drug and water diffusivities are
Mt Bðtd;eq þ t⁎

pffiffiffiffiffiffiffi pffiffiffi exponentially dependent on the concentration of the
¼ ð 2at þ BtÞ ð42Þ
Ml 2l swelling polymers due to their viscosity-inducing
1288 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

Fig. 7. The temporal variation of (a) normalized gel layer thickness, δgel = (S − R) / l, where l is the half of slab thickness or the initial position of
both interfaces R and S; and (b) cumulative mass release of diprophylline. The tablet consists of 50% diprophylline, 10% PVA
ðPMn ¼ 130; 000Þ, and 40% mannitol. The gel layer thickness and cumulative release profile are predicted by Eqs. (41) and (42), respectively
(reprinted from [34] with permission from Wiley-Liss, Inc., a subsidiary of John Wiley & Sons, Inc. Copyright © 1997).

capabilities. In addition, the transport analysis is im- the maximum concentration and diffusion coefficient
proved into two-dimensional (axial and radial) model of species i in the swelling polymer matrix, respec-
and integrated with the polymer swelling and disso- tively, at the interface between the matrix and release
lution. The decrease of dry polymer matrix mass is medium where the polymer disentanglement takes
characterized by the dissolution rate constant (kdiss), place. This model also accounts for the polymer
which is a function of the type of polymer and dissolution process quantitatively. The dissolution
dissolution medium. However, in current model, the process is characterized by a dissolution rate constant
polymer dissolution is not confined at the surface (kdiss) as follows:
boundary condition. Instead, the total polymer mass is
calculated over the release period to result in time- Mp;t ¼ Mp;0 −kdiss St t ð46Þ
variant matrix composition, matrix dimensions, and
drug and water diffusion pathways. where Mp,t and Mp,0 are the dry polymer matrix mass
The transport model applied in the cylindrical at time t and at initial time (t = 0), respectively, and St is
coordinate, without taking into account the variation in the surface area of the system at time t. It is important
θ-direction, is expressed as follows: to note that the polymer disentanglement (dissolution)
  constant is defined in different ways as the previous
∂Ci ∂ ∂Ci Di ∂Ci ∂ ∂Ci
¼ Di þ þ Di ð44Þ models. In the previous models, it is defined as the
∂t ∂r ∂r r ∂r ∂z ∂z polymer disentanglement rate constant (kd) in the
In the above equation, the concentration-dependent boundary condition at the interface with the surround-
diffusivity can be expressed as follows: ing medium.
In this model, the matrix swelling is assumed to be
  
Ci ideal, in which the total volume of the system is always
Di ¼ Di;crit exp −bi 1− ð45Þ equal to the sum of water, drug, and polymer volumes.
Ci;crit
The volume of each component is calculated at each
Here, i and βi indicate the diffusing species (i.e., 1 for time step, and by assuming ideal homogenous swelling
water and 2 for drug) and dimensionless constants and accounting for the polymer dissolution, the new
characterizing the concentration dependence of both volume is calculated. In this case, the system is
species diffusivities, respectively. Ci,crit and Di,crit are allowed to have volume increase (swelling) and
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1289

volume decrease (shrinking) due to imbibing water the concentrations of all species as well as influence the
into the system and polymer dissolution at the surface, mobility of the species (increasing diffusion coeffi-
respectively. However, the two moving fronts R and S cients of water and drug). Similar to the previous model
in the previous models are not observed in this model [39], only water and drug are considered in the
as the change in volume is only attributed to change of transport model, whereas the polymer undergoes a
outer boundaries. In addition, only two species (water reduction in molecular weight characterized by the
and drug) are considered to follow the Fick's second constant dissolution rate constant (kdiss). This model
law of diffusion as polymer dissolution process is has obvious improvements in accounting for the
simplified to be the overall mass decrease of the volume change in the system that affects concentration
polymer. Therefore, the polymer mass balance is of species in the system as well as the increasing
basically uncoupled from the drug and water transport. mobility of species due to water penetration. However,
Since the model is later used to explain drug release similar to the previous model, this model does not
from HPMC tablet, the model improvement in result in an explicit mathematical model and so it is
utilizing a two-dimensional cylindrical coordinate difficult to highlight the important swelling parameters
system is substantial. The model was able to explain that affect the drug release. In addition, the existence of
the propranolol-hydrochloride release profile from moving interface R for glassy region is neglected and
cylindrical HPMC matrices [39]. The agreement diffusion of drug and water is assumed throughout the
between model and experimental data is achieved by whole system. But, apparently the diffusion in the
simultaneous fitting of no more than two parameters, glassy core region can be negligible and the drug
e.g., βi and Di,crit. The study also shows that the effect release is dominated by the enhanced diffusion through
of the initial tablet radius is more influential on drug the gel layer.
release rather than that of initial thickness. It is The “sequential layer” model is able to fit the
expected since the former dictates more on the several experimental drug release profiles from HPMC
available surface area in the radial direction, which is tablets [40,41]. The model drugs used are chlorphe-
the main diffusional pathline for the drug release. From niramine maleate, theophylline, acetaminophen, di-
a practical view, however, this model relies on colofenac sodium, and propranolol-hydrochloride.
computational power and does not offer an explicit The model has also proved its capability to predict
mathematical solution. the unknown release profile in an independent experi-
By also utilizing Fick's second law in cylindrical ment for theophylline release from HPMC tablets
coordinates and the Fujita-type exponential-depen- (Fig. 8b). In addition, the applicability of this model to
dence of the diffusion coefficients for both water and fitting several model drugs from HPMC tablets with
drug, Siepmann et al. [40,41] further developed the different molecular weights suggests that this model
previous model into a “sequential layer” model can be used to understand the effect of HPMC tablet
(Fig. 8a), in which the tablet system is assumed to design (e.g., initial radius, height, and size) on the drug
have a certain amount of single layers penetrated by release profile [41,42]. In concordance with the
water. Introducing the concept of “sequential layer”, previous model's conclusion [39], simulation results
this model is performed in a computational grid and the also suggest that the available surface area of the
modified structure of the grid is required for numerical device for diffusion is the most crucial parameter that
analysis. An advantage of using computational grid is determines the drug release profile, in which varying
that it allows the modeling of inhomogeneous swelling. device geometry does not provide a straightforward
In contrast to the previous model, swelling is result. The model can be improved by coupling the
considered to take place layer by layer, in which the polymer species with drug and water species in the
outermost layer swells first followed by the neighbor- transport equation as has already been done before
ing inner layers. Here, the model is able to capture the [32,34]. It would be interesting to see in the future, the
major feature of swelling-controlled system, which is development of an additional feature of glassy core
the substantial change in volume of the system in the region in the “sequential layer” model that may enable
outer layer. This improvement is important since the the observation of the variation of two moving fronts
substantial change in the system volume will change during the drug release process.
1290 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

4. Mathematical models for erosion-controlled mer backbone cleavage and autocatalytic process, are
systems involved in the model.
Since complex phenomena take place in erosion-
Bioerodible polymers are versatile materials for a controlled systems, physiochemical characterizations
variety of biomedical applications, especially for drug on drug devices are important to understand the
delivery systems, since their chemistry and surfaces possible essential mechanism of drug release. The
can be tailored to stabilize macromolecular agents and identification of dominant drug release mechanism, in
enhance the tissue site-targeting. More importantly, the turn becomes the basis of mathematical model devel-
erosion kinetics can be tailored by careful selection of opment. In this particular objective, transient character-
polymer and a variety of techniques of encapsulation izations are often necessary to conduct during drug
to control the drug release profile. In the simplest release and erosion processes. Several characterization
manner, the erosion kinetics can be altered by techniques, commonly carried out in vitro, have been
modifying copolymer composition or the degree of used widely to investigate the controlled release
crystallinity as crystalline and amorphous polymers properties of polymeric drug devices as reported in
erode at different rates. literatures [43,44]. Gel permeation chromatography
It is important to distinguish the two different terms (GPC) can be used to monitor the polymer molecular
commonly used in describing the polymer erosion weight changes during the drug release and erosion.
phenomena, namely degradation and erosion. In a The cumulative polymer erosion and degradation, i.e.
simplistic point of view, degradation refers to the the monomer release, can be related to cumulative drug
polymer chain/bond cleavage/scission reaction (chem- release profiles to provide insights on how polymer
ical process), whereas the erosion designates the loss erosion and degradation are involved in drug release.
of polymer material in either monomers or oligomers Several mechanistic models relate the increase of drug
(chemical and physical process). Here, the erosion diffusion coefficient with the decrease of polymer
may consist of several chemical and physical steps, molecular weight due to chemical degradation which
including degradation. Since erosion is a more general will be discussed later. Differential scanning calorim-
term to capture the overall mechanism of the bio- etry (DSC) can measure the degree of crystallinity and
erodible system, this section will mainly utilize the glass temperature (Tg) changes. Aforementioned in the
term erosion. But the term of degradation will still be swelling-controlled system section, Tg is important in
used when specific degradation processes, e.g. poly- polymeric system since above Tg, the polymer is in

Fig. 8. (a) A schematic of “sequential layer” structure of the matrix in symmetry planes in axial and radial directions; and (b) the validation of the
model by independent experiment, in which the model predicts the result before the experimental study. The theoretical prediction shows an
agreement with the experimental cumulative release profile of theophylline from HPMC tablets with R0 = 0.6 cm, Z0 = 0.23 cm and 40% initial
drug loading. Samples were prepared as 500-mg tablets in phosphate buffer pH 7.4 medium at 37 °C (reprinted from [40] with kind permission
from Springer Science and Business Media).
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1291

rubbery state characterized by high mobility of polymer ical characteristic of these polymers has been re-
chains such that more free volume is available for drug viewed in literature [44]. On the other hand, polymer
diffusion. In contrast, below Tg, the polymer is in glassy with less reactive functional group, e.g. polyesters,
state, where the drug diffusion is more difficult as tends to degrade more slowly compared to solvent
compared to that in the rubbery region. Scanning penetration rate and thus bulk erosion is expected. The
electron microscopy (SEM) allows investigation of the most important of these esters for drug delivery
microstructure of polymer surface and matrix. If the application are poly(lactic acid) (PLA), poly(lactic-
cross-section cutting of the microspheres is possible, the co-glycolic acid) (PLGA), and poly(ε-caprolactone)
evolving internal morphology during the release can (PCL).
also be investigated using this technique. The characteristic time for water diffusion into the
There are two ideal scenarios for polymer erosion, polymer bulk (tdiff) is as follows:
namely surface (heterogeneous) and bulk (homoge-
neous) erosions (as shown in Fig. 9). In bulk erosion, phxi2 1
tdiff ¼ ð47Þ
the microsphere has a constant diameter size and 4 De
external fluid is allowed to penetrate into the micro-
where De is the effective diffusivity of water inside the
sphere, during which erosion of the polymer occurs. On
polymer matrix and 〈x〉 the mean distance for the
the other hand, in surface erosion, the microsphere has
diffusion. The effective Fickian diffusion is based on
an evolving shrinking diameter as the erosion of the
the continuum formulation with effective diffusion
polymer takes place at the external matrix boundary.
coefficient in liquid within fine pores [46]. The effective
It is suggested that the manner in which polymer
diffusion coefficient is defined as the result of hindered
erodes depends on the erosion number (ψ), which
diffusion that may be complicated by the size, shape,
is the ratio of characteristic time of diffusion of water
and electrical charge of the solutes and pores as
into the polymer drug to that of the degradation rate
described in [47].The characteristic degradation time
of the polymer backbone [45]. This study also implies
E(tn), for polymer functional groups is given by:
that the types of erosion behavior depend on the types of
polymer that strongly dictate the degradation rate of the "  P 1=3 #
polymer backbone. Therefore, polymer with reactive 1 1 Mn
Eðtn Þ ¼ lnðnÞ ¼ lnhxi−ln
functional groups, e.g. polyanhydrides, is expected k k NA ðNp −1Þq
to degrade faster and surface erosion is expected. ð48Þ
The most important of these polyanhydrides are poly
(1,3-bis-p-carboxyphenoxypropane-co-sebacic acid) where the pseudo first-order rate constant λ, is the
(p(CPP-SA)) and poly(1,6-bis-p-carboxyphenoxyhex- kinetic rate constant that accounts for the reactivity of
ane-co-sebacic acid) (p(CPH-SA)). The physiochem- polymer functional groups and therefore related to the
P
half-life of polymer bonds, Mn is the number-average
polymer molecular weight, Np is the degree of
polymerization, NA is Avogrado's number, and ρ is
the density of the polymer. Thus, the erosion number
(ψ) is defined as follows:

tdiff phxi2 k
w¼ ¼   P 1=3  ð49Þ
Eðtn Þ Mn
4De lnhxi−ln NA ðNp −1Þq

If the polymer matrix characteristics and type are the


same, the expression for ψ can be simplified as follows:
Fig. 9. Schematic illustration of surface- and bulk-erosion processes  
tdiff k k
in the polymer matrix (reprinted from [45] with permission from w¼ ¼K ¼ LðhxiÞ ð50Þ
Elsevier). Eðtn Þ De De
1292 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

This analysis suggests that polymer erosion depends on the latter considers the erosion process as a random
the matrix geometry, i.e., shape and size, since the event.
characteristic length (L) contributes to the magnitude of
ε. If λ and De changes are negligible during the erosion 4.1. Empirical models
process when erosion medium and fabricated polymer
matrix characteristics (e.g. porosity and type) can be Weibull [48] provided an empirical equation that
kept constant, the characteristic length (L) will be the was adopted for the drug release profile by Langen-
only parameter that determines how polymer erodes. In bucher [49] and applied to fit the drug release by
this case, there will be a critical characteristic length Vudathala and Rogers [50] and Dokoumetzidis et al.
(Lcrit) that defines the boundary between two possible [51]. The Weibull equation defines the cumulative
erosion types for each polymer type. When L N Lcrit, fractional drug release expression as a function of time
polymer will undergo surface erosion, whereas, when as follows:
L b Lcrit, polymer will undergo bulk erosion. This analy-
sis predicts that fast-eroding polymers, like polyanhy- " #
Mt −ðt−tlag Þb
drides, are already surface-eroding at micrometer scale ¼ 1−exp ð51Þ
as Lcrit ≈ 10− 4 , which is in agreement with the Ml tscale
observation from experimental studies [44]. On the
other hand, polymers like PLA and PLGA, which have In Weibull equation, tlag is the lag time before the drug
relatively less reactive functional groups, tend to have release takes place, tscale is the time scale of the release
bulk-eroding characteristics since they have higher process, and b characterizes the shape of the release
Lcrit (≈ 10− 2 ). curve. The case of b = 1 gives an exponential curve,
Several mathematical models have been developed to b N 1 gives a sigmoid (S-shaped) curve, and b b 1
explain the drug release from the erosion-controlled gives a parabolic curve with high initial slope followed
systems. Theoretically, the erosion mechanism is a com- by an exponential decay.
bination of mass transport and chemical reaction Hopfenberg [52] developed an empirical drug
phenomena. It involves several important mechanisms, release model for erosion-controlled polymer by as-
which are drug dissolution, polymer degradation, suming that the overall release behaves as a zero-order
porosity creation, micro-environmental pH change due process. This zero-order process is essentially a combi-
to polymer degradation, diffusion of drug in polymer nation result of dissolution and erosion processes at the
matrix, and autocatalytic effect during polymer degra- polymer surface. Therefore, this empirical equation is
dation. This complex interplay hinders the development appropriately applied for the surface-eroding particles
of a useful and accurate mathematical model that is able since this model assumes that the release rate is
to predict all the mechanism contributions on the controlled by the dissolution process on the surface.
resulting drug release kinetics from a bioerodible This model was also suggested by Hixson and Crowell
polymer. In the simplest manner, mathematical models [53] with the assumption that the shrinking spherical
for erosion-controlled systems are classified into empir- particle area is proportional of the cubic root of its
ical models and mechanistic models. The former does volume. For spherical geometry, the mathematical
not take into account the complex physicochemical equation is written as follows:
phenomena and are commonly developed for surface-
eroding systems that exhibit zero-order release process.  
Mt kero;0 t 3
On the other hand, the latter accounts for the physico- ¼ 1− 1− ð52Þ
chemical phenomena that basically involve diffusional Ml C0 R
mass transfer and chemical reaction processes. Based
on the approach of developing models, the latter can where kero,0 is the surface erosion rate constant, C0 is
be sub-classified into diffusion-and-reaction models the initial concentration of the drug in the matrix, and
and cellular-automata models. The former models the R is the initial radius of the sphere.
erosion process as a transport process of combined El-Arini and Leuenberger [54] modified the Hopfen-
diffusion and chemical reaction processes, whereas berg model by accounting for the lag time (tlag) in the
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1293

early time of drug release process. The modified drug 4.2. Mechanistic models
release fraction equation is written as follows:
 3 Mechanistic models are primarily divided into two
Mt kero;0 main groups based on how the erosion process is
¼ 1− 1− ðt−tlag Þ ð53Þ
Ml C0 R modeled. The first group is the diffusion-and-reaction
model, which suggests the description of the erosion
Cooney [55] considered two main steps in the process as a combination of polymer diffusion and
polymer matrix that undergo surface erosion, which reaction. Until now, these models are primarily deve-
are the detachment of an atom or molecule, followed loped for the bulk-eroding microspheres with exception
by the diffusion of this agent through a diffusion of several modeling efforts by Zhang et al. [56] and
boundary layer into the bulk. For the case of surface Larobina et al. [57]. The second group, which is the
erosion (slow dissolution rate) in a sphere, the surface cellular-automata model, assumes the erosion process as
dissolution/detachment controls the overall process a random event. Here, the matrix surface detachment is
and the mass balance is given by: commonly assumed the rate-controlling step, therefore
  the models are best applied for surface-eroding systems.
d 43 pr3 For this reason, contrary to the diffusion-and-reaction
−qs ¼ kero ð4pr2 ÞDC ð54Þ
dt models, the cellular-automata models are primarily built
to explain the surface erosion process, except for a
With the assumption of constant concentration differ- model by Siepmann et al. [58] that utilized the Monte-
ence existing between the surface and the bulk me- Carlo technique for a bulk-eroding system.
dium (ΔC), the previous equation has a solution as
follows: 4.2.1. Diffusion-and-reaction models
  Harland et al. [59] developed the first dissolution
kero DC r kero DC
r ¼ R− t or ¼ 1− t ¼ ð1−sÞ model for a polymer matrix that undergoes bulk ero-
qs R qs R sion. The model is solved for both infinite and finite
ð55Þ mass transfer boundary conditions at the polymer sur-
face. The model accounts for the effective Fickian
However, even though the surface concentration can diffusion and dissolution of the solute into liquid-filled
be maintained, the surface area will decrease with time pores. Aforementioned, the effective Fickian diffusion
leading to a lower dissolution rate. The ratio of the is based on the continuum formulation by applying the
instantaneous dissolution rate to the initial dissolution effective diffusivity to account the diffusion in liquid
rate ( f ) and the total drug release rate can be defined within fine pores [46]. Thus, the transport model of the
as follows: drug solute is expressed as follows:
 2
∂C ∂ C 2 ∂C
f ¼ ð1−sÞ2 ð56Þ ¼ De þ þ kðeCs −CÞ ð58Þ
∂t ∂r2 r ∂r
If the surface-eroding matrix has a uniform drug Here, C and De are the drug concentration and effective
distribution, the cumulative drug release expression is diffusivity in liquid-filled pores respectively, k is the drug
similar to that of Hopfenberg, which is as follows: dissolution rate constant, ε is the porosity of the polymer
 matrix and εCs is the equivalent drug saturation
Mt  r 3 kero DC 3 concentration in the solution found in the pores. The
¼ 1− ¼ 1− 1− t ð57Þ
Ml R qs R second term in the right-hand side is the dissolution term
and is negligible when the drug loading (C0) is smaller
The final result is basically similar to that of Hopfenberg. than the solubility concentration (εCs). In case of a non-
The only slight difference between these two models is constant diffusivity, the diffusion coefficient is assumed to
that Hopfenberg's model does not account explicitly the be exponentially concentration-dependent and the bound-
contribution of concentration gradient, which is lumped ary condition at the surface is the infinite convective mass
to be a constant (kero,0). transfer under a perfect sink medium.
1294 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

Performing dimensionless analysis by scaling This relationship takes places as long as undissolved
naturally so that all variables are bounded between 0 drug in the microparticles still exists.
and 1, a new dimensionless number, the dissolution/ Heller and Baker [60] used the classical Higuchi
diffusion number (Di), is defined as follows: model as a basis to develop a mathematical model for
bulk-eroding polymers. The main process that controls
kR2 drug release is assumed to be the hydrolytic backbone
Di ¼ /2s ¼ ð59Þ
De cleavage before the drug is solubilized in the medium.
The modified Higuchi model is as follows:
Following a chemical engineering point of view, Di is
the square of Thiele modulus (ϕs) since similar 
phenomenon is observed when reaction and diffusion dMt S 2PC0 1=2
¼ ð63Þ
take place altogether inside a catalyst. dt 2 t
The solution of fractional drug release for infinite
mass transfer at the surface is as follows: In this equation, S is the surface area on both sides of
the planar film, C0 is the initial drug concentration
Mt (loading) in the polymer, and P is the permeability of
eCs 43 pR3 the drug inside the polymer matrix.
Xl
ðDi þ n2 p2 ÞDis þ n2 p2 f1−exp½−ðDi þ n2 p2 Þsg
¼6 Contrary to the classical Higuchi model, this model
n¼1 ðDi þ n2 p2 Þ2 accounts the effect of the drug permeability in the
ð60Þ biodegradable polymer, which is not constant, but
increases with time, as more pores are created during
where the dimensionless time (τ) is defined as τ =Det /R2. the erosion. The following equation was used to
On the other hand, for the release under a finite account for the change of permeability with time as a
convective mass transfer condition, the Sherwood num- function of number of remaining bonds.
ber (Sh), which is the ratio of the rate of convective
mass transfer to that of diffusion, appears in the P initial number of bonds N
solution. The solution is as follows: ¼ ¼ ð64Þ
P0 remaining number of bonds N −Z
Mt
eCs 43 pR3 where N is the initial number of bonds and Z is the
Xl
ðDi þ a2n R2 ÞDis−a2n R2 fexp½−ðDi þ a2n R2 Þs−1g number of bonds that have undergone cleavage.The
¼ 6Sh2
n¼1 ðDi þ a2n R2 Þ2 ½a2n R2 þ ShðSh−1Þ bond cleavage process is assumed to follow a first-
order kinetics that can be expressed as follows:
ð61Þ

where Sh = hR / De and α n are the roots of the dZ


¼ KðN −ZÞ ð65Þ
transcendental equation of αnRcot(αnR) + Sh − 1 = 0. dt
Interestingly, the simplified solution for longer times
of drug release, during which dissolution controls the where K is the first-order kinetic rate constant for bond
process, reveals that that the amount of drug release is cleavage process.The final equation for the rate of drug
simply a first-order function of time as follows: released from the system is as follows:

"rffiffiffiffiffiffi sffiffiffiffiffiffi! 
2pR3 eCs De k dMt S 2P0 expðKtÞC0 1=2
Mt ¼ pffiffiffi coth R ¼ ð66Þ
k R D e
dt 2 t
sffiffiffiffiffiffi!#
pffiffiffi 2 k
− k csc h R þ 8pRDe eCs ð62Þ Thombre and Himmelstein [61] developed a math-
De ematical model for a bulk-eroding polymer matrix of
" rffiffiffiffiffiffi sffiffiffiffiffiffi! #
poly(orthoester) (Fig. 10). The model is based on slab
kR De k 1
 coth R − t geometry and a perfect sink condition is assumed. In
2De k De 2
this model, the acid generation from acid hydrolytic
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1295

Fig. 10. (a) Schematic illustration of the diffusion-reaction model for poly (orthoester) drug delivery system by Thombre and Himmelstein; and
(b) fractional release of each species as a function of time based on the proposed model (reprinted from [61] with permission from Wiley-Liss,
Inc., a subsidiary of John Wiley & Sons, Inc. Copyright © 1985).

reaction is taken into account. The resulting species dence of drug diffusivity in the polymer matrix to
mass balance is as follows: account for increase in drug release during erosion.
The advantage of these models is the variation of

∂Ci ∂ ∂Ci polymer molecular weight with time due to erosion and
¼ De;i ðx; tÞ þ ti ð67Þ the dependence of drug diffusivity on polymer molec-
∂t ∂x ∂x
ular weight can be independently determined and
therefore, the number of fitted parameters is mini-
In this equation, υi is the net sum of synthesis and
mized. Here, however, it is important to note that the
degradation rate of species i, whereas A, B, C, and E
pore creation that enhances the drug diffusion is not
denote water, acid generator, acid, and drug, respec-
mechanistically modeled.
tively.The diffusion coefficient of each species is also
Charlier et al. [62] assumed that the polymer degra-
updated during the erosion as an exponential function of
dation, which is quantitatively regarded as the polymer
the respective species concentration as follows:
molecular weight, follows a first-order kinetic process
  of polymer chain cleavage. However, the model is
ðCi;0 −Ci Þ developed to quantify drug released from PLGA film
De;i ¼ Di;0 exp l ð68Þ
Ci;0 with planar geometry.

where Di,0 and Ci,0 are initial diffusion coefficient and dMw
concentration of component i when the polymer is yet ¼ −kdegr Mw YMw ¼ Mw;0 expð−kdegr tÞ ð69Þ
dt
hydrolyzed, and μ is a constant.
Several models prefer to apply the simple Fick's The drug diffusion coefficient in polymer matrix during
second law, but define the molecular weight depen- erosion is assumed to correlate inversely to the polymer
1296 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

molecular weight, which suggests the following drug release is obtained from a combination between
equation: Baker and Lonsdale's diffusion solution (Eqs. (9) and
(10)) and the pure matrix erosion (FE) of Fitzgerald
De Mw;0 and Corrigan [65]. For drug release from a sphere, this
¼ YDe ¼ D0 expðkdegr tÞ ð70Þ is expressed as follows:
D0 Mw
rffiffiffiffiffiffiffi  
This model assumes a linear concentration gradient as Mt De t De t expðkdegr ðt−tmax ÞÞ
¼4 −3 þ FE
predicted in the classical Higuchi model. The final Ml pr2 r2 1 þ expðkdegr ðt−tmax ÞÞ
expression for the rate of drug release is as follows: ð74Þ

2C0 Cs D0 ðekdegr t −1Þ where tmax is the time to maximum matrix erosion rate.
Mt ¼ S ð71Þ
kdegr This model is able to describe the triphasic drug
release profile that comprises: (i) initial “burst” due to
At early times, the resulting equation can be reduced the short diffusion pathways, (ii) intermediate phase of
into Higuchi's solution for C0 N Cs as ekdegr t ≈ 1 + kdegr t, zero-order release due to simultaneous drug diffusion
whereas, at longer times, the exponential term will and polymer degradation, and (iii) second rapid drug
control the drug release and Eq. (71) still holds. The release due to final matrix erosion that releases the
solution for the former case is expressed mathematically remaining drug. The resulting release profile is similar
as follows: to the “S” erosion model for bulk erosion by Zhang
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi et al. [56] that will be described later in this section.
Mt ¼ S 2C0 Cs D0 t ð72Þ A similar approach was also used by Siepmann
et al. [66] to explain the 5-fluorouracil (5-FU) release
Raman et al. [63] used the simple diffusion model from bulk-erosion PLGA microspheres. However, this
for spherical geometry with diffusivity dependence on model requires two different constants to express the
molecular weight to explain the piroxicam release evolving drug diffusivity as the diffusivity is not a
from bulk-erosion poly(lactide-co-glycolide) (PLG) simple ratio relationship as in Charlier et al. [62]. A
microspheres. The molecular weight loss due to the similar approach was previously done by Wada et al.
erosion is expressed as follows: [67] for aclarubicin hydrochloride (ACR) release
from PLA-based microspheres. The variations of
Mw;0 when tbtlag drug diffusivity with molecular weight and molecular
Mw ¼ ð73Þ
Mw;0 ekdegr ðt−tlag Þ when tztlag weight with time are expressed as follows:

Here, tlag is the lag time before the polymer matrix k


De ¼ D0 þ ð75Þ
erodes. The diffusivity dependence on molecular Mw
weight is estimated to be an empirical cubic polyno-
mial fit that relates ln(De) and ln(Mw). Another impor-
Mw ðtÞ ¼ Mw;0 expðkdegr tÞ ð76Þ
tant feature of the model is the effect of non-uniform
drug distribution using data from confocal fluores-
cence microscopy. Here, intensity plots obtained from In addition, the difference with previous models is that
confocal microscopy are converted to initial radial the amount of drug release utilized the approximate
distribution of drug. explicit solution by Koizumi and Panomsuk [7] for
He et al. [64] also used a similar approach by spherical delivery system. The amount of drug
combining Baker and Lonsdale's model [4] for released equation is expressed as follows:
diffusion-controlled release and the concept of in- 
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 4Cs De t
creased diffusion coefficient due to polymer chain Mt ¼ 4pR2 2ðC0 −Cs ÞCs De t þ
9R
scissions. In their model, similar to Charlier's   ð77Þ
approach, the variation of De with time is defined as Cs
 −3
an exponential function. However, the overall fraction 2C0 −Cs
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1297

Ehtezazi and Washington [68] combined the per- to contact with water. The resulting acids will diffuse
colation theory and classical Crank diffusion model to out of the microparticles to get neutralized. On the other
study the macromolecule (FITC-dextran) release from hand, bases also diffuse into the microparticles to
porous PLA microspheres. Here, the kinetic model is neutralize the acids. However, since both diffusional
basically lumped in the evolving drug diffusion processes are comparably slower than the rate of acids
coefficient through the percolation theory. In this generation, the micro-pH inside the system still drops
approach, the pores are divided into conducting during the bulk-erosion process. Here, the hypothesis is
(accessible) and discrete (isolated) regions. The that, with the increasing microparticles size, the diffu-
accessible pore region is connected to the exterior sion pathways for acids and bases increase and thereby,
surface that allows mass transport from the system to the neutralization process is less pronounced. This will
the surrounding. On the other hand, the isolated pore lead to a drop in micro-pH inside the matrix and the
space is disconnected from the surrounding medium. It accelerated polymer degradation. When polymer deg-
is assumed that, below the percolation threshold (ρC), radation is assumed to follow first-order kinetics, the
the accessible porosity is not available and, thus, the experimental results are fitted into the equation to
porosity equals to the isolated porosity. The Bethe obtain the polymer degradation rate constant (kdegr).
lattice theory relates the percolation threshold with the Having assumed the drug release as a diffusion-
coordination number z, which is given by: controlled process with effective diffusivity (Eq. (58)
without the dissolution term in the right-hand side of the
1
qC ¼ ð78Þ equation) into a perfect sink medium with a finite
ðz−1Þ convective mass transfer coefficient (h), the fitting
Having determined the Bethe lattice coordination shows that the effective diffusivity (De) increased sub-
number for molecule of which the diameter is smaller stantially with increasing microsphere diameter. This is
than the porous diameter, the effective diffusion modeled by defining the effective diffusivity as a
coefficient is related to the transport coefficient of function of microsphere diameter. This observation
the porous structure (εE) that is determined as follows: suggests that the simulation results without autocatalytic
effect fail to predict the lidocaine release profile (Fig.
De ¼ D0 eE ð79Þ 11). It is explained by the autocatalytic effect that causes
the increasing polymer degradation rate with increasing
For a Bethe lattice model, εE can be calculated as microsphere diameter so that it allows higher mobility of
follows: the drug to be released into the medium. This is con-
 firmed by SEM study that shows the smaller micro-
z−1 C Vð0Þ
eE ¼ − ð80Þ spheres were more resistant to degradation than larger
z−2 D0
microspheres [69] as shown in Fig. 12. In this case,
where C′(x) is the derivative of a non-linear integral using the classical diffusion model, it is initially assumed
equation defined in Ehtezazi and Washington [68]. that the dissolution of the drug in this bulk-eroding
After determination of De, the fraction of drug system is negligible, whereas the autocatalytic effect is
released is calculated using the classical Crank model more significant. Though it is often acceptable that for a
for diffusion in microspheres. Furthermore, in order to hydrophobic agent, like lidocaine, since the dissolution
take into account the non-uniform distribution of contribution is marginal, the fitting could have been
microsphere sizes, the equation is slightly modified by done better by assuming both dissolution and diffusion
introducing the fraction of microspheres with certain processes take place together using a Harland model
size ranges. [59]. If the magnitude of the dissolution term turns out to
Siepmann et al. [69] provided a hypothesis to be negligible, it can be confirmed that the drug was
incorporate the autocatalytic effect to explain the lido- released dominantly by the diffusion process. Never-
caine release from PLGA microspheres. The PLGA theless, this finding unveils an interesting discussion on
polymer is known to undergo the bulk-erosion process, how important the size effect on the drug release rate.
where its polyester chains are cleaved into shorter Batycky et al. [70] developed a model to predict the
chains of alcohols and acids throughout the matrix due water-soluble glycoprotein released from bulk-eroding
1298 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

Fig. 11. (a) Theoretical lidocaine release profiles for different sizes of PLGA microspheres without autocatalytic effects (De = 4.6 × 10− 14 cm2/s);
and (b) experimental (symbols) and theoretical (solid lines, with autocatalytic effects) lidocaine release profiles for different sizes of PLGA
microspheres in phosphate buffer pH 7.4 (reprinted with permission from [69]. Copyright 2005 American Chemical Society).

system of PLGA and PLA by taking into account the burst (Fig. 13). Since the protein is water-soluble and
evolving populations of micro- and mesospores and has a relatively large size, the model incorporates the
the possible fractional drug release due to the initial induction phase that takes place when the available

Fig. 12. SEM pictures (surfaces and cross-sections) of small (R = 8 μm) and large (R = 80 μm) lidocaine-loaded PLGA microspheres in phosphate
buffer pH 7.4 at day 7 (reprinted with permission from [69]. Copyright 2005 American Chemical Society).
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1299

Fig. 13. (a) Schematic illustration of a microsphere that undergoes hydration, degradation, and erosion in bulk-erosion system; and (b)
experimental (symbols) and model fitting (solid lines) of fractional release profile of MN rgp 120 protein (reprinted from [70] with permission
from Wiley-Liss, Inc., a subsidiary of John Wiley & Sons, Inc. Copyright © 1997).

initial porosity is low, and thus prevents release of the is assumed to be constant, which may not be true as the
protein via diffusion until mesoscopic pores are avail- bulk erosion increases the porosity and therefore, the
able in sufficient quantity. This also suggests that there drug diffusivity should increase during the erosion. The
is an initial burst before the induction time (td) due to model also does not account for the effects of variable
the water-soluble protein because of the protein de- pH on the degradation constants as the acidic monomers
sorption in mesospores connected to external surface are produced during the erosion. More importantly, the
of the microsphere. The model involves an explicit fraction of drug release from the burst effect must be
evolving microstructural model of pores development obtained experimentally through the fitting of release
during the erosion. This is related to the mass and curve. Therefore the model can actually be considered
molecular weight change of the polymer that was de- as a semi-empirical model though the diffusion-and-
scribed by a combined model of random end-scission. reaction concept is still retained after the induction
The fractional protein release from the microspheres is period. However, this effort illustrates the difficulties of
expressed as follows: modeling the release of hydrophilic drug, i.e. protein
from hydrophobic polymer, i.e. PLA and PLGA, and
Mt justifies the use of a semi-empirical equation to fit the
¼ 1−b½expð−kdegr tÞ experimental data. The unique problem of modeling the
Ml  2 2 
−i p De ðt−td Þ ð81Þ hydrophilic drug released from nanospheres will also be
6X l exp
R
discussed later in Section 6.
−ð1−bÞ 2
p i¼1 i2 Zhang et al. [56] developed mathematical models for
bulk- and surface-eroding polymeric microspheres by
In this equation, b is the drug fraction release in the considering diffusion, dissolution, and erosion process-
initial burst. The last term will be negligible when t b td es altogether. The bulk-erosion models are solved for
as the drug is not available to diffuse yet. Following the both infinite and finite mass transfer cases, whereas the
induction time (t N td), the drug undergoes Fickian boundary movement in surface erosion model is as-
diffusion through the porous microspheres. sumed to be a first-order process with time. This study
This model has a good agreement with the experi- also includes the sensitivity analysis of several para-
mental data in terms of polymer mass loss fraction and meters that may affect the drug release. The drug release
protein release fraction. However, there are several profiles are fitted to several experimental data, even
limitations in this model. For instance, drug diffusivity though it is mainly performed for the bulk-erosion case.
1300 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

For the bulk-erosion model, the concept of effective amorphous part is blocked by the crystalline domain,
(V1) and virtual (V0) solid phases was proposed. The which makes the degradation rate slow at the beginning.
effective solid phase is the volume of solid that is When the crystalline region is hydrolyzed, the amor-
eroding, in which the change pattern is related to the phous polymer chains are free to degrade therefore both
type of erosion. Erosion reduces the part of the degradation and drug release rates are accelerated.
effective solid phase and causes drug release from For surface-erosion model, two compartments are
matrix entrapment. As a result, the drug concentration considered, namely the solid and pore (liquid) com-
in effective solid phase is reduced. The virtual solid partments. The latter refers to the available compart-
phase is the initial volume of microspheres before ment for water penetration. The size of the microsphere
erosion occurs. Therefore, three phases are taken into subsequently decreases, but the average polymer mole-
account in this model, which are the liquid phase that cular weight does not change appreciably. This suggests
has constant volume of V0, the virtual solid phase that that erosion is confined at the external boundary. The
has constant volume of V0, and the effective solid model assumed that the amount of water penetrating the
phase that has a variable volume of V1. At the initial microspheres is low so it does not cause erosion in the
time, the volume of the effective solid phase V1 is interior of microspheres. The dissolution from the solid
equal to that of the virtual solid phase V0. The concen- phase to the liquid phase is taken into account in the
tration in virtual solid phase is expressed as follows: Fick's second law for the liquid phase as proposed in
 Eq. (58) by Harland et al. [59]. Under an infinite
V1
CS ¼ CSE ð82Þ convective mass transfer condition, the initial and
V0
boundary conditions for the equation are given by:
where CS and CSE are drug concentrations in virtual
and effective solid phases, respectively. t ¼ 0 and 0 V r V R0 CL ¼ Cb
In the effective solid phase, concentration decreases ∂CL
t N 0 and r ¼ 0 ¼0 ð83Þ
only due to the dissolution process. When the liquid ∂r
phase concentration is higher than this concentration, t N 0 and r ¼ R0 ð1−kero tÞ CL ¼ Cb
the drug will deposit back to the effective solid phase. where CL is the drug concentration in the liquid phase
In addition, the diffusion coefficient varies with the and the constant rate of movement of erosion boundary
changes of polymer porosity and tortuosity. Here, the is given as kero.
tortuosity changes are proportional to the volume On the other hand, the governing equation and
change of effective solid phase. initial condition for solid phase that only accounts for
Three types of erosion are modeled, namely linear the dissolution term can be expressed as follows:
erosion, “S” erosion, and hyperbolic erosion. The pat-
tern the polymer matrix erodes depends on several ∂CS
¼ −kðeCsat −CÞ ð84Þ
physical properties, such as degree of crystallinity (the ∂t
amorphous region is more prone to degradation), t ¼ 0 and 0 V r V R0 CS ¼ C0 ð85Þ
molecular weight distribution (wide distribution would
accelerate the degradation rate), the portion of glycolide where R0 is the initial radius of the moving front of the
(more glycolide monomer increases the degradation eroding interface.
rate), and glass transition temperature (polymer chains The dissolution term is set to zero when no driving
in the rubbery state are more responsive to erosion with force is present (CL = Csat or CS = Cb), in which case
the presence of liquid water). For linear erosion, the ratio dissolution does not take place anymore. Csat is the drug
of amorphous region to crystalline region and free saturation concentration. This condition applies for both
chains to rigid chains are comparable. For hyperbolic governing equations of liquid and solid phases. These
erosion, most of the polymer is in the amorphous state equations are solved by the transformation technique to
and the polymer chains are loose. This causes fast convert the moving-boundary problem to a fixed-
degradation at the initial stage followed by slow boundary problem, which is computationally easier to
degradation at the later stage due to the presence of handle. In this surface erosion model, even though the
crystalline region. For “S” erosion, it is possible that the model is assumed to be a two-compartment model, the
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1301

solid and liquid compartment equations are actually 4.2.2. Cellular-automata models
uncoupled. In the liquid compartment, the additional Zygourakis [72] first developed a cellular-automata
drug concentration is attributed to the dissolution term model to simulate the drug release from surface-eroding
and also washed away to the surface due to its con- polymer matrices. The model used a two-dimensional
centration gradient. grid. Each pixel in the grids represents drug, polymer,
Larobina et al. [57] developed a new model for filler, or pore. The erosion process relies on the “life
surface-eroding polymer system based on the polymer expectancy” that is defined for each type of pixel. The
phase separation phenomenon observed by Shen et al. lifetime starts counting when the solvent is in contact
[71]. This phenomenon is observed in the system of with the polymer, which makes the pixels in surface
CPP–SA and CPH–SA when the fraction of one com- more prone to polymer erosion rather than any other
ponent is in excess of ∼80%-mol. In this system, SA is site. But, this model does not consider the mass trans-
the fast-eroding phase with rate constant kA, whereas port phenomena, such as the water and drug diffusion
CPP or CPH is the slow-eroding phase with rate inside the polymer matrix during erosion. The cross-
constant kB. In addition, tA and tB are the times sections of the matrix grid for 0, 25, 50, and 75% drug
required for both fast-eroding and slow-eroding phases released are shown in Fig. 14. The dissolution rates of
to degrade, respectively. The surface erosions for both the drug (Rd) and polymer (Rp) are defined as:
species are described by zero-order reactions. Since
the drug will be released when the polymer phase that
dVd kd Sd ðCd;s −Cd;b Þ kd Sd DCd
entraps it degrades, the drug partitioning into either of Rd u ¼ ¼ ð87Þ
these phases influences the drug release from this dt qd qd
polymer system. Due to the inability of the whole
polymer system to entrap the entire drug in compatible dVp kp Sp ðCp;s −Cp;b Þ kp Sp DCp
domains, in which the over-saturation condition Rp u ¼ ¼ ð88Þ
dt qp qp
occurs, this model also accounts for the burst effect.
The cumulative drug release from this model can be
expressed as follows: When the diffusion resistance is assumed to be negli-
gible, the drug release rate is equal to the drug disso-
lution rate. Therefore, the drug release rate in a constant
Mt ðxA ðtÞ/A;0 þ xBVðtÞ/B;0 PÞ volume medium (V0) is given as follows:
¼ ð1−bÞ þb ð86Þ
Ml /A;0 þ /B;0 P
dMt 1 dVd 1 kd Sd DCd
¼ ¼ ð89Þ
In this equation, xA and xB′ are the released mass dt V0 dt V0 qd
fractions of fast- and slow-eroding monomers, respec-
tively, ϕA,0 and ϕB,0 are the surface fractions of fast- These equations are solved using cellular automata and
and slow-eroding domain, respectively, b is the discrete iterations by updating the lifetime of each pixel
fraction of drug released in the burst, and P is the for each time point.
ratio of drug concentration in the slow-eroding phase Zygourakis and Markenscoff [73] improved the
to that in the fast-eroding phase. Here, kA is fitted to model by analyzing the effect of several structural
the monomer release data, whereas the other three parameters on the erosion and release rates. The
constants (kB, tA, and tB) are calculated after kA is parameters that are analyzed in this model are given
determined. Then, the experimental drug release data by:
is fitted to Eq. (86) to estimate P.
Here, the model has shown good agreements in 1. The ratio of dissolution rate constant (χ), which is
terms of drug release profile with experimental results defined as χ = ν1 / ν2.
[71]. However, P is a semi-empirical parameter that is 2. The initial loading of the first component (λ1),
difficult to obtain experimentally and may vary with which is defined as λ1 = V1,0 / (V1,0 + V2,0), where
time as erosion progresses with time. The fitting V1,0 and V2,0 are the initial volumes of components
should suggest whether the estimated P is reasonable. 1 and 2, respectively.
1302 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

Fig. 14. Zygourakis cellular-automata model [72] for surface-eroding system simulation. The pixel types are drug, polymer, filler, and pores.
There are four stages of surface erosion as shown, referring to the time points when the drug released is (a) 0%, (b) 25%, (c) 50%, and (d) 75%
(reprinted from [72] with permission from Elsevier).

3. The dispersion parameter (ξ), defined as the aspect is important as the crystalline region erodes slower
ratio ξ = d / h, where d is the characteristic length of than the amorphous region as observed experimen-
component 1 and h is the characteristic length of tally [75]. Similar to the Zygourakis model, the
the device. simulation is conducted in the two-dimensional grid.
4. The porosity of the device (ε). In the subsequent modeling development, this model
5. The pore size parameter (γ), whose length is was also used as the basis for the three-dimensional
defined as γ = l / h, where l is the pore size. erosion of polymer cylinder [76]. Each pixel is
assigned its initial lifetime. The initial lifetime
The first two parameters are already incorporated in expectancy for each pixel is assigned randomly
the previous model [72], whereas the remaining three according to Erlang distribution. To account the
parameters are new to account for the characteristic effect of crystallinity, the crystalline and amorphous
geometry of the device as well as the dynamics of pixels are assigned different Erlang distribution
erosion process that affects the porosity of the matrix. constants (e):
The simulation results show the effect of these param-
eters on the release profile of surface-erosion matrix
eðtÞ ¼ ke−kt ð90Þ
system.
Gopferich and Langer [74] utilized the Monte-
Carlo technique to build a mathematical model for where λ is the first-order degradation rate constant.
the polymer matrix erosion. The model takes into Pixels that are in contact with either water or the
account several parameters that are correlated with eroded neighbor pixels will start to erode and its
how the polymer erodes, such as crystallinity, lifetime is calculated from that point. The pixel will be
geometry, and porosity. The degree of crystallinity removed from the grid when its lifetime ends. The
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1303

Fig. 15. (a) Schematic illustration of surface polymer degradation with monomer release mechanism incorporated; and (b) the model fitting to
experimental data of CPP and SA monomers released from p(CPP-SA) 20:80 polymer matrix discs (reprinted from [78] with permission from
Elsevier).

matrix porosity (ε) is updated for each time step schematic illustration of this modeling and its fitting
according to the following equation: results of monomer release are shown in Fig. 15. This
block copolymer has been widely investigated to have
1X yn
the surface-eroding characteristic [77]. In such a
eðidDx; tÞ ¼ sði; jÞ for 1 V i V nx ð91Þ
ny j¼1 system, the diffusion of monomer in polymer matrix
is derived from Fick's second law of diffusion:
1 for ‘eroded’ pixels
where sði; jÞ ¼  
0 for ‘noneroded’ pixels. ∂ ∂ ∂Ci ðx; tÞ
½Ci ðx; tÞd eðx; tÞ ¼ De;i ðCi Þdeðx; tÞd
As diffusion takes place in the x-direction, i · Δx is ∂t ∂x ∂x
equal to x. In the above equation, ε(x,t) is the discrete ∂Si
porosity distribution; nx and ny are the numbers of þ
∂t
pixels in x and y direction in the grid, respectively; and ð92Þ
s(i, j) is a function that takes into account the status of a
pixel to be either ‘eroded’ or ‘non-eroded’. The poros- where Ci(x,t) is the concentration of the diffusing SA
ity dynamically increases as mentioned by the in- or CPP monomer, ε(x,t) is the porosity of each pixel,
creasing number of ‘eroded’ pixels. The matrix De,i(Ci) is the effective diffusion coefficient of either
porosity affects the magnitude of the effective monomer, and @S @t is the dissolution rate.
i

diffusion coefficient of the drug in the polymer matrix. The complete species balances for monomer SA
Since the monomer properties are found to be im- and CPP that include diffusion and dissolution
portant to the erosion mechanism of polyanhydrides- processes are as follows:
based polymer matrix [75,77], the previous model is
then developed in a more extensive detail by in- ∂ ∂CSA ðx; tÞ ∂De;SA ðCSA Þ
CSA ðx; tÞ ¼
corporating the monomer release [78]. Here, it turns ∂t ∂x ∂x
out that the drug release profile is more correlated to De;SA ðCSA Þ ∂CSA ðx; tÞ ∂eðx; tÞ
þ
the monomer release profile than the overall polymer eðx; tÞ ∂t ∂x
molecular weight decrease due to erosion. The model ∂2 CSA ðx; tÞ
þ De;SA ðCSA Þ
is originally applied for the SA-CPP copolymer that is ∂x2
widely used at that time for the Gliadel® wafer. The þ kSA mSA ðx; tÞ½CSA;s −CSA ðx; tÞ ð93Þ
1304 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

 
∂ ∂CCPP ðx; tÞ ∂De;CPP ðCCPP Þ phous polymer of type one  ¼ tky , κ1 and κ2 are
CCPP ðx; tÞ ¼
a1

∂t ∂x ∂x the dimensionless dissolution rates for polymer of


De;CPP ðCCPP Þ ∂CCPP ðx; tÞ ∂eðx; tÞ types one and two, respectively, ka1 is the dissolu-
þ
eðx; tÞ ∂t ∂x tion rate for amorphous polymer of component one,
∂2 CCPP ðx; tÞ and ρ is the total density of the polymer. The deg-
þ De;CPP ðCCPP Þ ð94Þ radation rate constants are normalized to that of
∂x2
þ kCPP mCPP ðx; tÞ amorphous polymer of type one by γ, β, and δ for
n 1=2
the crystalline polymer of type one, amorphous
 ad10−b½pKa;SA −logCSA ðx;tÞ polymer of type two, and crystalline polymer of type
−CCPP ðx; tÞ g two, respectively.
The dimensionless dissolution rate for each mono-
mer i (κi) is defined as follows:
where kSA and kCPP are the dissolution rate constants
of SA and CPP crystals; mSA and mCPP are the mass of ki ðCsat;i −Ci Þ
suspended SA and CPP crystals; CSA,s and pKa,SA are ji ¼ ð96Þ
ka1 Csat;i
the solubility and pKa value of SA, respectively.
A more detailed monomer modeling for surface-
eroding homopolymers and copolymers tablets of CPP- where ki is the dissolution rate of monomer i and Csat,i
SA and CPH-SA was developed by Kipper and is the saturation concentration of monomer i.
Narasimhan [79]. This model accounts for the micro- The dimensionless concentrations (χ1 and χ2) of
phase separation between crystalline and amorphous the monomer of types one and two follow the diffusion
regions and is conducted in the assumption of planar equation with additional source term of dissolution
system with characteristic length scale of y. Therefore, from the pores as follows:
the model sacrifices the geometrical effect for better
understanding in monomer release to explain the ∂v1 ð fma1 þ fmc1 Þj1 r ∂ 2 v1
polymer erosion. In addition, this model involves the ¼ þ D⁎1
∂s e ∂n2 ð97Þ
monomers solubility changes and can be developed ∂v2 ð fma2 þ fmc2 Þj2 r ∂ 2 v2
further to include the pH changes during erosion due to ¼ þ D⁎2
∂s e ∂n2
release of acidic monomers.
There are in total eight fractional changes due to the
where ε is the total porosity of the polymer, D1* and
number of phases considered in the model (including
D2* are the dimensionless diffusivities of polymer
the monomer phases), which are the fractions of
type one and two, respectively, and σ is the dimen-
amorphous polymer ( fa1, fa2) and monomer ( fma1,
sionless surface area per unit volume. ξ = x / y is the
fma2) of components one and two as well as the
dimensionless distance from the original surface of
fractions of crystalline polymer ( fc1, fc2) and monomer
the polymer ( y).
( fmc1, fmc2) of components one and two. The transient
The total porosity of the polymer increases as the
equations for monomer are as follows:
pore radius increases due to the monomer release. It is
expressed as follows:
∂fma1
¼ fa1 −fma1 j1
∂s
∂fma2 ∂e 8ð1−eÞR⁎ ∂R⁎ 8ð1−eÞR⁎
¼ g fa2 −fma2 j2 ¼ ¼ ½ð fma1 þ fmc1 Þj1
∂s ð95Þ ∂s S2 Ss S2
∂fmc1 ð98Þ
¼ bfc1 −fmc1 j1 þ ð fma2 þ fmc2 Þj2 
∂s
∂fmc2
¼ dfc2 −fmc2 j2 where R* is the average dimensionless radius of the
∂s
pore and ℓ is the dimensionless long-period of the
where the dimensionless time parameter (τ) is crystalline lamellae (ℓ = l / y), where l is the long-
normalized by the degradation rate constant of amor- period of crystalline lamellae [79].
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1305

Finally, the cumulative polymer fractional mass ties in radial and axial direction, ε(r,t) and ε(z,t),
loss is computed by integrating the porosity at a certain respectively. They are calculated as follows:
time point in the whole polymer matrix (0 ≤ ξ ≤ ξmax).
This can be expressed as follows: 1X nz
eðr; tÞ ¼ 1− sðiðrÞ; j; tÞ
nz j¼1
Z ð101Þ
Mp;t nmax
1X nr
¼ eðs; nÞdn ð99Þ eðz; tÞ ¼ 1− sði; jðzÞ; tÞ
Mp;l 0 nr i¼1

The cumulative polymer fractional mass flow can be As a result of increasing porosity, diffusion coefficient
fitted to the result from experimental studies. Since of the drug also increases and is calculated as follows:
several values can be taken from previous experimental
studies, unknown parameters that are used for fitting De ðr; tÞ ¼ Dcrit eðr; tÞ
ð102Þ
purpose may give insights on the important mechanisms De ðz; tÞ ¼ Dcrit eðz; tÞ
affecting the polymer erosion. The development of this
model to include the exact geometry and the encapsu- where Dcrit is the critical (maximum) diffusivity
lated drug release due to polymer erosion can be characterized for a specific drug–polymer interaction.
beneficial. Their results also show that the above model pro-
Interestingly, Siepmann et al. [58] applied a vides a reasonable fitting to the in vitro 5-FU release
Monte-Carlo simulation to predict the drug release rate from PLGA-based microparticles in a phosphate
from bulk-erosion PLGA microspheres in a perfect buffer solution with a pH value around 7.4.
sink medium with the experimentally determined
initial drug loading (24% w/w). Due to symmetry 5. Irradiation effect on drug release profile from
condition, a quarter of the sphere is used and extended polymeric microspheres
in the θ-direction to obtain a cylindrical 2D simu-
lation domain. This assumption allows the simula- Irradiation sterilization, i.e., γ-irradiation, of phar-
tion to be conducted in cylindrical coordinates. The maceutical products is popular in recent years due to its
quarter is divided into rectangular pixels and results ease of operation. Conventional sterilization processes,
in cylindrical rings when the rotation around the such as heat and chemical sterilizations often lead to
z-axis is done. Since there is no concentration gradient degradation of polymer backbones and toxicological
in θ coordinate, the drug diffusion in this domain is problems respectively. Dose of 25 kGy (2.5 Mrad) is
described by the Fick's second law for cylindrical essentially accepted to be satisfactory for pharmaceuti-
coordinate without the θ-dependent. cal products sterilization. However, γ-irradiation steril-
Each pixel in the domain has a lifetime and as soon ization is also known to alter the properties of drug
as a pixel is in contact with water, its lifetime (tlifetime) delivery formulations. The other aspect that needs
starts decreasing as follows: caution is that the radiation sterilization may form
toxic degradation products though this possibility is still
 controversial.
ð−1Þ# #
tlifetime ¼ taverage þ ln 1− ð100Þ Due to their biocompatibility, polyesters, which
k 100 exhibit bulk-erosion behavior, are widely used as
polymeric systems for drug delivery applications.
where taverage is the average lifetime of the pixels, λ is They are essentially homo- or copolymer derived from
a constant characterizing the type and physical state of monomers of lactic and glycolic acid. The family
the polymer, and ϑ is a random variable (integer be- members of this type of polymer are PLA, poly(glycolic
tween 0 and 99). acid) (PGA), and the copolymer PLGA with different
Given the initial status for each pixel, the status of lactide to glycolide ratios. However, polyesters are
each pixel at each time step is updated from the Monte- prone to γ-irradiation due to a reduction in mechanical
Carlo simulations that result in the increase of porosi- properties, which leads to chain scission [80]. The theory
1306 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

of “cage effect” was also proposed where the scission of ponential function of the irradiation dose (Fig. 16a).
the polymer chain could be more dominant in the However, the dependence of the diffusivity on the
amorphous regions. PLA and PLGA are more suscep- irradiation dose was not significant and these studies
tible to irradiation degradation than PGA even though concluded that the irradiation dose up to 25 kGy will
the former is more resistant to hydrolytic degradation. not change the erosion behavior of PLGA polymer,
This is due to the presence of the methyl group on lactic but slightly alter the diffusional transport within the
acid which sterically hinders the formation of radical polymer matrix. This conclusion was supported by
pairs and increases the probability of degradation [81]. results from Montanari et al. [89] who observed that
It was observed that the initial molecular weight the decrease in Mw is negligible for irradiation dose
distribution of polymer microspheres decreases sub- below 15 kGy, while a 10%-decrease in Mw was
stantially with the irradiation dose of 0 to 55 kGy [82]. observed for irradiation dose of 25 kGy, which is
When the first 15 kGy of irradiation was applied, slight possibly caused by the increase of initial drug
decrease was observed in weight-average molecular diffusivity. Fitting both non-cellular-automata model
weight (Mw) of 14% as compared to number-average of Koizumi and Panomsuk [7] and cellular-automata
molecular weight (Mn) of 26%. This observation was model of Siepmann et al. [58] (Fig. 16b), the ex-
more pronounced for higher irradiation dose. There- ponential dose-dependent initial diffusivity may be
fore, the increasing irradiation dose resulted in the expressed in the following form:
increasing polydispersity index (I = Mw / Mn) of poly-
mer. In contrast, this phenomenon was not observed by D0 ¼ aexpðbIÞ ð103Þ
Volland et al. [83], Mohr et al. [84], and Yoshioka et al.
[85,86] since the chain scission mechanism is a random where I is the irradiation dose, and a and b are fitting
process by γ-irradiation. Furthermore, Yoshioka et al. constants.
[85,86] suggested that irradiation up to 25 kGy has no Since b is found to be in the order of 0.02 1/kGy in
significant changes in the glass transition temperature the range of 11–33 kGy irradiation for both fittings, it
and the initial drug release rate. implies that the initial diffusivity is not a strong
In another study, Faisant et al. [87,88] correlated the function of the irradiation dosage. This supports the
5-fluorouracil (5-FU) release profile from PLGA previous experimental observation that, below 25 kGy
microspheres with irradiation effect (11–33 kGy) and irradiation, no significant change in PLGA erosion
proposed that the initial drug diffusivity is an ex- behavior is observed.

Fig. 16. (a) The dependence of the initial diffusivity of 5-FU in the PLGA microparticles on the applied γ-irradiation (60Co source); and (b)
experimental results (symbols) and model predictions (dotted line: Koizumi and Panomsuk [7] non-cellular-automata model, solid line: Siepmann
cellular-automata model [58]) of 5-FU release profile from PLGA microparticles after exposure to 4.4 kGy γ-irradiation in phosphate buffer pH
7.4 (reprinted from [88] with permission from Elsevier).
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1307

6. Going to nanoscale: the implication on drug 6.1. Drug targeting and biodistribution
release mechanism
One of the important applications of nanoparticles
Biodegradable microparticulate systems have been for drug delivery is a site-specific delivery [98].
well-studied for controlled release of various types of Biodistribution of drug carriers is very much depen-
drugs, both hydrophobic and hydrophilic. In recent dent on the size of the particles and the properties of
years, there is an obvious trend of increasing interest in the particle surfaces [91] in oral and intravenous
development of nanoparticles for drug delivery applica- delivery. Effect of the size of biodegradable micro and
tions especially in intravenous delivery. Nanoparticles nanoparticles on gastrointestinal has been investigated
may range from sizes 10 nm to 1000 nm [90] and in the studies by Desai et al. [99]. PLGA particles of
include liposomes, micelles, polymer–drug conjugates, size 100 nm, 500 nm, 1 μm and 10 μm were used for
and polymer particles [91]. Novel fabrication processes animal studies (rats) in situ. The uptake efficiency
have also been widely developed to produce nanoscale depends on the type of tissue, but generally the uptake
drug delivery systems. Fabrication techniques such as efficiency of 100 nm size particles in intestinal tissue
modified solvent evaporation methods [92,93], super- was higher as compared to the larger size micro-
critical fluid techniques [94–96] and dialysis methods particles. Histological studies showed that particles at
[97] were some of the feasible routes to achieve drug- 100 nm could diffuse through the submucosal layers
loaded polymeric nanoparticles. For example, several while the larger sized particles (500 nm–10 mm) were
fabrication methods have been adopted and developed found to be concentrated in the epithelial lining of the
for the controlled release of a model hydrophobic tissue. This provides insight to the design of nano-
chemotherapeutic agent, paclitaxel in the form of particles for oral drug delivery systems.
nanoparticles in biodegradable polymer matrix. The For intravenous injection, nanoparticles are general-
various fabrication techniques and size range of ly cleared from the blood rapidly and found to be
particles obtained for paclitaxel delivery system are concentrated in the spleen, liver and blood marrow
summarized in Table 1. [100]. To increase the blood circulation time of

Table 1
Summary of nanofabrication methods for a model hydrophobic drug of paclitaxel
Method Group Year Materials Particle size Remarks
range (nm)
Solvent evaporation Chen et al. [102] 2001 Brij78 + paclitaxel 103.5 ± 29.2 Solid lipid nanocores
Poloxamer F68 + paclitaxel 220 ± 98
Spray drying Mu and Feng [103] 2001 PLGA + DPCC + 800–1000 –
cholesterol + paclitaxel
Temperature induced Lee et al. [104] 2002 F127 a PLGA + paclitaxel 150–600 –
phase transition
Interfacial deposition Fonseca et al. [105] 2002 PLGA + paclitaxel 120–160 –
method
Modified single emulsion/ Mu and Feng [106–108] 2002, 2003 PLGA + paclitaxel 300–900 PVA, DPPC as emulsifier;
solvent evaporation PLA + paclitaxel 600–1000 vitamin E TPGS as emulsifier
and matrix material
Zhang and Feng [109] 2006 PLA–vitamin E 290–333 –
TPGS copolymer + paclitaxel
Controlled solvent Potineni et al. [110] 2003 Poly 1 b + pluronic 100–150 w/o pluronic F-108,
displacement F-108 + paclitaxel particle ∼ 400 nm
Dialysis (self-assembled Xie and Wang [97] 2005 PLGA + paclitaxel 286–291 Vitamin E TPGS as emulsifier
nanoparticles) PLA + paclitaxel 310
Kim et al. [111] 2006 HGC c + paclitaxel 200–416 –
a
F127: poly(ethylene oxide)–poly(propylene oxide)–poly(ethylene oxide) triblock copolymer.
b
Poly 1: poly(ethylene oxide)-modified poly(β-amino ester).
c
HGC: Hydrophobically modified glycol chitosan.
1308 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

nanoparticles, a major challenge would be to avoid its and lidocaine, estimated drug diffusion coefficient in
opsonization and uptake by phagocytic cells. Biode- PLA polymer matrix has been reported in literature
gradable long circulating polymeric nanospheres were ranging from 10− 15 m2/s [112] to 10− 20 m2/s [114]
developed using amphiphilic copolymers of PLGA or respectively. In a study by Kang and Schwendeman
with poly(ethylene glycol) (PEG) in the studies by Gref [115], the diffusion coefficient of a small hydrophobic
et al. [101]. In vivo studies in mice showed an increase probe in PLGA microparticles was measured using
in blood circulation time as the molecular weight of laser scanning confocal microscopy (LSCM) and esti-
PEG used increases. This is most likely due to an mated values of De from 3 to 10 × 10− 16 m2/s were
increase in the protective layer (PEG layer) thickness on obtained.
the outer surface of the nanospheres which could better The release profile of highly hydrophobic drug from
prevent opsonization. A corresponding reduction in polymeric systems is mainly influenced by drug
liver uptake was also achieved. diffusion mechanism. For hydrophobic drug molecules,
due to the very low drug solubility in aqueous medium,
6.2. Drug release from nanoparticles the dissolution constant will generally be very small.
Drug dissolution mechanism has little effect on the drug
6.2.1. Hydrophobic drugs release as the solubility is low and the dissolution rate
A major challenge in drug delivery for anticancer constant (k) will be very small as compared to the
agents is the poor aqueous solubility of many anti- diffusion rate. In this respect, even though one may
cancer drugs. Due to the low aqueous solubility of the expect the diffusion of drug from polymeric micro-
drug and low diffusivity within the polymer matrix, the spheres to be very slow, the time scale of diffusion due
release profile is generally very slow. An example is in to relatively small size of nanoparticles is still smaller
the delivery of a promising anticancer agent paclitaxel than that of dissolution and, therefore, the encapsulated
which is found to be useful against a wide spectrum of drug is preferably released through diffusion in the
carcinomas, especially in breast and ovarian cancer. Its polymer matrix. Table 2 summarizes the estimated time
low aqueous solubility and high crystallinity make it scales for drug release for diffusion coefficients of
difficult to be encapsulated in biodegradable micro- 10− 15 to 10− 20 m2/s ranging from micro- to nanopar-
particles at reasonably high drug loading (N 30%) and ticles. In the case for a crystalline polymer matrix such
the drug release by diffusion is generally very low. as PLA, drug release is often very slow due to a
The use of nanoparticles provides a solution to the corresponding low diffusion coefficient.
slow release profile by providing a much larger surface
area, but compromised with the slower diffusivity in the
Table 2
polymer matrix due to the more compact structures
Diffusional time scale (t / τ) using Harland's dissolution-controlled,
resulted from the fabrication techniques. For a mono- diffusional drug release ( Eq. (61)) from non-swellable polymeric
lithic system with uniform drug distribution with the microspheres
polymeric matrix, release is generally governed by Hydrophobic drug Effective diffusion coefficient, De (m2/s)
diffusion or erosion mechanisms. Drug release from
De = 10− 15 De = 10− 20
common polymeric nanoparticles, such as PLA and
PLGA, is mainly due to diffusion, drug dissolution and t / τ (s) t / τ (s)
subsequently surface erosion or bulk degradation. The 10 μm 2.5 × 104 2.5 × 109
effective diffusion coefficient of drug follows the 1 μm 2.5 × 102 2.5 × 107
combined effects of diffusion through particle pores 100 nm 2.5 2.5 × 105
and diffusion through an intact polymer matrix and is a Hydrophilic drug Effective diffusion coefficient, De (m2/s)
function of the tortuosity and porosity of the polymer.
De = 10− 11 De = 10− 13
Adopting the drug dissolution and diffusion model
by Harland et al. [59] as described earlier in Eqs. (58)– t / τ (s) Di t / τ (s)
−4
(62), the effect of varying particle size on the drug 10 μm 2.5 2.5 × 10 2.5 × 102
release profile of micro and nanoscale particles can be 1 μm 2.5 × 10− 2 2.5 × 10− 6 2.5
100 nm 2.5 × 10− 4 2.5 × 10− 8 2.5 × 10− 2
investigated. For hydrophobic drugs, such as paclitaxel
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1309

From a simple Fickian diffusion model, one may coefficient values of 10− 21 and 10− 22 m2/s were esti-
expect that the release profile to be significantly altered mated for average particle sizes of 1.95 μm [103] and
by changes in particle size as illustrated by the 273 nm [108] respectively. The application of the
difference in diffusion time scale as shown in Table 2. diffusion model is still limited to fully explain the
However, this is not reflective of actual experimentally hydrophobic drug release. In Fig. 17, Fickian diffusion
determined release profiles shown in many studies model was used to fit experimental data for paclitaxel
especially for hydrophobic drug release, for example in release from PLA micro- and nanospheres. It is shown
Liggins and Burt [112] for paclitaxel release from PLA that for particles ranging from 310 nm to 105 μm
polymer matrix and Siepmann et al. [69] for lidocaine although drug release tends to be faster for smaller
release from PLGA polymer matrix. In fact, as shown particles, the change in release profile was not as sig-
earlier in Fig. 11, the release profile for microparticles nificantly different as predicted using the Fickian
ranging from 7.2 μm to 53 μm was not significantly diffusion model using a constant drug diffusion
different as predicted using the same effective diffusion coefficient. The estimated drug diffusion coefficients
coefficient. A better fit of the experimental release used to fit the experimental data are shown in Fig. 17.
profiles was obtained using the mathematical model Here, for a reduction of size from 35–105-μm micro-
with the autocatalytic effect to account for the influence spheres to 310-nm nanospheres, the drug diffusion
of particle size on drug diffusivity. It is found that the coefficient is found to be reduced from 8 × 10− 17 to
drug diffusivity decreases with decreasing size of micro- 8 × 10− 21 m2/s. Therefore, when using the diffusion
spheres since the small diffusion length scale allows acid model, one also has to consider the relationship between
and base molecules to travel from bulk medium to particle size and drug diffusion coefficient.
neutralize the autocatalytic effect. Furthermore, SEM Another important characteristic to be highlighted is
analysis on the surface and cross-section of different size the effect of drug loading on the release profile. The
particles after suspension in physiological solution also diffusion model generally predicts the release profile
showed a corresponding increase in the porosity of based on the fraction of initial loaded drug released over
larger microparticles due to the more pronounced time. However, in many studies, it was observed that the
autocatalytic effect as earlier illustrated in Fig. 12. release profile is dependent on the drug loading in the
Similar trends were also observed for the in vitro polymer matrix. In Mu and Feng [107,108], the
release profiles of paclitaxel from PLGA (75:25) cumulative release percent of paclitaxel from PLGA
polymer matrix for micro- and nanoparticles in the nanospheres generally decreases with an increase in
study by Mu and Feng where effective drug diffusion drug loading, and this is also in consistency with the

Fig. 17. Release profiles for paclitaxel-loaded microparticles ranging from 310 nm to 105 μm are modeled using a pure diffusion model. Release
data are cited from Refs. [97] and [113].
1310 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

observations by Polakovic et al. [114] and Gref et al. of drug from hydrophilic drugs such as proteins also
[101] for the release profiles of lidocaine from PLA and poses a challenge in the control of the drug release
PLGA nanoparticles polymer matrices, respectively. profile. A large initial burst is often associated with
Fig. 18 shows the experimental and fitted release polymeric controlled release devices for hydrophilic
profiles of paclitaxel from PLGA 5050 nanospheres molecules, and this is followed by a phase of very slow
from formulations with drug loading ranging from 2 to drug release. The diffusion coefficient of a model
12% w/w. It can be clearly observed that when drug protein drug bovine serum albumin (BSA) in polymer
loading was increased to 12%, the total fraction of drug systems has been reported to be in the range of
release after a long time was much lower as compared to 10− 11 m2/s [116] to 10− 13 m2/s [117] which is typically
the lower drug loading formulations. The estimated much higher than the diffusion coefficient of hydro-
effective diffusion coefficient for four different drug phobic molecules. Similarly, drug dissolution rate is
loading nanospheres was very close and ranged from 4 also much higher for hydrophilic molecules as
to 9 × 10− 21 m2/s. The possible explanation for this compared to hydrophobic molecules. A typical release
phenomenon is the encapsulation of drug as a molec- profile for hydrophilic drug release from biodegradable
ularly dispersed phase within the polymer matrix at polymer microspheres is illustrated in Fig. 13b. An
lower drug loadings. When the drug loading is high (e.g. initial burst is observed followed by a period of lag
at 30% w/w), the drug loading may exceed the drug phase and subsequent release following polymer
solubility in the polymer matrix and small drug crystals erosion. In contrast to the release profile of hydrophobic
will be embedded heterogeneously in the polymer drugs which is mainly dominated by drug diffusion
matrix [101,114]. In this case, the release of drug will be mechanism, the release of hydrophilic drugs and
much slower than predicted by the diffusion model. proteins is influenced by both drug diffusion and drug
Hence, in the design of controlled release devices dissolution mechanisms. This is illustrated in the studies
for hydrophobic drugs, both factors of drug diffusion by Wong et al. [118] for sustained release of human
coefficient dependence on particle size and the effect of immunoglobulin G (IgG) from biodegradable micro-
drug loading on fraction of drug available for diffusion spheres. Harland's diffusion and dissolution model for
from the polymer matrix play an important role. drug release was found to fit the actual release data
better than for the pure diffusion model within the time
6.2.2. Hydrophilic drugs frame where no bulk degradation was observed.
In contrast to the difficulties encountered in For smaller particle in the nanometer size range,
hydrophobic drug delivery of slow release, the release drug will be released even more rapidly in less than

Fig. 18. Release profiles for different paclitaxel-loaded PLGA 50:50 nanoparticles ranging from 2% to 12% drug loadings. Effective diffusion
coefficients used to fit the models ranged from 4 to 9 × 10− 21 m2/s.
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1311

1 ms. The approximated diffusion time scales as defined where ρ is the density of the interstitial fluid, v is the
earlier in Eq. (61), for hydrophobic and hydrophilic convective velocity vector of the interstitial fluid, ε is
drug release from different size particles are shown in the porosity of the tissue, μ is the mean viscosity of the
Table 2. For both micro- and nanoscale particles, an interstitial fluid, f is the body force, k is the Darcy's
initial burst phase can be predicted by the diffusion– permeability, and CV is the inertial loss coefficient.
dissolution model and observed in experimental data. The two terms on the left-hand side of Eq. (104)
This poses problems in applications where sustained describe the transient and convective contributions,
delivery of drug over longer time periods is desired. The respectively. On the right-hand side of Eq. (104), the
subsequent drug release may be modeled using semi- corresponding terms are the contribution by pressure
empirical models as described in Eq. (81). gradient, viscous loss, body force, Darcy's resistance,
One way to overcome this problem of initial burst and inertial loss, respectively.
effect from water-soluble molecules is the encapsula- The mass conservation (continuity) equation of the
tion of the drug in double-walled composite particles to interstitial fluid can be formulated by considering the
dampen the initial burst effects to allow sustained interstitial fluid source and sink in the tissue as
release over a longer period of time. Double-walled described by Starling's law [121–123]:
microspheres have been fabricated for the sustained
delivery of a water-soluble radiosensitizer drug, etani- ∂q
þ jd ðqvÞ ¼ qðFV −FL Þ ð105Þ
dazole [119]. A possible challenge in delivery of ∂t
hydrophilic drug molecules would be the develop-
ment of nanoscale particles capable of site-specific where
targeting with sustained release over a longer period 
of time. This may be achieved by the development of S
FV ¼ Lp ½ pv −pi −rT ðpv −pi Þ
double-walled nanoparticles with a less permeable V

outer shell made of a hydrophobic matrix to control SL
FL ¼ Lp;L ½ pi −pL 
the initial burst effect. Surface modifications could be V
carried out to enhance the site-specific delivery of the
nanoparticles. The right-hand side of the continuity equation is the
application of Starling's law, where the two terms
7. Simulation of drug delivery in tissue: linking account for the fluid source from the blood vasculature
drug release profile and tissue elimination kinetics and the fluid drainage (loss) to the lymphatic system,
to predict temporal and spatial drug transport respectively. In the source term, Lp and Lp,L are the
hydraulic conductivity of interstitial fluid from the
In recent years, in vitro release profile of drug from vasculature to the interstitium and from the interstitium
controlled release devices has been combined with to the lymphatic system, respectively, (S / V) and (SL /
state-of-the-art computational fluid dynamics simula- V) are the fluid exchange area per unit volume of the
tion to predict the spatial and temporal variation of vasculature and lymphatic system, respectively, pv and
drug transport in the living tissues. Macroscopically, pL are the vasculature and lymphatic pressure,
the tissue, which can be either normal or tumor tissue, respectively, pi is the interstitial fluid pressure in the
is ideally assumed as an isotropic porous medium, interstitium, σT is the osmotic reflective coefficient, πv
which is described by Darcy's law [120] for the is the osmotic pressure of the vasculature, and πi is the
balance of linear momentum in the tissue interstitium. osmotic pressure of the interstitium. The variables in
The full-form of the momentum equation is expressed the sink term are also similar, except that hydraulic
as follows: conductivity, surface area of exchange, and the
 pressures are specific to the lymphatic system.
q ∂v l
To account for the drug distribution in the tissue, the
þ vdjv ¼ −jpi þ j2 v
e ∂t e species continuity equation is coupled with the
l 1
momentum and overall continuity equations and solved
þ qf− v− CV qjvjv ð104Þ
k 2 simultaneously. The solute transport in biological
1312 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

systems can be assumed to be a combination of When solute transport is assumed to be the


convection and diffusion in the porous medium. Here, combination of convection, diffusion, and elimination
the concept of percolation theory for the solute transport kinetics, the prediction of transport parameter values,
in the porous medium [125,126] is applied with i.e. diffusion coefficient in the tissue and the
additional feature of solute elimination kinetics. In elimination rate constant is also crucial. These two
addition to the drug elimination due to degradation and important transport parameters were initially deter-
uptake by the tissue cells, species conservation also mined in a coupled parameter of the diffusion/
involves the possible source of drug transported from the elimination modulus or Thiele modulus (ϕs) by fitting
vasculature (Fs) and loss due to convective to the the species balance equation with the transient and
lymphatic system (F1s), as shown in Fig. 19. For the spatial drug concentration obtained from in vivo
drug (solute) transport, a diffusion/kinetics model is animal experiments. This parameter can be uncoupled
used as follows: by conducting a separate diffusivity measurement in
ex vivo tissue, in which the elimination can be assumed
∂Ci
¼ De:tissue j2 Ci −jd ðrF vCi Þ þ Fs −F1s −RðCi Þ to be negligible, by using several techniques, i.e.
∂t
integrative optical imaging [127], fluorescence recov-
ð106Þ ery after photobleaching (FRAP) [128], and multipho-
where rF is the retardation factor (the ratio of the drug ton microscopy [129]. Recently, some developments
velocity to the fluid velocity), De:tissue and Ci are dif- for in vivo diffusivity measurements have been made
fusion coefficient and concentration of drug molecules [128–132], even though a new and accurate model and
in the tissue interstitium, respectively. It is important to its assumptions should be carefully employed to obtain
note that De:tissue is different from the previous effective the effective diffusivity without possible contribution
diffusivity in polymer matrix (De) since De:tissue is used by drug elimination during experimental time scale.
to account for the drug diffusivity in the tissue This review section will be sub-sectioned based on
interstitium when it is assumed as a porous medium. the targeted tissue for drug delivery. The rationale
On the right-hand side of Eq. (106), the first two terms behind this classification is the differences in the
account for the diffusion and convective contributions, nature of tissues such that different considerations
the third and fourth terms describe the possible drug must be taken into account. Each sub-section will
source and sink from the fluid exchange with the discuss the development of the modeling works of
vasculature and lymphatic systems, and the last term is drug delivery in the respective targeted tissue of
the rate of drug degradation and uptake in the tissue. interest. The mathematical models developed in this

Fig. 19. The solute transport mechanisms in the tissue, given by convection of fluid, diffusion, elimination in the extracellular space, receptor- and
non-receptor-based cells internalization, and intracellular elimination (reprinted from [124] with kind permission from Springer Science and
Business Media).
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1313

area are mostly about the drug release from polymer is plasma concentration, Vmax and Km are Michaelis–
implants that have certain characteristics of release Menten parameters, ke is the first-order elimination
profile. The discussion of the drug penetration and constant due to non-enzymatic reactions, and kapp is
dominant transport mechanism will be highlighted. the lumped first-order elimination rate constant.
On the right-hand side, the first term designates the
7.1. Simulation of drug delivery in brain drug loss due to the BBB permeability, the second and
third terms are the drug elimination due to the
The administration of drug in brain tissue or central enzymatic and non-enzymatic processes, respectively.
nervous system (CNS) poses a challenge to overcome Since the drug permeability of the BBB is low
the very selective permeability of the blood capillaries, (CpL ≪ Ci) and the concentration of the solute in the
which is widely known as the blood–brain barrier brain is very low so that the enzymatic process is in the
(BBB). The BBB exists due to capillaries that have order of first-order magnitude (Ci ≪ Km), the overall
continuous lining of endothelial cells held together by reaction can be assumed to be a first-order kinetics
tight junctions. Therefore, non-lipid solutes, either with an apparent elimination rate constant (kapp).
drug or protein, enter the brain slower as compared to This model was applied to describe the in vivo drug
other tissues. In addition, since most of the targeted release from the implant in the rat [134], rabbit [135],
Drug Delivery in the brain tissue is aimed to treat the and monkey brain tissues [136]. In these models, the
brain tumor, there exists an additional barrier that aforementioned one-dimensional diffusion-and-kinet-
needs to be overcome. It is the presence of elevated ics models are used for carmustine delivery from
interstitial fluid pressure in the tumor that reduces the polymer implants with diffusion process as the main
driving force of the solutes to be extravasated into the mechanism. In the absence of the convective or bulk
brain tissue from the blood vasculatures [123]. Since flow, the only fitting parameter is the Thiele modulus
these two barriers are difficult to overcome for drug to (ϕs), as given below:
penetrate the vasculature in the brain tissue, several
sffiffiffiffiffiffiffiffi
invasive approaches of localized drug delivery, i.e.
kapp
polymer implants and localized injection become more /s ¼ a ð108Þ
popular. Even for the site-specific implantation of De
controlled drug delivery devices, such as polymer
implants, the high interstitial fluid pressure barrier, where a is the length scale of the polymer geometry.
when tumor tissue is present, still poses a challenge of In brief, high ϕs results in a steeper local
reducing the extent of drug reaching the tumor region concentration gradient due to the rapid elimination
due to the outward flow of interstitial fluid. when drug diffuses in the extracellular matrix.
Saltzman and Radomsky [133] first developed a Assuming the steady-state solution, the experimental
diffusion/kinetics model for the drug release from data of drug spatial distribution fit well to the model's
polymer implants into brain tissue. The transport is prediction. Here, the steady-state assumption is valid
assumed to be mainly by diffusion, whereas the over the period of study since the mean steady-state
complex elimination reactions due to irreversible ϕs is close to the transient ϕs. An example of the
metabolism, reversible binding to the fixed tissue fitting result for carmustine distribution in rat brain
components, and capillaries partitioning, are finally is shown in Fig. 20. In addition, the summary of
simplified to be first-order kinetics. The complex and these works has been also nicely described in the
simplified forms of the elimination term are given by: literature [137].
 However, this simplified model may not be valid for
Ci Vmax Ci the case when the influence of convective flow is
−RðCi Þ ¼ kBBB −CpL þ þ ke Ci ckapp Ci
e Km þ Ci significant, i.e., when vasogenic edema takes place
ð107Þ [138–140]. Vasogenic edema is the condition of
significant increase of production of fluid in the
where kBBB is the permeability of the BBB (defined extracellular space (ECS) of the brain due to a surgical
based on extracellular matrix drug concentration), CpL trauma that damages the BBB and hence increases the
1314 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

Fig. 20. Carmustine concentration profile in the rat brain tissue in the vicinity of a polymer implant at time point of (a) day 1, (b) day 3, (c) day 7,
and day 14. The experimental results are denoted by symbols, whereas the steady-state diffusion/kinetic model with fitted ϕs for each time point is
shown in solid lines (reprinted from [134] with kind permission from Springer Science and Business Media).

blood capillaries permeability. Thus, at the site of model drug used in this 2D simulation is interleukin-2
edema, there is a substantial increase of pressure (IL-2), which is administered via both bolus and
gradient that leads to the bulk flow, which may carry microspheres injection as a comparison study. The
the drug released from the implant site. This hy- pressure and velocity profiles are tracked from mag-
pothesis is also bolstered by simulation results that netic resonance imaging (MRI) after contrast agent
showed the importance of the convective flow when administration. In the modeling perspective, this
edema takes place that causes a deeper penetration of study has included the flow in porous medium using
carmustine. the extended form of Darcy's law and a constant
Kalyanasundaram et al. [141] utilized the finite lumped parameter in the species balance due to
element method to develop a two-dimensional (2D) clearance into capillary and enzymatic elimination.
model of drug delivery in rabbit brain (Fig. 21). The However, the transient flow profile in the brain and its
grid for this model uses realistic brain geometry and influence on drug penetration were not described in
incorporated the differences in transport properties of detail.
the white and gray matters, the effect of ventricle at the The three-dimensional simulation of human brain
boundary, and the effect of edema, which may cause a tumor of primitive neuroectodermal tumor (PNET) was
significant increase in interstitial fluid pressure in the initiated by Wang et al. [142,143]. Here, the simulation
brain. The ventricle, in which approximately half of study is aimed to analyze and compare several car-
the total cerebrospinal fluid (CSF) is produced in the mustine delivery techniques to a real human brain
brain (besides from blood vessels), becomes an in- tumor case, i.e. systemic administration (injection) and
ternal boundary to provide an escape route for the drug controlled release from polymer. Using a simplified
through the bulk flow of interstitial fluid. As the white geometry of an isolated tumor, the simulation is con-
matter has more regular arrangement of nerve fibers ducted in computational fluid dynamics software
than in gray matter, the transport resistance in the white package to solve simultaneously continuity, momen-
matter is relatively lower than in the gray matter. The tum, and drug species equations. For instance, the
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1315

Fig. 21. (a) The 2D finite element meshes of the rabbit brain (1 mm anterior to the bregma). The grid contains 800 triangular elements and includes
the implant and different brain regions (white matter, gray matter, and ventricle). (b) The IL-2 distribution profile (in g/cm3) involving diffusion,
convective transport, and edema effect at time point of hour 1 (A), hour 6 (B), and hour 12 (C) (reprinted from [141] with permission from The
American Physiological Society).

drug source term (Fs) in Eq. (106) is modeled by the sure time, and reduced systemic toxicity as com-
following: pared to bolus injection treatment. This study
provides insight that the computer simulation can
Fs ¼ Fv ð1−rÞCv þ PS=V ðCv −Ci ÞPev =ðePev −1Þ be employed to optimize the type and dosage
ð109Þ of drug delivery in the tissue before any experimen-
tal work is conducted for validation. Using a 2D
where σ and Cv are the osmotic reflection coefficient simplified geometry of PNET together with discrete
and drug concentration in the vascular space, placement of Gliadel® wafer discs, the simulation
respectively. P is the transcapillary permeability of to investigate the transient interstitial fluid flow due
the drug molecule. The transcapillary Peclet number to surgical cavity with the presence of post-surgery
(Pev) is given by: edema was conducted by Teo et al. [144]. The study
evaluated the impact of surgical excision of tumor
Fv ð1−rÞ on the resulting transient interstitial fluid flow field
Pev ¼ ð110Þ
PS=V and efficiency of drug delivery. It further suggests
that the presence of the post-surgery edema
This model examines in detail the contribution of increases the interstitial pressure and fluid velocity,
convective transport of macromolecular and small thereby causing higher relative toxicity in the
molecular drugs in the vicinity of tumor and repro- surrounding normal tissue.
duced Baxter and Jain's model prediction for Tan et al. [145,146] improved the 3D human brain
elevated pressure profile in the tumor. However, tumor model by reconstructing the 3D isolated tumor
the transport properties are only distinguished in geometry from a magnetic resonance image (MRI)
terms of tumor and normal brain tissue and the and incorporating the effect of vasogenic edema (Fig.
possible effect of edema has not been incorporated. 22). In both studies, the model drug is a radio-
Simulation results suggest that, since the penetration sensitizer, etanidazole. Since there is no functioning
depth of carmustine is very short due to high lymphatics in the brain, F1s is also negligible
elimination rate, the polymer implantation approach (F1s ∼ 0). In view of the order of magnitude for
provided higher mean concentration, longer expo- Pev, a simplified expression for Fs is used. The
1316 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

Fig. 22. (a) Contour plot of etanidazole concentration (in kg/m3) in a planar cut-section parallel to the z-axis at hour 20; and (b) mean etanidazole
concentration profiles for zero-order release system (A) and double-burst release system (B). The dotted lines represent the fictitious minimum
threshold concentration of etanidazole (reprinted from [145] with permission from Wiley-Liss, Inc., a subsidiary of John Wiley & Sons, Inc.
Copyright © 2003).

constitutive equations for Fs and R terms in Eq. interstitial fluid to the open site. This leads to the non-
(104) for various regimes are given by: uniform distribution of the drug in the tumor zone;
8 9 thus, eventually decreases the drug therapeutic index.
−t=tr
>
< S0 e in wafers >
= However, since in both studies the model geometry is
PS
Fs;etanidazole ¼ − C in tumor and tissues the isolated tumor, the overall picture of the fluid flow
>
: V >
; in the brain, e.g. the contribution of fluid flow from
8 0 in cavity 9
< kapp;cavity C in cavity = ventricle, has not been taken into account.
Retanidazole ¼ kapp;tissue C in tumor and normal tissues The drug simulation in brain tissue poses great
: ;
0 elsewhere challenges to provide more meaningful results for
ð111Þ biomedical applications. They include the incorporation
of the accurate determination of geometric reconstruc-
where S 0 and t r can be obtained by fitting tion and transport parameters by fitting to the model
experimental etanidazole release profile. firstly with the normal case. In the former case, the
In the first study, the different mechanisms of drug automatic reconstruction of normal tissue and tumor
release from polymer implants, i.e. linear and double- geometry from medical images, i.e. magnetic resonance
burst release systems, were investigated [145]. The imaging (MRI) and computed tomography (CT) scan, is
simulation results suggest that the double-burst release now possible owing to recent development of biomed-
system provided a higher drug penetration depth and ical imaging software. This can be utilized to generate
reasonably similar therapeutic index as compared to an actual modeling grid that allows a patient-specific
those of linear release system. In this case, the modeling work to be conducted. In the latter case, the
therapeutic index is defined as the ratio of mean drug difficulty to obtain transport parameters experimentally
concentration in the tumor region to that of the normal can be overcome by fitting these parameters to the
tissue region. model with actual modeling grid owing to accurate
The second study by Tan et al. [146] emphasized on boundary conditions of physiological variables, such as
the effect of the surgical opening on the delivery ventricular (choroid plexus) pressure, which are easier
efficiency of drug and polymer implants (Fig. 23). It is to measure experimentally. In addition, different therapy
shown that the drug efficacy was significantly lowered strategies, i.e. polymer implant, site-specific direct
for open tumor case due to the convective transport of injection, systemic administration from vasculature,
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1317

Fig. 23. (a) Pressure (Pa) and (b) velocity (m/s) contour plots at hour 20; (c) the velocity (m/s) vector plot showing the fluid loss due to surgical
opening; and (d) drug distribution (kg/m3) contour plot at hour 20 (reprinted from [146] with permission from Wiley-Liss, Inc., a subsidiary of
John Wiley & Sons, Inc. Copyright © 2003).

also give variations to the modeling works in tissue drug trans-catheter administration via branches of hepatic
delivery. artery. The transient vascular pressure ( pv) is assumed
to be a step function followed by a decreasing
7.2. Simulation of drug delivery in liver exponential function, which can be written as follows:

An example of drug delivery in liver is given by the   


t
trans-catheter oily chemoembolization for the treat- pv ¼ pv;0 1 þ Aexp ð112Þ
tc
ment of hepatoma. The procedure involves the
insertion of a catheter through an opening in the
groin of the patient into the hepatic artery or its In this equation, pv,0 is the initial vascular pressure, A
branches. The tube is then guided up the artery until it is the fraction increase from the initial vascular pres-
reaches the hepatoma in the liver. The anticancer drug sure due to injection of the drug and fluid, and tc is the
doxorubicin is mixed with lipiodol, an oily contrast time-constant of decay, which depends on the tumor
medium, and is injected through the tube into the blood vasculature. The term A can be estimated for a bed of
capillaries. Here, one simulation study of doxorubicin uniform parallel capillaries where the pressure drop in
delivery to hepatoma was performed by Goh et al. a laminar blood flow is related to the mean capillary
[147]. The model also utilizes computational fluid flow velocity and capillary density for a given bulk
dynamics to solve the pressure and velocity profile in flow rate. Furthermore, the plasma pharmacokinetics
the liver tissue. However, the drug is released from of doxorubicin is obtained from the literature data and
1318 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

Fig. 24. (a) Simulation geometry of femur segment with implantation of PLGA drug discs; and (b) the gentamycin concentration profile released
from the discs along the line x = 0 in the xy-plane at different time points after the implantation (reprinted from [147] with permission from Wiley-
Liss, Inc., a subsidiary of John Wiley & Sons, Inc. Copyright © 2001).

fitted as piecewise-smooth functions for the temporal gradable polymer poly(methylmethacrylate) (PMMA)
variation of doxorubicin concentration. beads (Fig. 24). In this simulation, effect of the clotting
This mathematical model is used to evaluate the process due to surgical trauma is investigated. The
sensitivity of the following few physiology and clotting process is modeled by varying Darcy's
operating parameters for actual clinical treatments, permeability with time (anisotropic in nature, relative-
such as therapeutic index, injection rate, extent of ly higher permeability is observed in the axial
angiogenesis, lymphatic drainage, and DNA binding direction), where the initial permeability is assumed
kinetics. It is found that, in the trans-catheter treatment, to be comparable to that of in vitro gel and the final
increasing the injection volume does not result in a permeability is predicted to be the cortical bone axial
substantial drug concentration increase in the tumor, permeability.
core, and normal tissues. This is due to the small The simulation model provides quantitative predic-
volume injection compared to fluid extravasation from tions for the region of effective therapy for gentamycin
the blood vessel so that the fluid volume acts as a concentration greater than MIC (minimum inhibitory
“perfectly-stirred vessel” that dissolves the drug at a concentration) and relates this to the contribution of
very rapid time without increasing its extravasation various transport properties and interstitial fluid
rate. Diffusion is the main mechanism of transport in distribution. Comparison between the baseline simu-
the interstitium for free doxorubicin molecules. A lation and the implantation of PLGA discs and PMMA
smaller vascular exchange area results in lower beads is also investigated where the results show a
interstitial drug concentration. Lymphatic drainage in comparable overall increase in drug concentration
the tumor causes negligible reductions in the mean in all zones. It is shown that the implantation of
concentrations in all three different zones. Cellular PMMA beads into the bone led to higher mean con-
metabolism and DNA binding kinetics also decrease centration in the marrow that may be due to the
the mean concentrations of drug by about 15 to 40% difference in the geometry between the disc and the
respectively, as compared to the baseline case. bead. The study performs sensitivity analysis on two
different formulations of discs. One of them is fab-
7.3. Simulation of drug delivery in bone ricated from pure PLGA 50:50 only while the other has
30% b-tricalcium phosphate blended in it. The salient
Lee et al. [148] developed a simulation model for feature of both release profile is that they are biphasic,
antibiotic (gentamycin) delivery to bone tissue for exhibiting an initial burst followed by a second burst
osteomyelitis treatment that was delivered by either about 3 weeks later. However, the differences between
biodegradable polymer (PLGA) discs or non-biode- the two profiles include the amount of initial burst and
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1319

the time when the second burst takes place. However, tissue due to an increase in its collagen content [151].
PMMA carriers are not biodegradable, and have to be Theoretically, the former will reduce the drug delivery
surgically removed at the end of treatment, causing efficacy as the additional convective effect will “wash”
unnecessary additional trauma to patient. In contrast, the drug away from the targeted site. On the other
PLGA discs do not involve any second surgery for hand, the latter can help control the drug penetration
removal. confined only in smaller region, i.e. tumor. Here, the
CFD simulation can be used to optimize the interplay
7.4. Challenges ahead the simulation of drug delivery effect of the irradiation to the Drug Delivery efficacy,
in tissue for instance, by the scheduling of irradiation and
chemotherapy treatments.
Since the patient validation for tissue modeling is On the other hand, the anti-angiogenic therapy is
still not fully attainable at least until now, the one of the most recent important strategies to combat
modeling plays a role in suggesting the appropriate tumor since theoretically it will help inhibit the blood
release profile in order to optimize the therapeutic vessel vascularization that is required for tumor
efficiency of an agent. This refers to the highest growth. Jain [152] hypothesized that the anti-
toxicity for the targeted tissue (e.g. tumor) and the angiogenic therapy normalized the tumor vasculari-
least for the surrounding normal tissue. The simula- zation so it inhibited the tumor growth and helped the
tion results can also be reverted back to provide Drug Delivery to the tumor site. Kunkel et al. [153]
recommendation on the required (optimized) release confirmed the inhibition of glioma growth in
profile at the drug device fabrication stage. Here, intracranially implanted animal tumor models, in
there are two important issues that have to be which tumor volume is reduced by 59% as compared
addressed. Firstly, it is important to fit the modeling to the control. It was also suggested that the
concept and results to the existing normal tissue fluid interstitial fluid pressure decreases in subcutaneously
flow data. It can be achieved by allowing some implanted U87 (glioblastoma) tumor in nude mice
undetermined variables that are difficult to obtain [154]. From the transport perspective, this concept
experimentally to fit the baseline (normal) case of the can be interpreted as the reduced convective effect
tissue fluid flow. Secondly, the emerging capability of from the tumor site, thereby, increasing the chemo-
medical imaging equipment and software to create a therapy efficiency in the tumor site. This particular
patient-specific simulation grid is also crucial. The effect is also interesting to be investigated through
patient-specific simulation grid can provide a patient- future simulation studies.
specific recommendation in terms of required dosage
and release profile. 8. Conclusions
The emerging adjunct therapies for tumor treatment
(besides chemotherapy), i.e. radiotherapy and anti- The mathematical model and simulation of drug
angiogenic therapy, become an important area where release from polymeric microspheres have developed
the modeling effort can play a crucial role. Radiother- and evolved to various approaches and concepts.
apy helps the chemotherapeutic agent eliminates the Based on the nature of the polymeric matrices used
tumor cells when patient undergoes irradiation after and the behavior during drug release, these models
surgery. Qin et al. [149] suggested that irradiation can be distinctly categorized to diffusion-controlled,
induced the opening of BBB and, when it was swelling-controlled, and erosion-controlled systems.
combined with chemotherapy, helped enhance the For all types of systems, chemical reactions and
survival rates by 200% (means that the patient with mass transfer processes, which are affected by
combined therapy had three times longer survival polymer and drug type, device size, shape, compo-
time) with statistically significant improvement. Sev- sition, and encapsulation techniques, are crucial in
eral studies suggest that irradiation improves the controlling the drug release. In this case, the choice
hydraulic conductivity of BBB by increasing the of an appropriate mathematical model for a specific
fluid production from blood vessels [150] and de- drug delivery system has to be carried out with
creasing the interstitial drug diffusion in the tumor caution.
1320 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

Both molecular-level study (e.g. diffusivity, drug– Acknowledgements


polymer interaction parameters) and macroscopic study
(e.g. drug release profile, molecular weight decrease This work was supported by Science and Engineer-
profile) are important to integrate the understanding of ing Research Council (SERC), Singapore and National
drug release mechanisms from polymeric systems with University of Singapore under the grant number R279-
mathematical model development. The challenges 000-208-305. The authors are grateful to Professor
posed here are to test the drug release in the in vivo Kenneth A. Smith (Department of Chemical Engineer-
condition. The surrounding environment, such as pH, ing, MIT) for helpful discussion on this project. We
osmotic pressure, and tissue elimination, will influence would also thank J. Xie and L.K. Lim for their technical
the drug release profile, especially for bioerodible support in the course of this study.
polymeric systems, in which the degradation process
would be strongly affected. One may expect that, References
especially for bioerodible polymer matrix system, the in
vivo release is faster than in vitro release as degradation [1] K.W. Leong, R. Langer, Polymeric controlled Drug Delivery,
Advanced Drug Delivery Reviews 1 (1987) 199–233.
rate is enhanced due to enzymatic reactions. Other
[2] J.-M. Vergnaud, Controlled Drug Release of Oral Dosage
important challenge that has been addressed in this Forms, Ellis Horwood Limited, Chichester, 1993.
review is to account for the possibility of altering the [3] J. Crank, The Mathematics of Diffusion, 2nd ed., Clarendon
drug release profile by applying the sterilization pro- Press, Oxford, 1975.
cedures, i.e. γ-irradiation. [4] R.W. Baker, H.K. Lonsdale, Controlled release: mechanisms
Even though the use of simple empirical or semi- and rates, in: A.C. Tanquarry, R.E. Lacey (Eds.), Controlled
Release of Biologically Active Agents, Plenum Press, New
empirical models is sufficient, the detailed mechanism York, NY, 1974, pp. 15–71.
regarding the drug release process is not fully [5] R. Baker, Controlled Release of Biologically Active Agents,
elucidated. Therefore, there is still room for improve- John Wiley & Sons, 1987.
ment for mechanistic model to highlight the important [6] T. Higuchi, Mechanism of sustained-action medication:
drug release mechanism for different systems. In this theoretical analysis of rate of release of solid drugs dispersed
in solid matrices, Journal of Pharmaceutical Sciences 52
case, when reliable and detailed information on drug (1963) 1145–1149.
release process is available, mechanistic models are the [7] T. Koizumi, S.P. Panomsuk, Release of medicaments from
best to apply. However, it does not mean that the more spherical matrices containing drug in suspension: theoretical
complex the mechanistic model is, the better it is. The aspects, International Journal of Pharmaceutics 116 (1995)
most ideal model is the simplest model that is able to 45–49.
[8] D.S. Cohen, T. Erneux, Controlled drug release asymptotics,
satisfy the theory of step-by-step drug release mechan- SIAM Journal on Applied Mathematics 58 (4) (1998) 1193–1204.
isms for general cases and highlight the important [9] P.I. Lee, Diffusional release of a solute from a polymeric
process affecting the drug release profile. matrix — approximate analytical solutions, Journal of
Future drug delivery modeling efforts would be Membrane Science 7 (1980) 255–275.
mainly focused on the drug transport in tissue after it is [10] D.R. Paul, S.K. McSpadden, Diffusional release of a solute
from a polymer matrix, Journal of Membrane Science 1
released from systemic administration or implanted (1976) 33–48.
polymeric devices. Linked to the concept of compu- [11] M.J. Abdekhodaie, Y.-L. Cheng, Diffusional release of a
tational fluid dynamics, the modeling involves the dispersed solute from a spherical polymer matrix, Journal of
complex interplay of possible processes of drug Membrane Science 115 (1996) 171–178.
[12] M.J. Abdekhodaie, Y.-L. Cheng, Diffusional release of a
diffusion and convective transport in extracellular
dispersed solute from planar and spherical matrices into finite
matrices, drug extravasation from blood vessels (when external volume, Journal of Controlled Release 43 (1997)
the device is systematically administered), tissue 175–182.
elimination by lymphatic system, and intracellular [13] X.Y. Wu, Y. Zhou, Studies of diffusional release of a dis-
internalization and degradation. In this case, the in persed solute from polymeric matrixes by finite element
vitro drug release profile may not be appropriate for method, Journal of Pharmaceutical Sciences 88 (10) (1999)
1050–1057.
this purpose of modeling as the realistic physiological [14] Y. Zhou, X.Y. Wu, Theoretical analyses of dispersed-drug
condition may substantially change the mechanism of release from planar matrices with a boundary layer in a finite
drug release. medium, Journal of Controlled Release 84 (2002) 1–13.
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1321

[15] Y. Zhou, X.Y. Wu, Modeling and analysis of dispersed-drug [30] J.S. Papanu, D.S. Soane, A.T. Bell, D.W. Hess, Transport
release into a finite medium from sphere ensembles with a models for swelling and dissolution of thin polymer films,
boundary layer, Journal of Controlled Release 90 (2003) Journal of Applied Polymer Science 38 (1989) 859–885.
23–26. [31] P. Colombo, R. Bettini, P. Santi, A.D. Ascentiis, N.A.
[16] J. Siepmann, A. Peppas, Modeling of drug release from Peppas, Analysis of the swelling and release mechanisms
delivery systems based on hydroxypropyl methylcellulose from Drug Delivery systems with emphasis on drug sol-
(HPMC), Advanced Drug Delivery Reviews 48 (2001) ubility and water transport, Journal of Controlled Release
139–157. 39 (1996) 231–237.
[17] B. Narasimhan, Mathematical models describing polymer [32] R.S. Harland, A. Gazzaniga, M.E. Sangalli, P. Colombo, N.A.
dissolution: consequences for Drug Delivery, Advanced Drug Peppas, Drug–polymer matrix swelling and dissolution,
Delivery Reviews 48 (2001) 195–210. Pharmaceutical Research 5 (1988) 488–494.
[18] P. Costa, J.M.S. Lobo, Modeling and comparison of [33] U. Conte, P. Colombo, A. Gazzaniga, M.E. Sangalli, A. La
dissolution profiles, European Journal of Pharmaceutical Manna, Swelling-activated Drug Delivery systems, Bioma-
Sciences 13 (2001) 123–133 P.I. terials 9 (1988) 489–493.
[19] Lee, N.A. Peppas, Prediction of polymer dissolution in [34] B. Narasimhan, N.A. Peppas, Molecular analysis of Drug
swellable controlled-release systems, Journal of Controlled Delivery systems controlled by dissolution of the polymer
Release 6 (1987) 207–215. carrier, Journal of Pharmaceutical Sciences 86 (3) (1997)
[20] T. Alfrey Jr., E.F. Gurnee, W.G. Lloyd, Diffusion in glassy poly- 297–304.
mers, Journal of Polymer Science. Part C 12 (1966) 249–261. [35] J.S. Vrentas, C.M. Vrentas, Energy effects for solvent self-
[21] R.W. Korsmeyer, S.R. Lustig, N.A. Peppas, Solute and diffusion in polymer–solvent systems, Macromolecules 26
pentrant diffusion in swellable polymers. I. Mathematical (1993) 1277–1281.
modeling, Journal of Polymer Science. Polymer Physics [36] J.S. Vrentas, C.M. Vrentas, Solvent self-diffusion in rubbery
Edition 24 (1986) 395–408. polymer–solvent systems, Macromolecules 27 (1994)
[22] R.W. Korsmeyer, E. von Meerwall, N.A. Peppas, Solute and 4684–4690.
penetrant diffusion in swellable polymers. II. Verification of [37] J.S. Vrentas, C.M. Vrentas, Solvent self-diffusion in glassy
theoretical models, Journal of polymer science. Polymer polymer–solvent systems, Macromolecules 27 (1994)
physics edition 24 (1986) 409–434. 5570–5576.
[23] P.L. Ritger, N.A. Peppas, A simple equation for description of [38] P.J. Flory, J. Rehner, Statistical mechanics of cross-linked
solute release. I. Fickian and non-Fickian release from non- polymer networks. II. Swelling, The Journal of Chemical
swellable devices in the form of slabs, spheres, cylinders or Physics 11 (1943) 521–526.
discs, Journal of Controlled Release 5 (1987) 23–26. [39] J. Siepmann, H. Kranz, R. Bodmeier, N.A. Peppas, HPMC-
[24] P.L. Ritger, N.A. Peppas, A simple equation for description of matrices for controlled Drug Delivery: a new model
solute release. II. Fickian and anomalous release from combining diffusion, swelling, and dissolution mechanisms
swellable devices, Journal of Controlled Release 5 (1987) and predicting the release kinetics, Pharmaceutical Research
37–42. 16 (11) (1999) 1748–1756.
[25] H. Kim, R. Fassihi, Application of binary polymer system in [40] J. Siepmann, N.A. Peppas, Hydrophilic matrices for controlled
drug release rate modulation. 2. Influence of formulation Drug Delivery: an improved mathematical model to predict the
variables and hydrodynamic conditions on release kinetics, resulting drug release kinetics (the “sequential layer” model),
Journal of Pharmaceutical Sciences 83 (1997) 323–328. Pharmaceutical Research 17 (10) (2000) 1290–1298.
[26] J.L. Ford, K. Mitchell, P. Rowe, D.J. Armstrong, P.N.C. [41] J. Siepmann, A. Sterubel, N.A. Peppas, Understanding and
Elliott, C. Rostron, J.E. Hogan, Mathematical modeling of predicting Drug Delivery from hydrophilic matrix tablets
drug release from hydroxypropylmethylcellulose matrices: using the “sequential layer” model, Pharmaceutical Research
effect of temperature, International Journal of Pharmaceutics 19 (3) (2002) 306–314.
71 (1991) 95–104. [42] J. Siepmann, H. Kranz, N.A. Peppas, R. Bodmeier,
[27] N.A. Peppas, J.J. Sahlin, A simple equation for the description Calculation of the required size and shape of hydroxypropyl
of solute release. III. Coupling of diffusion and relaxation, methylcellulose matrices to achieve desired drug release
International Journal of Pharmaceutics 57 (1989) 169–172. profiles, International Journal of Pharmaceutics 201 (2000)
[28] R.T.C. Ju, P.R. Nixon, M.V. Patel, Drug release from hydro- 151–164.
philic matrices. 1. New scaling laws for predicting polymer [43] A.G. Hausberger, P.P. DeLuca, Characterization of biode-
and drug release based on the polymer disentanglement con- gradable poly(D,L-lactide-co-glycolide) polymers and micro-
centration and the diffusion layer, Journal of Pharmaceutical spheres, Journal of Pharmaceutical and Biomedical Analysis
Sciences 84 (12) (1995) 1455–1463. 13 (6) (1995) 747–760.
[29] R.T.C. Ju, P.R. Nixon, M.V. Patel, D.M. Tong, Drug release [44] N. Kumar, R.S. Langer, A.J. Domb, Polyanhydrides: an over-
from hydrophilic matrices. 2. A mathematical model based on view, Advanced Drug Delivery Reviews 54 (2002) 889–910.
the polymer disentanglement concentration and the diffusion [45] F. Von Burkersroda, L. Schedl, A. Göpferich, Why degradable
layer, Journal of Pharmaceutical Sciences 84 (12) (1995) polymers undergo surface erosion or bulk erosion, Biomater-
1464–1477. ials 23 (2002) 4221–4231.
1322 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

[46] C.N. Satterfield, C.K. Colton, W.H. Pitcher Jr., Restricted evaluations, International Journal of Pharmaceutics 200
diffusion in liquids within fine pores, AIChE Journal 19 (3) (2000) 115–120.
(1973) 628–635. [63] C. Raman, C. Berkland, K. Kim, D.W. Pack, Modeling small-
[47] W.M. Deen, Hindered transport of large molecules in liquid- molecule release from PLG microspheres: effects of polymer
filled pores, AIChE Journal 33(9) 1409–1425. degradation and nonuniform drug distribution, Journal of
[48] W. Weibull, A statistical distribution of wide applicability, Controlled Release 103 (2005) 149–158.
Journal of Applied Mechanics 18 (1951) 293–297. [64] J. He, C. Zhong, J. Mi, Modeling of drug release from bioerodible
[49] F. Langenbucher, Linearization of dissolution rate curves by polymer matrices, Drug Delivery 12 (2005) 251–259.
the Weibull distribution, Journal of Pharmacy and Pharma- [65] J.F. Fitzgerald, O.I. Corrigan, Mechanisms governing drug
cology 24 (1972) 979–981. release from poly α-hydroxy aliphatic esters, diltiazem base
[50] G.K. Vudathala, J.A. Rogers, Dissolution of fludrocortisones release from poly-lactide-co-glycolide delivery systems, in: M.
from phospholipid coprecipitates, Journal of Pharmaceutical A. El-Nokaly, D.M. Piatt, B.A. Charpentier (Eds.), Polymeric
Sciences 82 (1992) 282–286. Delivery Systems, Properties and Applications, ACS Sympo-
[51] A. Dokoumetzidis, V. Papadopoulou, P. Macheras, Analysis sium Series, vol. 520, American Chemical Society, Washington,
of dissolution data using modified versions of Noyes– 1993, pp. 311–326.
Whitney equation and the Weibull function, Pharmaceutical [66] J. Siepmann, N. Faisant, J. Akiki, J. Richard, J.P. Benoit,
Research 23 (2) (2006) 256–261. Effect of the size of biodegradable microparticles on drug
[52] H.B. Hopfenberg, Controlled release from erodible slabs, release: experiment and theory, Journal of Controlled Release
cylinders, and spheres, in: D.R. Paul, F.W. Harris (Eds.), Con- 96 (2004) 123–134.
trolled Release Polymeric Formulations, ACS Symposium [67] R. Wada, S.-H. Hyon, Y. Ikada, Kinetics of diffusion-
Series, vol. 33, American Chemical Society, Washington, 1976, mediated drug release enhanced by matrix degradation,
pp. 26–31. Journal of Controlled Release 37 (1995) 151–160.
[53] A.W. Hixson, J.H. Crowell, Dependence of reaction velocity [68] T. Ehtezazi, C. Washington, Controlled release of mac-
upon surface and agitation, Industrial and Engineering romolecules from PLA microspheres: using porous struc-
Chemistry Research 23 (1931) 923–931. ture topology, Journal of Controlled Release 68 (2000)
[54] S.K. El-Arini, H. Leuenberger, Dissolution properties of 361–372.
praziquantel–PVP systems, Pharmaceutica Acta Helvetiae [69] J. Siepmann, K. Elkharraz, F. Siepmann, D. Klose, How
73 (2) (1998) 89–94. autocatalysis accelerates drug release from PLGA-based
[55] D.O. Cooney, Effect of geometry on the dissolution of microparticles: a quantitative treatment, Biomacromolecules
pharmaceutical tablets and other solids: surface detachment 6 (2005) 2312–2319.
kinetics controlling, AIChE Journal 18 (2) (1972) 446–449. [70] R.P. Batycky, J. Hanes, R. Langer, D.A. Edwards, A
[56] M. Zhang, Z. Yang, L.L. Chow, C.H. Wang, Simulation of theoretical model of erosion and macromolecular drug release
drug release from biodegradable polymeric microspheres with from biodegrading microspheres, Journal of Pharmaceutical
bulk and surface erosions, Journal of Pharmaceutical Sciences Sciences 86 (12) (1997) 1464–1477.
92 (10) (2003) 2040–2056. [71] E. Shen, R. Pizsczek, B. Dziadul, B. Narasimhan, Microphase
[57] D. Larobina, G. Mensitieri, M.J. Kipper, B. Narasimhan, separation in bioerodible copolymers for Drug Delivery,
Mechanistic understanding of degradation in bioerodible Biomaterials 22 (2001) 201–210.
polymers for Drug Delivery, AIChE Journal 48 (12) (2002) [72] K. Zygourakis, Development and temporal evolution of
2960–2970. erosion fronts in bioerodible controlled release devices,
[58] J. Siepmann, N. Faisant, J. Benoit, A new mathematical model Chemical Engineering Science 45 (8) (1990) 2359–2366.
quantifying drug release from bioerodible microparticles using [73] K. Zygourakis, P.A. Markenscoff, Computer-aided design of
Monte Carlo simulations, Pharmaceutical Research 19 (12) bioerodible devices with optimal release characteristics: a cellu-
(2002) 1885–1893. lar automata approach, Biomaterials 17 (2) (1996) 125–135.
[59] R.S. Harland, C. Dubernet, J.-P. Benoit, N.A. Peppas, A [74] A. Gopferich, R. Langer, Modeling of polymer erosion,
model of dissolution-controlled, diffusional drug release from Macromolecules 26 (1993) 4105–4112.
non-swellable polymeric microspheres, Journal of Controlled [75] A. Gopferich, R. Langer, The influence of microstructure and
Release 7 (1988) 207–215. monomer properties on the erosion mechanism of a class of
[60] J. Heller, R.W. Baker, Theory and practice of controlled Drug polyanhydrides, Journal of Polymer Science. Part A, Polymer
Delivery from bioerodible polymers, in: W. Baker (Ed.), Chemistry 31 (1993) 2445–2458.
Controlled Release of Bioactive Materials, Academic Press, [76] A. Gopferich, R. Langer, Modeling of polymer erosion in
New York, 1980, pp. 1–18. three dimensions: rotationally symmetric devices, AIChE
[61] A.G. Thombre, K.J. Himmelstein, A simultaneous transport- Journal 41 (10) (1995) 2292–2299.
reaction model for controlled Drug Delivery from catalyzed [77] A. Gopferich, Mechanisms of polymer degradation and
bioerodible polymer matrices, AIChE Journal 31 (5) (1985) erosion, Biomaterials 17 (2) (1996) 103–114.
759–766. [78] A. Gopferich, R. Langer, Modeling monomer release from
[62] A. Charlier, B. Leclerc, G. Couarraze, Release of mifepristone bioerodible polymers, Journal of Controlled Release 33
from biodegradable matrices: experimental and theoretical (1995) 55–69.
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1323

[79] M.J. Kipper, B. Narasimhan, Molecular description of erosion and in vitro anti-tumoral activity, Journal of Controlled
phenomena in biodegradable polymers, Macromolecules 38 Release 83 (2002) 273–286.
(2005) 1989–1999. [94] P. Chattopadhyay, R.B. Gupta, Supercritical CO2 based
[80] C.C. Chu, N.D. Campbell, Scanning electron microscopic study production of magnetically responsive micro- and nanopar-
of the hydrolytic degradation of poly(glycolic acid_suture, ticles for drug targeting, Industrial and Engineering Chemistry
Journal of Biomedical Materials Research 16 (4) (1982) Research 41 (2002) 6049–6058.
417–430. [95] T.W. Randolph, A.D. Randolph, M. Mebes, S. Yeung, Sub-
[81] M.B. Sintzel, A. Merkli, C. Tabatabay, R. Gurny, Influence of micrometer-sized biodegradable particles of poly (L-lactic
irradiation sterilization on polymers used as drug carriers — a acid) via the gas antisolvent spray precipitation process,
review, Drug Development and Industrial Pharmacy 23 (9) Biotechnology Progress 9 (1993) 429–435.
(1997) 857–878. [96] D.J. Jarmer, C.S. Lengsfeld, T.W. Randolph, Manipulation of
[82] A.G. Hausberger, R.A. Kenley, P.P. DeLuca, Gamma particle size distribution of poly(L-lactic acid) nanoparticles
irradiation effects on molecular weight and in vitro degrada- with a jet-swirl nozzle during precipitation with a compressed
tion of poly(D,L-lactid-co-glycolide microparticles, Pharma- antisolvent, Journal of Supercritical Fluids 27 (2003)
ceutical Research 12 (6) (1995) 851–856. 317–336.
[83] C. Volland, M. Wolff, T. Kissel, The influence of terminal [97] J. Xie, C.H. Wang, Self-assembled biodegradable nanoparti-
gamma-sterilization on captopril containing poly(D,L-lactide- cles developed by direct dialysis for the delivery of paclitaxel,
co-glycolide) microspheres, Journal of Controlled Release 31 Pharmaceutical Research 22 (12) (2005) 2079–2090.
(1994) 293–305. [98] R.A. Roman, A. Naik, Y.N. Kalia, R.H. Guy, H. Fessi, Skin
[84] D. Mohr, M. Wolff, T. Kissel, Gamma irradiation for terminal penetration and distribution of polymeric nanoparticles,
sterilization of 17β-estradiol loaded poly(D,L-lactide-co- Journal of Controlled Release 99 (2004) 53–62.
glycolide) microparticles, Journal of Controlled Release 61 [99] M.P. Desai, V. Labhasetwar, G.L. Amidon, R.J. Levy, Gastro-
(1999) 203–217. intestinal uptake of biodegradable microparticles: effect of par-
[85] S. Yoshioka, Y. Aso, T. Otsuka, S. Kojima, The effect of ticle size, Pharmaceutical Research 13 (12) (1996) 1838–1845.
γ-irradiation on drug release from poly(lactide) microspheres, [100] L.B. Peppas, Recent advances on the use of biodegradable
Radiation Physics and Chemistry 2 (1995) 281–285. microparticles and nanoparticles in controlled drug delivery,
[86] S. Yoshioka, Y. Aso, S. Kojima, Drug release from poly(DL- International Journal of Pharmaceutics 116 (1995) 1–9.
lactide) microspheres controlled by γ-irradiation, Journal of [101] R. Gref, Y. Minamitake, M.T. Peracchia, V. Trubetskoy, V.
Controlled Release 37 (1995) 263–267. Torchilin, R. Langer, Biodegradable long-circulating poly-
[87] N. Faisant, J. Siepmann, P. Oury, V. Laffineur, E. Bruna, J. meric nanospheres, Science 263 (1994) 1600–1603.
Haffner, J.P. Benoit, The effect of gamma-irradiation on drug [102] D.B. Chen, T.Z. Yang, W.L. Lu, Q. Zhang, In vitro and in vivo
release from bioerodible microparticles: a quantitative treatment, study of two types of long-circulating solid lipid nanoparticles
International Journal of Pharmaceutics 242 (2002) 281–284. containing paclitaxel, Chemical and Pharmaceutical Bulletin
[88] N. Faisant, J. Siepmann, J. Richard, J.P. Benoit, Mathematical 49 (11) (2001) 1444–1447.
modeling of drug release from bioerodible microparticles: [103] L. Mu, S.S. Feng, Fabrication, characterization and in vitro
effect of gamma-irradiation, European Journal of Pharmaceu- release of paclitaxel (Taxol®) loaded poly (lactic-co-glycolic
tics and Biopharmaceutics 56 (2003) 271–279. acid) microspheres prepared by spray drying technique with
[89] L. Montanari, M. Costantini, E.C. Signoretti, L. Valvo, M. lipid/cholesterol emulsifiers, Journal of Controlled Release 76
Santucci, M. Bartolomei, P. Fattibene, S. Onori, A. Faucitano, (2001) 239–254.
B. Conti, I. Genta, Gamma irradiation effects on poly(DL- [104] K.E. Lee, B.K. Kim, S.H. Yuk, Biodegradable polymeric
lactide-co-glycolide) microspheres, Journal of Controlled nanospheres formed by temperature-induced phase transition
Release 56 (1998) 219–229. in a mixture of poly(lactide-co-glycolide) and poly(ethylene
[90] K.S. Soppimath, T.M. Aminabhavi, A.R. Kulkarni, W.E. oxide)–poly(propylene oxide)–poly(ethylene oxide) triblock
Rudzinski, Biodegradable polymeric nanoparticles as Drug copolymer, Biomacromolecules 3 (2002) 1115–1119.
Delivery devices (review), Journal of Controlled Release 70 [105] C. Fonseca, S. Simoes, R. Gaspar, Paclitaxel-loaded PLGA
(2001) 1–20. nanoparticles: preparation, physicochemical characterization
[91] J.P. Vasir, M.K. Reddy, V.D. Labhasetwar, Nanosystems in and in vitro anti-tumoral activity, Journal of Controlled
drug targeting: opportunities and challenges, Current Release 83 (2002) 273–286.
Nanoscience 1 (2005) 47–64. [106] L. Mu, S.S. Feng, Vitamin E TPGS used as emulsifier in the
[92] T. Niwa, H. Takeuchi, T. Hino, N. Kunou, Y. Kawashima, solvent evaporation extraction technique for fabrication of
Preparations of biodegradable nanospheres of water-soluble polymeric nanospheres for controlled release of paclitaxel
and insoluble drugs with D,L-lactide/glycolide copolymer by a (Taxol®), Journal of Controlled Release 80 (2002) 139–144.
novel spontaneous emulsification solvent diffusion method [107] L. Mu, S.S. Feng, PLGA/TPGS nanoparticles for controlled
and the drug release behavior, Journal of Controlled Release release of paclitaxel: effects of the emulsifier and drug loading
25 (1993) 89–98. ratio, Pharmaceutical Research 20 (11) (2003) 1864–1872.
[93] C. Fonseca, S. Simoes, R. Gaspar, Paclitaxel-loaded PLGA [108] L. Mu, S.S. Feng, A novel controlled release formulation for
nanoparticles: preparation, physiochemical characterization the anticancer drug paclitaxel (Taxol®): PLGA nanoparticles
1324 D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325

containing vitamin E TPGS, Journal of Controlled Release 86 [123] L.T. Baxter, R.K. Jain, Transport of fluid and macromolecules
(2003) 33–48. in tumors. I. Role of interstitial pressure and convection,
[109] Z. Zhang, S.S. Feng, Nanoparticles of poly(lactide)/vitamin E Microvascular Research 37 (1989) 77–104.
TPGS copolymer for cancer chemotherapy: synthesis, [124] M.F. Haller, W.M. Saltzman, Localized delivery of proteins in
formulation, characterization and in vitro release, Biomater- the brain: can transport be customized? Pharmaceutical
ials 27 (2006) 262–270. Research 15 (3) (1998) 377–385.
[110] Potineni, D.M. Lynn, R. Langer, M.M. Amiji, Poly(ethylene [125] R.G. Larson, L.E. Scriven, H.T. Davis, Percolation theory of
oxide)-modified poly(b-amino ester) nanoparticles as a pH- two phase flow in porous media, Chemical Engineering
sensitive biodegradable system for paclitaxel delivery, Journal Science 36 (1981) 57–73.
of Controlled Release 86 (2003) 223–234. [126] W.M. Saltzman, R. Langer, Transport rates of proteins in
[111] J.H. Kim, Y.S. Kim, S. Kim, J.H. Park, K. Kim, K. Choi, H. porous materials with known microgeometry, Biophysical
Chung, S.Y. Jeong, R.W. Park, I. Kim, I.C. Kwon, Hydro- Journal 55 (1989) 163–171.
phobically modified glycol chitosan nanoparticles as carriers [127] C. Nicholson, L. Tao, Hindered diffusion of high molecular
for paclitaxel, Journal of Controlled Release 111 (2006) weight compounds in brain extracellular microenvironment
228–234. measured with integrative optical imaging, Biophysical
[112] R.T. Liggins, H.M. Burt, Paclitaxel loaded poly(L-lactic acid) Journal 65 (1993) 2277–2290.
microspheres: properties of microspheres made with low [128] W.M. Saltzman, M.L. Radomsky, K.J. Whaley, R.A. Cone,
molecular weight polymers, International Journal of Pharma- Antibody diffusion in human cervical mucus, Biophysical
ceutics 222 (2001) 19–33. Journal 66 (1993) 508–515.
[113] R.T. Liggins, H.M. Burt, Paclitaxel loaded poly(L-lactic acid) [129] M. Stroh, W.R. Zipfel, R.M. Williams, W.W. Webb, W.M.
microspheres II: the effect of processing parameters on Saltzman, Diffusion of nerve growth factor in rat striatum as
microsphere morphology and drug release kinetics, Interna- determined by multiphoton microscopy, Biophysical Journal
tional Journal of Pharmaceutics 281 (2004) 103–106. 85 (2003) 581–588.
[114] M. Polakovic, T. Gorner, R. Gref, E. Dellacherie, Lidocaine [130] D.K. Binder, M.C. Papadopoulos, P.M. Haggie, A.S. Verk-
loaded biodegradable nanospheres II. Modelling of drug man, In vivo measurement of brain extracellular space
release, Journal of Controlled Release 60 (1999) 169–177. diffusion by cortical surface photobleaching, The Journal of
[115] C. Kang, S.P. Schwendeman, Determination of diffusion Neuroscience 24 (37) (2004) 8049–8056.
coefficient of a small hydrophobic probe in poly (lactide-co- [131] R.G. Thorne, S. Hrabětová, C. Nicholson, Diffusion of
glycolide) microparticles by laser scanning confocal micros- epidermal growth factor in rat brain extracellular space
copy, Macromolecules 36 (2003) 1324–1330. measured by integrative optical imaging, Journal of Neuro-
[116] S.R. Van Tomme, B.G. De Geest, K. Braeckmans, S.C. De physiology 92 (2004) 3471–3481.
Smedt, F. Siepmann, J. Siepmann, C.F. van Nostrum, W.E. [132] R.G. Thorne, C. Nicholson, In vivo diffusion analysis with
Hennick, Mobility of model proteins in hydrogels composed quantum dots and dextrans predicts the width of brain
of oppositely charged dextran microspheres studied by protein extracellular space, PNAS 103 (14) (2006) 5567–5572.
release and fluorescence recovery after photobleaching, [133] W.M. Saltzman, M.L. Radomsky, Drugs released from
Journal of Controlled Release 110 (1) (2005) 67–78. polymers: diffusion and elimination in brain tissue, Chemical
[117] Y. Zhang, C.C. Chu, Biodegradable dextran–polylactide Engineering Science 46 (10) (1991) 2429–2444.
hydrogel network and its controlled release of albumin, [134] L.K. Fung, M. Shin, B. Tyler, H. Brem, W.M. Saltzman,
Journal of Biomedical Materials Research 54 (1) (2000) 1–11. Chemotherapeutic drugs released from polymers: distribution
[118] H.M. Wong, J.J. Wang, C.H. Wang, In vitro sustained release of 1,3-bis(2-chloroethyl)-1-nitrosurea in the rat brain, Phar-
of human immunoglobulin G from biodegradable micro- maceutical Research 13 (5) (1996) 671–682.
spheres, Industrial and Engineering Chemistry Research 40 [135] J.F. Strasser, L.K. Fung, S. Eller, S.A. Grossman, W.M.
(2001) 933–948. Saltzman, Distribution of 1,3-bis(2-chloroethyl)-1-nitrosurea
[119] T.H. Lee, F.J. Wang, C.H. Wang, Double-walled microspheres and tracers in the rabbit brain after interstitial delivery by
for the sustained release of a highly water soluble drug: biodegradable polymer implants, The Journal of Pharmacology
characterization and irradiation studies, Journal of Controlled and Experimental Therapeutics 275 (3) (1995) 1647–1655.
Release 83 (2002) 437–452. [136] K.W. Fung, M.G. Ewend, A. Sills, E.P. Sipos, R. Thompson,
[120] J.F. Gross, A.S. Popel, Mathematical models of transport M. Watts, O.M. Colvin, H. Brem, W.M. Saltzman, Pharma-
phenomena in normal and neoplastic tissue, in: H.I. Peterson cokinetics of interstitial delivery of carmustine, 4-hydroper-
(Ed.), Tumor Blood Circulation, CRC Press, Boca Raton, oxycyclophosphadime, and paclitaxel from a biodegradable
1979, pp. 169–183. polymer implant in the monkey brain, Cancer Research 58
[121] E.H. Starling, On the absorption of fluids from the connective (1998) 672–684.
tissue spaces, Journal of Physiology 19 (1986) 312–326. [137] A.B. Fleming, W.M. Saltzman, Pharmacokinetics of the
[122] F.E. Curry, Mechanics and thermodynamics of transcapillary carmustine implant, Clinical Pharmacokinetics 41 (6) (2002)
exchange, in: E.M. Renkin, C.C Michel (Eds.), Handbook of 403–419.
Physiology, Section 2: the Cardiovascular System), Amer. [138] H.J. Reulen, R. Graham, M. Spatz, I. Klatzo, Role of
Physiol. Soc., Bethesda, 1984, pp. 309–374. pressure gradients and bulk flow in dynamics of vasogenic
D.Y. Arifin et al. / Advanced Drug Delivery Reviews 58 (2006) 1274–1325 1325

brain edema, Journal of Neurosurgery 46 (1) (1977) [148] C.G. Lee, Y.-C. Fu, C.H. Wang, Simulation of gentamicin
24–35. delivery for the local treatment of osteomyelitis, Biotechnol-
[139] T. Nagashima, B. Horwitz, S.I. Rapoport, A mathematical ogy and Bioengineering 91 (5) (2005) 622–635.
model for vasogenic brain edema, Advances in Neurology 52 [149] D. Qin, G. Ou, H. Mo, Y. Song, G. Kang, Y. Hu, X. Gu,
(1990) 317–325. Improved efficacy of chemotherapy for glioblastoma by
[140] T. Nagashima, T. Shirakuni, S.I. Rapoport, A two-dimen- radiation induced opening of blood–brain barrier: clinical
sional, finite element analysis of vasogenic brain edema, results, International Journal of Radiation Oncology, Biology,
Neurologia Medico-Chirurgica (Tokyo) 30 (1990) 1–9. Physics 51 (2001) 959–962.
[141] S. Kalyanasundaram, V.D. Calhoun, K.W. Leong, A finite [150] M.V. Vulpen, H.B. Kal, M.J.B. Taphoorn, S.Y. El Sharouni,
element model for predicting the distribution of drugs de- Changes in blood–brain barrier permeability induced by
livered intracranially to the brain, American Journal of radiotherapy: implications for timing of chemotherapy?
Physiology—Regulatory, Integrative and Comparative Phys- (Review), Oncology Reports 9 (2002) 683–688.
iology 273 (1997) 1810–1821. [151] C.A. Znati, M. Rosenstein, T.D. McKee, E. Brown, D. Turner,
[142] C.H. Wang, J. Li, Three-dimensional simulation of IgG W.D. Bloomer, S. Watkins, R.K. Jain, Y. Boucher, Irradiation
delivery to tumors, Chemical Engineering Science 53 (20) reduces interstitial fluid transport and increases the collagen
(1998) 3579–3600. content in tumors, Clinical Cancer Research 9 (2003)
[143] C.H. Wang, J. Li, C.S. Teo, T. Lee, The delivery of BCNU to 5508–5513.
brain tumors, Journal of Controlled Release 61 (1999) 21–41. [152] R.K. Jain, Normalization of tumor vasculature: an emerging
[144] C.S. Teo, W.H.K. Tan, T. Lee, C.H. Wang, Transient concept in antiangiogenic therapy, Science 307 (2005) 58–62.
interstitial fluid flow in brain tumors: effect on Drug Delivery, [153] P. Kunkel, U. Ulbricht, P. Bohlen, M.A. Brockmann, R.
Chemical Engineering Science 60 (2005) 4803–4821. Fillbrandt, D. Stavrou, M. Westphal, K. Lamszus, Inhibition
[145] W.H.K. Tan, F.J. Wang, T. Lee, C.H. Wang, Computer of glioma angiogenesis and growth in vivo by systemic
simulation of the delivery of etanidazole to brain tumor from treatment with a monoclonal antibody against vascular
PLGA wafers: comparison between linear and double burst endothelial growth factor receptor-2, Cancer Research 61
release systems, Biotechnology and Bioengineering 82 (3) (2001) 6624–6628.
(2003) 278–288. [154] R.T. Tong, Y. Boucher, S.V. Kozin, F. Winkler, D.J. Hicklin,
[146] W.H.K. Tan, T. Lee, C.H. Wang, Simulation of intratumoral R.K. Jain, Vascular normalization by vascular endothelial
release of etanidazole: effects of the size of surgical opening, growth factor receptor 2 blockade induces a pressure gradient
Journal of Pharmaceutical Sciences 92 (2003) 773–789. across the vasculature and improves drug penetration in
[147] Y.M.F. Goh, H.L. Kong, C.H. Wang, Simulation of the tumors, Cancer Research 64 (2004) 3731–3736.
delivery of doxorubicin to hepatoma, Pharmaceutical Re-
search 18 (6) (2001) 761–770.

You might also like