You are on page 1of 12

REVIEW

Rate- and Extent-Limiting Factors of Oral Drug Absorption: Theory


and Applications
KIYOHIKO SUGANO, KATSUHIDE TERADA

Department of Pharmaceutics, Faculty of Pharmaceutical Sciences, Toho University, Funabashi, Chiba 274-8510, Japan

Received 18 December 2014; revised 23 January 2015; accepted 23 January 2015


Published online 24 February 2015 in Wiley Online Library (wileyonlinelibrary.com). DOI 10.1002/jps.24391

ABSTRACT: The oral absorption of drugs has been represented by various concepts such as the absorption potential, the maximum
absorbable dose, the biopharmaceutics classification system, and in vitro–in vivo correlation. The aim of this article is to provide an
overview of the theoretical relationships between these concepts. It shows how a simple analytical solution for the fraction of a dose
absorbed (Fa equation) can offer a theoretical base to tie together the various concepts, and discusses how this solution relates to the rate-
limiting cases of oral drug absorption. The article introduces the Fa classification system as a framework in which all the above concepts
were included, and discusses its applications for food effect prediction, active pharmaceutical ingredient form selection, formulation
design, and biowaiver strategy.  C 2015 Wiley Periodicals, Inc. and the American Pharmacists Association J Pharm Sci 104:2777–2788,

2015
Keywords: intestinal absorption; salt selection; bioequivalence; biopharmaceutics classification system (BCS); in silico modeling; disso-
lution; food effects; formulation; gastrointestinal transit; oral absorption

INTRODUCTION provide a comprehensive overview of theoretical relationships


between these concepts. A brief history in this scientific area is
The theories of oral drug absorption have been actively investi-
shown in Table 1.
gated in the last three decades.1 Various concepts have been in-
troduced to represent the oral absorption of drugs, for example,
the absorption potential (AP),2 the maximum absorbable dose DIFFERENTIAL EQUATIONS OF ORAL DRUG
(MAD),3 the biopharmaceutics classification system (BCS),4 the ABSORPTION
developability classification system (DCS),5 and in vitro–in vivo
correlation (IVIVC).6 These concepts originate from the same The oral absorption of a drug can be expressed by the two dif-
differential equations that describe the dissolution and intesti- ferential equations that describe the dissolution and intestinal
nal membrane permeation of drugs. The aim of this article is to membrane permeation of a drug (except for supersaturation).
The dissolution of a drug is usually expressed by the Noyes–
Whitney equation introduced in 1897 (Eq. (1)).7,14 The intesti-
Abbreviations used: AP, absorption potential; API, active pharmaceutical in-
gredient; AUC, area under the curve; AUC BE, bioequivalence of AUC; BCS, bio-
nal membrane permeation of a drug is usually expressed by the
pharmaceutics classification system; BCS-BWS, BCS-based biowaiver scheme; first order equation (Eq. (2)). These are typically expressed as:
BE, bioequivalence; COAS, computational oral absorption simulation; Cmax BE,
bioequivalence of Cmax ; Cdissolv , concentration of a dissolved drug; Cdissolv,ss ,  
dX undissolv 2 1 X dissolv
concentration of a dissolved drug at the steady state; DCS, developability clas- = −kdiss Dose 3 X undissolv
3
1− (1)
sification system; DF, degree of flatness; DRL, dissolution rate limited; Deff , dt Sdissolv V
effective diffusion coefficient; Dn, dissolution number; Dncrit , critical dissolution
number; Do, dose number; Docrit , critical dose number; Dose, dose strength; EIE,
equivalent-in-effect; EIP, equivalence of independent parameter; Fa, fraction of
a dose absorbed; FaCS, Fa classification system; GI, gastrointestinal; IR, imme- dX abs
diate release; IVIVC, in vitro–in vivo correlation; MAD, maximum absorbable = kperm X dissolv (2)
dose; MDT, mean dissolution time; PBPK, physiologically based pharmacokinet- dt
ics; PL, permeability limited; PL-E, epithelial membrane permeability limited;
PL-U, unstirred water layer permeability limited; Pncrit , critical permeation
number; PUWL , unstirred water layer permeability; Papp , apparent permeabil-
3Deff Sdissolv
ity; Peff , effective intestinal membrane permeability; Pep , epithelial membrane kdiss = (3)
permeability of unbound drug molecules; PKs, pharmacokinetics; Pn, perme- rp2 D
ation number; S1I7, GI compartment model with one stomach compartment
and seven intestinal compartments; SL, solubility–permeability limited; SL-E,
solubility–epithelial membrane permeability limited; SL-U, solubility–unstirred
water layer permeability limited; SMF, safe margin factor; Sblank , solubility in a 2DF
blank buffer; Sdissolv , solubility in the GI tract (free and bile micelle bound drug kperm = Peff (4)
molecules); Sn, saturation number; Tabs , transit time through the absorption R
site in the GI tract; UWL, unstirred water layer; V, fluid volume in the GI tract;
VE, villi expansion factor; Xdissolv , dissolved amount; fu , free fraction at the where kdiss is the dissolution rate coefficient, kperm is the per-
surface of the intestinal epithelial membrane; in vitro T85% , time to reach 85%
dissolution in an in vitro dissolution test; kabs , absorption rate coefficient; kdiss , meation rate coefficient, Dose is the dose strength, Xundissolv is
dissolution rate coefficient; kel , elimination rate coefficient; kperm , permeation the undissolved amount, Xdissolv is the dissolved amount, Xabs
rate coefficient; rp , particle radius of a drug; T1/2 , elimination half-life. is the absorbed amount, Sdissolv is the solubility in the intesti-
Correspondence to: Kiyohiko Sugano (Telephone: +81-47-472-1494; Fax: +81-
47-472-1337; E-mail: kiyohiko.sugano@phar.toho-u.ac.jp) nal fluid (free and bile micelle bound drug molecules), Deff is
Journal of Pharmaceutical Sciences, Vol. 104, 2777–2788 (2015) the effective diffusion coefficient, Peff is the effective intesti-

C 2015 Wiley Periodicals, Inc. and the American Pharmacists Association nal membrane permeability (free and bile micelle bound drug

Sugano and Terada, JOURNAL OF PHARMACEUTICAL SCIENCES 104:2777–2788, 2015 2777


2778 REVIEW

Table 1. Brief History of Theoretical Biopharmaceutics oral absorption of the drug becomes permeability limited (PL).
The trivial answer for Eq. (2) is:
Year Events Reference

1897 Noyes–Whitney equation 7  


Fa = 1 − exp −kperm Tabs = 1 − exp (−Pn) (7)
1985 Absorption potential 2
1986 Mixed tank model 8
1993 Plug-low model 9 This equation is often used to correlate in vitro mem-
1995 Biopharmaceutics classification system 4 brane permeability values with Fa, for example, the Caco-2
1996 Maximum absorbable dose 3 cell assay,16,17 the Madin–Darby canine kidney cell assay,18
Compartment absorption transit model 10
and the parallel artificial membrane permeation assay.19–21 We
1999 Absorption-limiting step classification 11
2005 Biopharmaceutics drug disposition 12
can also introduce a dimensionless parameter for permeation
classification system (Pn = kperm × Tabs ).
2009 Fa equation 13
2010 Developablility classification system 5 Sequential First-Order Approximation
When the oral absorption of a drug can be represented as the
sequential first-order process of dissolution and permeation,
molecules), D is the true density, and rp is the particle radius the analytical solution becomes
of a drug. R is the radius of the gastrointestinal (GI) tract, DF
is the degree of flatness, and V is the fluid volume in the GI
kperm
tract. Eq. (1) is for monodispersed spherical particles smaller Fa = 1 − exp (−kdiss Tabs )
than 60 :m. To calculate the fraction of a dose absorbed (Fa), kperm − kdiss
Eqs. (1) and (2) have to be integrated simultaneously. However, kdiss   Pn
the exact analytical solution for general cases has not been dis- − exp −kperm Tabs = 1 − exp (−Dn)
kdiss − kperm Pn − Dn
covered. The quest for the answer(s) that leads to Fa is one of
the main themes of theoretical biopharmaceutics. Dn
− exp (−Pn) (8)
Dn − Pn

QUEST FOR ANSWERS: A BRIEF HISTORY


Solubility - Permeability Limited Case
A simple analytical solution can usually be derived from a
The above three trivial answers were simply derived from
differential equation(s) by applying the initial and boundary
the differential equations. However, the concentration gradi-
conditions for a special (limiting) case. Historically, this ap-
ent term [1−Xdissolv /(Sdissolv × V)] makes it difficult to solve the
proach was first applied to the dissolution and permeation
equations for the cases of great interest in the pharmaceu-
equations for the three absorption limiting cases, that is, the
tical sciences. As no trivial answer is provided for this case,
dissolution rate, permeability, and solubility–permeability lim-
pharmaceutical insights are required to solve the equations.
ited (SL) cases. An illustrative explanation of these limiting
The concept of solubility - permeability limited absorption first
cases is available elsewhere in the literature.15
emerged as the AP in 1985,2 and then eventually developed
to the concept of the MAD3 and BCS.4 When the dissolution
Trivial Answers rate of a drug is much faster than the permeation rate and
Dissolution Rate-Limited Cases the dose to solubility ratio (Dose/Sdissolv ) exceeds the intestinal
fluid volume (V), the concentration of a dissolved drug (Cdissolv =
In the case of dissolution rate-limited (DRL) absorption, Fa Xdissolv /V) in the GI tract reaches close to the equilibrium solu-
can be calculated by integrating the dissolution equation. By bility of the drug (Sdissolv ). In this case, Eq. (2) can be integrated
applying Xdissolv = 0 (perfect sink condition), we obtain by applying a constant value of Xdissolv = Sdissolv × V as:
  32   32
2 2 kperm Sdissolv VTabs MAD Pn
Fa = 1 − 1 − kdiss Tabs = 1 − 1 − Dn (5) Fa = = = (9)
3 3 Dose Dose Do

where Tabs is the transit time through the absorption site in


the GI tract. Tabs can be approximated as the small intestinal Dose
transit time for many cases. We can introduce a dimensionless Do = (10)
Sdissolv V
parameter for dissolution [dissolution number (Dn) = kdiss ×
Tabs ]. This equation can be approximated as a first-order pro-
cess, where Do is the dose number. This oral absorption pattern is
often referred as “solubility limited” in the literature. However,
Fa = 1 − exp (−kdiss Tabs ) = 1 − exp (−Dn) (6) to explicitly recognize the role of permeability, we refer it as
“solubility - permeability limited” in this article. In the AP, kperm
is represented by the lipophilicity of a drug. The maximum
Permeability Limited Cases
absorbable dose is defined as MAD = kperm × Sdissolv × V ×
When the whole amount of a drug instantly and completely Tabs . BCS categorizes a drug by Do and Pn. The classification
dissolves after administration in the GI tract, the dissolution boundary of DCS is based on Fa = Pn/Do (Fig. 1). Therefore,
process will not become the rate-limiting step. In this case, the AP, MAD, BCS, and DCS are related to each other via Eq. (9).

Sugano and Terada, JOURNAL OF PHARMACEUTICAL SCIENCES 104:2777–2788, 2015 DOI 10.1002/jps.24391
REVIEW 2779

Pn Computational Numerical Integration

The wide spread of personal computers after the late 1980s pro-
BCS I BCS II vided new research tools for many scientists. Computational
10 numerical integration of differential equations (computer sim-
ulation) can provide answers to more general cases compared
with analytical solutions. Computer simulation is now used in
various research areas such as astronomy, meteorology, engi-
neering, economics, and so on. Pharmaceutical sciences have
been reaping many benefits from computer simulation as well.
Computational oral absorption simulations (COASs)13 have
0.1 1 10 been used since the mid-1980s. A simple mixed tank model
1 Do and a plug flow model were first employed to represent the
GI tract.8,9 The dissolution and permeation equations were nu-
merically integrated with these models. The compartment ab-
sorption transit model was then introduced in 1996 to repre-
sent the distribution of a drug in the GI tract.10,11,24 The other
multicompartment models such as the gastrointestinal transit
BCS III BCS IV absorption model25 and the advanced dissolution absorption
and metabolism model26 were later introduced. In most cases,
0.1
the stomach and the small intestine are expressed as one and
Figure 1. The BCS plane (Pn–Do plane) and DCS lines. The solid and seven compartments, respectively [the GI compartment model
gray lines corresponds to Fa = Pn/Do = 0.3 and 0.7 for SL cases. with one stomach compartment and seven intestinal compart-
ments (S1I7) model]. In each GI compartment, the dissolution
and permeation processes of a drug are simulated. The GI tran-
Approximate Fa Equation for General Cases sit of polydispersed particles between the compartments can be
In 2009, by applying the steady-state approximation, an ap- simultaneously simulated.13 By using a S1I7 model, various
proximate open analytical solution for Fa was discovered.13,22 scenarios of oral drug absorption can be simulated, such as the
This equation (the Fa equation, Eq. (11)) captures the essential effects of pH change in the GI tract.27
processes of oral drug absorption (except for supersaturation). One of the well-known disadvantages of computational sim-
ulation is that it might distract us from understanding the
  deep and often abstracted nature.28 Therefore, even in this age
1 of massive computing power, analytical solutions will not lose
Fa = 1 − exp − 1 Do if Do < 1, set Do = 1.
Dn
+ Pn their importance. Computer simulation and analytical solution
are complementary to each other. Because of its complexity,
1
≈ for Fa < 0.7 (11) COAS is often used as a black box. However, commercial pro-
1
Dn
+ Do
Pn grams have not been sophisticated enough for nonexpert users.
Therefore, a simple analytical solution and/or a classification
By taking the limit, the Fa equation can be rearranged to system should be used alongside with COAS to understand the
Eqs. (6), (7) and (9). When comparing Eq. (11) with Eq. (7), we essence of oral drug absorption. Commercial COAS programs
can define the absorption rate coefficient (kabs ) as: are often used side by side with a physiologically based phar-
macokinetics (PBPK) model.29,30 Pharmacokinetic (PK) scien-
1 1 Do tists sometimes use oral absorption and formulation modules in
= + (12)
kabs kdiss kperm software, whereas formulation scientists merely use the PBPK
module on its own. A good collaboration between the biophar-
Some important aspects of Fa can be directly derived by the maceutics and PK scientists is vital for an appropriate use of
Fa equation. COAS. Even though the S1I7 models seem to provide signif-
icant advantages, a simple one compartment model is suffi-
r Fa is determined by three factors, i.e., Dn, Do, and Pn. cient to predict Fa in many cases.31 The Fa equation and the
r Dn is independent from Do and Pn. seven compartment model give similar Fa values for nonsuper-
r Pn and Do are conjugated as Pn/Do. saturable cases.22 To simulate the plasma concentration–time
r Dn or Pn/Do can be the dominant factor when one largely profile of a drug, kabs can be used with a one-compartment PK
exceeds the other. model.
As COAS is a relatively new research tool for pharmaceutical
The Fa equation enables us to predict Fa from the in vitro scientists, there is little consensus on how to write the method
data available in drug discovery by a back-of-the-envelope section in a research paper. There are many indefinite factors
calculation.23 The saturation number (Sn) that represents the in in vivo GI physiology and there are significant discrepancies
ratio of Cdissolv at a steady state (Cdissolv,ss ) and Sdissolv can be between in vitro and in vivo conditions. The awareness of this
defined as: issue initiated an international project in Europe (OrBiTo).32
Therefore, an almost perfect prediction of PK profiles reported
Cdissolv,ss 1 so far in many articles is at least questionable (many of them
Sn = = (13)
Sdissolv 1 + DnDo
Pn
employ a commercial software). A good simulation practice13

DOI 10.1002/jps.24391 Sugano and Terada, JOURNAL OF PHARMACEUTICAL SCIENCES 104:2777–2788, 2015
2780 REVIEW

Figure 2. Fa classification system and its relationship with BCS, food effect (by bile micelles), API and formulation design, and biowaiver
strategy.

should be provided as a guidance for the healthy development FA CLASSIFICATION SYSTEM


of this emerging scientific area.
The Fa equation can provide a classification system for oral
drug absorption (referred as the Fa classification system (FaCS)
BIOPHARMACEUTICS CLASSIFICATION SYSTEM in this article). FaCS includes the concept of BCS, but it pro-
vides a more mutually exclusive and collectively exhaustive
The biopharmaceutics classification system has been widely classification system for oral drug absorption (Fig. 2; Table 2).
used in drug discovery and development. BCS was derived from On the basis of the Fa equation, the limiting factors of oral drug
Eqs. (1) and (2) by Amidon and coworkers4,9 in the mid-1990s. absorption can be classified as:
BCS classifies a drug substance by Do and Pn. A drug can be
spotted on the Do–Pn plane (Fig. 1). BCS was originally intro-
r Dissolution rate limited (DRL)
duced for biowaiver consideration. However, as a simple and
r Permeability limited (PL)
basic classification system, BCS has inspired many scientists
r Solubility - permeability limited (SL)
to investigate its applicability to various situations in drug dis-
covery and development. Roughly speaking, Do = 1 and Pn = 2
(Fa = 0.85–0.9) have been used as the criteria for solubility The basic concept of this classification system might have
and permeability, respectively. However, the rationale for the been implicitly suggested in the literature before the mid-1990s
high/low-permeability boundary has not been rigorously estab- or earlier. In 1999, Yu11 explicitly explained this concept with
lished. It might have been derived from the bioequivalence (BE) the criteria shown in Table 2. Takano et al.36 experimentally
criteria in the biowaiver scheme (discussed later), or indefi- confirmed this concept.
nitely derived from the tacit knowledge among the experts.33 Fa classification system and the criteria based on the Fa
The biopharmaceutics drug disposition classification system equation were introduced in 2010 and 2011, respectively.23,37 In
suggests that the extent of metabolism can be used as a surro- FaCS, PL and SL can be further classified by the rate-limiting
gate for Fa.12,34 factors of intestinal membrane permeation, i.e., the unstirred
The Developability classification system also classifies a water layer (UWL) and the epithelial membrane.38,39 The Peff
drug by Do and Pn; however, the criteria are based on Fa = can be expressed as:
Pn/Do rather than Do = 1 and Pn = 2 (Fig. 1).5 DCS can be
considered as an improved descendant of MAD that has also
been widely used in drug discovery and development (cf. Pn/Do PE
Peff = (14)
= MAD/Dose).35
1
PUWL
+ VEf1u Pep

Table 2. Absorption-Limiting Steps in Oral Drug Absorption

Criteria

Absorption-Limiting Steps Yua FaCS Fa Equation

Dissolution rate limited Dn < 1.1 Dn < Pn/Do Fa = 1−exp(−Dn)


Pn > 5.7
Pn/Do >> 1
Permeability limited Dn > 4.4 Pn < Dnb Fa = 1−exp(−Pn)
Pn < 5.7 Do < 1
Pn/Do >> 1
Solubility–permeability limited Dn > 4.4 Pn/Do < Dnb Fa = Pn/Do
Pn > 5.7 Do > 1
Pn/Do << 1

a
Ref. 11. Converted to corresponding dimensionless numbers.
b
Further classified by the rate-limiting step of the effective permeability.

Sugano and Terada, JOURNAL OF PHARMACEUTICAL SCIENCES 104:2777–2788, 2015 DOI 10.1002/jps.24391
REVIEW 2781

where PE is the plica expansion factor, VE is the villi expan- DRL Cases
sion factor, fu is the free fraction at the surface of the intestinal
In the case of DRL, particle size reduction would be effective
epithelial membrane, PUWL is the unstirred water layer per-
to increase Fa. However, as the dissolution rate increases, the
meability (free and bile micelle bound drug molecules), and
absorption regime changes from DRL to SL. The critical par-
Pep is the epithelial membrane permeability of unbound drug
ticle radius to become DRL can be calculated based on the Fa
molecules. Taken together, FaCS classifies drug products into
equation.23 The discriminate equation between DRL and SL is:
the five classes as shown in Figure 2.23,37 FaCS explicitly in-
cludes the basic concept of BCS (classification by Pn and Do).
Pn
In addition, FaCS also explicitly includes the dissolution per- Dn < (15)
spective as represented by the Dn in its classification scheme Do
(Fig. 2; Table 2). In contrast to commonly speculated inter-
When the dimensionless numbers are broken down, this
pretation, BCS cannot determine whether Fa becomes DRL
equation becomes:
or not.5 For example, when the solubility of a drug is 0.001
mg/mL, the dose is 0.1 mg, the particle size is 20 :m, Deff
3Deff Sdissolv Sdissolv VSI 2DFPeff
is 6.6 × 10−6 cm2 /s, and Peff is 6.6 × 10−4 cm/s (correspond- < (16)
ing to BCS class I; Dn = 0.7, Pn = 18.9, Do = 0.8), Fa be- rp D
2 Dose RSI
comes DRL. The third discriminate point of FaCS is related
to the permeability criterion. Unlike BCS, the permeability Sdissolv can be cancelled out from both sides of Eq. (16), sug-
boundary in FaCS is based on the rate-limiting factors of in- gesting that the discrimination boundary between DRL and SL
testinal membrane permeation, i.e., UWL or the epithelial does not depend on the solubility of a drug (for Do > 1). Even
membrane. though it is counterintuitive, this can be interpreted as follows;
when Sdissolv is low, the dissolution rate becomes slow, and at the
same time, the ceiling of Cdissolv (i.e., Sdissolv ) becomes low. On
the contrary, when the solubility is high, the dissolution rate
FOOD EFFECT PREDICTION becomes fast and the ceiling of Cdissolv becomes high. Therefore,
the tendency of Cdissolv to reach Sdissolv does not depend on Sdissolv
The Fa equation and FaCS have been applied to elucidate the (cf. the Sn does not depend on Sdissolv as well). By solving this
food effect by bile micelles.37,40 In the DRL cases, the solubi- equation for rp , we obtain23
lization of a drug by bile micelles would increase the disso-
lution rate, leading to a positive food effect. For example, a 
positive food effect (2.6-fold) was observed in ivermectin.41 In 3Deff DoseRSI
rp > (17)
the epithelial membrane permeability limited (PL-E) cases, a 2DFPeff VSI
negative food effect can be observed. Bile micelle binding re-
duces the unbound drug concentration at the epithelial sur- By using a conventional milling method, the mean particle
face and reduces the effective permeability of a drug.42–44 size can be reduced to 10 :m or less. Therefore, according to
In the solubility–epithelial membrane permeability limited Eq. (17), even for relatively high Peff cases (e.g., 5 ×10−4 cm/s),
(SL-E) cases, Fa would not be increased by bile micelles. The when the dose is more than 20 mg, Fa becomes SL.
drug molecules bound to bile micelles cannot permeate across
the epithelial cell membranes. Therefore, the flux across the ep- SL-E Cases
ithelial membrane would not be increased (flux = Sdissolv × Peff ࣈ Theoretically, the free drug concentration at the epithelial
(Sblank /fu ) × (VE × fu × Pep ) = Sblank × VE × Pep . Solubility in a membrane surface should be increased to achieve an increase
blank buffer (Sblank ) = Sdissolv × fu ). Pranlukast was suggested to in Fa for the SL-E cases. Supersaturable API forms and for-
be a typical example for this case.37 The solubility of pranlukast mulation technologies would be effective in this case. Super-
was increased by approximately ninefold in the fed-state simu- saturable API crystalline forms are salts,52 cocrystals,53 and
lated intestinal fluid compared with that in the fasted state sim- anhydrates.54,55 Solid dispersions and some of digestible self-
ulated intestinal fluid.45 However, area under the curve (AUC) emulsifying drug delivery systems can also be employed to in-
increased only 1.5-fold when taken with a food (cf. Fa% in the crease the free drug concentration at the epithelial membrane
fed state is 11% at 300 mg). In the solubility–unstirred water surface.56–58 It should be noted that an in vitro permeability as-
layer permeability limited (SL-U) cases, a positive food effect say often underestimates the permeability of a highly lipophilic
would be observed as the drug molecules bound to bile micelles drug (the octanol water distribution coefficient > 1.5).59,60 This
could permeate across UWL. Many drugs in this class show could result in misclassification of permeability (i.e., a BCS
positive food effects,46 for example, griseofulvin,47 danazol,48 class II drug can be falsely classified as a BCS class IV drug). As
and atovaquone.49 the physicochemical principles of solubility and passive mem-
brane permeability suggest,61,62 a genuine low-solubility/low-
permeability drug rarely exists.
ACTIVE PHARMACEUTICAL INGREDIENT FORM SL-U Cases
SELECTION AND FORMULATION DESIGN
In addition to the solubilization technologies for the SL-E cases,
Fa classification system can be used as a guidance for active formulation technologies that increase the apparent solubility
pharmaceutical ingredient (API) form selection and formula- would be effective for the SL-U cases, such as cyclodextrin.63
tion design, which are critically important for successful drug Interestingly, particle size reduction has been shown to be ef-
development.5,50,51 fective in increasing Fa for the SL-U cases.64,65 The particle

DOI 10.1002/jps.24391 Sugano and Terada, JOURNAL OF PHARMACEUTICAL SCIENCES 104:2777–2788, 2015
2782 REVIEW

drifting effect has been hypothesized to explain this the small intestine was suggested to be sufficient to justify
phenomenon.66 It is well known that small particles can drift biowaiver.73,74 Therefore, in the following discussion, the gas-
into the UWL.67,68 In this case, the apparent thickness of UWL tric emptying is assumed instant. The following discussion is
can be reduced. not intended to suggest any change of the current regulatory
biowaiver guidelines. A biowaiver strategy is not only used for
a regulatory purpose but also used for routine formulation re-
BIOEQUIVALENCE searches. An in vitro dissolution test is routinely used as a
Problem Statement surrogate of in vivo oral absorption in drug development.

The equivalence of the PKs between two drug products con- Fa Congruent Condition
taining the same drug molecule is called “BE”. Clinical BE is On the basis of the Fa equation, the congruence of Fa can be
defined as “20% acceptance range (80%–125%) for the 90% con- written as:
fidence interval of the ratio between test and reference least
square means after log transformation of the PK parameters Fa (Dn1 , Do1 , Pn1 ) = Fa (Dn2 , Do2 , Pn2 ) (18)
of interest, Cmax , and AUC.”69 If it is possible to allow waivers
of clinical BE studies based on the in vitro data (“biowaiver”), where the subscript number indicates each formulation (e.g., 1,
it would be beneficial for both drug developers and patients. reference formulation; 2, test formulation). For BE, Fa should
However, an in vitro dissolution test is not an all-round tool be equivalent at any time point (cf. Fa is essentially a function of
and only reflects a certain dimension of complicated in vivo time even though it usually refers to the total absorbed amount
situations. Therefore, the problem statement for a biowaiver at an infinite time after dosing). Usually, the equivalence of Fa
scheme is “when and under what conditions can waivers of is judged by the representative parameters for the rate and
clinical studies be allowed?”4 extent of Fa.
BCS-Based Biowaiver Scheme
Cmax1 (Dn1 , Do1 , Pn1 ) = Cmax2 (Dn2 , Do2 , Pn2 ) (19)
The BCS-based biowaiver scheme (BCS-BWS) for oral
immediate-release (IR) dosage has been employed by US FDA,
EMA, WHO, and many other regulatory agents. There are AUC1 (Dn1 , Do1 , Pn1 ) = AUC2 (Dn2 , Do2 , Pn2 ) (20)
many excellent review articles regarding BCS-BWS.33 The ra-
tionale to allow biowaivers for BCS Class I drugs was described There are two types of conditions to satisfy these equa-
as, tions: the equivalence of independent parameter (EIP) and
equivalent-in-effect (EIE) conditions. When Dn, Do, and Pn of
“When the in vivo dissolution of an IR oral dosage form is two formulations are identical, it is obvious that Cmax and AUC
rapid in relation to gastric emptying, the rate and extent of become equivalent (Fig. 3a).
drug absorption is likely to be independent of drug disso-
lution. Therefore, similar to oral solutions, demonstration Dn1 = Dn2 , Do1 = Do2 , Pn1 = Pn2 (21)
of in vivo bioequivalence may not be necessary as long
as the inactive ingredients used in the dosage form do not We can also define the EIE condition in which AUC and Cmax
significantly affect the absorption of the active ingredient. become insensitive to the independent parameters. In a certain
Thus, for BCS Class I (high solubility- high permeability) range, even when the independent parameters differed (i.e.,
drug substances, demonstration of rapid in vitro dissolu- Dn1 ࣔ Dn2 , Do1 ࣔ Do2 , and/or Pn1 ࣔ Pn2 ), it could result in
tion using the recommended test methods would provide the bioequivalence of Cmax (Cmax BE) and AUC (AUC BE). For
sufficient assurance of rapid in vivo dissolution, thereby example, when Dn >> Pn, Do <1, and Pn > 2, AUC becomes
ensuring human in vivo bioequivalence.”33 insensitive to the differences of these parameters within this
range (Fig. 3b).
In BCS-BWS, the criterion for solubility is based on the sol-
ubility of an API in the plain buffers of the entire pH range in Dn1 , Dn2 > Dncrit ; Do1 , Do2 < Docrit ; Pn1 , Pn2 > Pncrit (22)
the GI tract (e.g., pH 1.2–7.4). The criterion for permeability
is based on the permeability value for nearly complete absorp- where the subscript “crit” indicates the critical values to satisfy
tion (e.g., Fa > 0.85–0.90) (there are some variations in each the EIE conditions.
regulatory guideline).
Discriminate Points in FaCS for Bioequivalence
Discussion Based on FaCS
In FaCS, each discriminate point determines whether an EIE
In this section, a biowaiver strategy is discussed based on FaCS. condition is applicable or not.
As described above, the dissolution criteria in BCS-BWS is
based on the ratio of the dissolution rate and the gastric emp- Dn Discriminate Point. The first discriminate point in FaCS
tying rate. However, the gastric-emptying rate largely depends is Dn versus Pn/Do (Fig. 2; Table 2). We can define the critical
on the phase of the migrating motor complexes, which are the dissolution number (Dncrit ) as the lowest Dn value that provides
waves of the GI walls that sweep through the stomach and the the Cmax and AUC ratios of more than 0.8 against a solution
intestine in a regular cycle during a fasting state.70 In addi- formulation.75 The worst-case scenario within this range is that
tion, the gastric emptying half time can be as short as 4 min one of the two formulations shows instant dissolution, whereas
in a fasted state.71,72 Furthermore, the solubility of a drug in the other shows slowest dissolution in this range. When both

Sugano and Terada, JOURNAL OF PHARMACEUTICAL SCIENCES 104:2777–2788, 2015 DOI 10.1002/jps.24391
REVIEW 2783

1
(a) 0.9
(b)
0.8 Insensitive
Formulation 1 Formulation 2 0.7
0.6

Fa
0.5
0.4
Dn1 Pn1 Dn2 Pn2 0.3
Do1 Do2 0.2
0.1
0
0 1 2 3 4 5
Pn

Figure 3. An illustration of (a) EIP and (b) EIE conditions for biowaiver.

Dn1 and Dn2 are larger than Dncrit , the dissolution rate does rate. Therefore, it is safe (conservative) to use Figure 4a for the
not affect the bioequivalence of AUC and Cmax even when Do > 1 cases as well. Some of the transporter and paracellular
Dn1 ࣔ Dn2 (Dn1 and Dn2 become “EIE”). When Dn1 and/or substrates are mainly absorbed from the upper small intestine.
Dn2 are smaller than Dncrit , AUC and/or Cmax become sensitive In this case, Tabs can be shorter than 3.5 h, so that a larger kdiss
to Dn. In this case, the EIP, that is, Dn1 = Dn2, has to be proved. should be used as the dissolution rate criteria for AUC BE (cf.
For the Cmax BE, Dncrit can be roughly calculated as: kdiss = Dn/Tabs ).
A similar approach can be applied to take the gastric emp-
1 tying rate into account. The dissolution rate should be approx-
Dncrit > T1/2
×4 (23) imately fourfold faster than the gastric-emptying rate for Cmax
Do
Pn
+ ln2 ×Tabs
BE of a short half-life drug. When using the gastric-emptying
half-life of 10 min, in vivo T85% should be shorter than 5 min
The elimination half-life (T1/2 ) was included as it affects Cmax (Dncrit > 84).
[cf. Cmax ∝ ≈ 1/(1 + kel /kabs ); kel is the elimination rate coeffi-
cient; for kabs , see Eq. (12)]. The multiplying number of four is
originated from the 80% threshold of BE. Do Discriminate Point. The discriminate point next to Dn >
For the AUC BE of a low-permeability drug, Dncrit can be Dncrit is the Do discriminate point (Fig. 2). We can define the
roughly calculated to be 5 as: critical dose number (Docrit ) as the highest Do value that pro-
vides a sink condition in vivo. Docrit < 1 is used in the BCS-BWS
kperm (Tabs − MDT) 1 guidelines. To ensure a sink condition, it might be safer to use
=1− > 0.8 (24) Docrit < 0.3 as the EIE condition. Thirty percent of the satu-
kperm Tabs Dncrit
rated solubility is often used as a criterion for a sink condition.
where MDT is the mean dissolution time. The Tabs –MDT term When Do > Docrit , Do1 = Do2 is required to prove the BE (to-
suggests that MDT reduces the time available for drug absorp- gether with Pn1 = Pn2 as the oral absorption becomes SL in
tion. this case).
Numerical integration can be used to calculate Dncrit more
accurately. Figure 4a shows the relationship among Pn, elim- Pn Discriminate Point. The discriminate point next to Do <
ination T1/2 , and Dncrit obtained by numerically solving three Docrit is the Pn discriminate point (Fig. 2). We can define the
first-order equations (with kdiss , kperm , and kel ) that are sequen- critical permeation number (Pncrit ) as the lowest Pn value that
tially connected (simulated for Tabs = 3.5 h; Do < 1). provides the Cmax and AUC ratios of more than 0.8 against the
Figure 4a suggests that in the case of short half-life theoretically highest permeability case of Pn = 20. Pn = 20 is
drugs (<120 min.), the dissolution rate criteria for a high- calculated based on the highest permeability value observed in
permeability drug should be larger than that for a low- humans (Peff = 10 × 10−4 cm/s for glucose) and adjusted for
permeability drug because of the sensitivity of Cmax to Dncrit .75,76 the average molecular weight of modern drugs (400).81 Figure
Figure 4b is the cross-section of Figure 4a at T1/2 = 60 min. The 5 shows the relationship between Pncrit and T1/2 . For drugs
Dncrit lines for Cmax and AUC show different trends. For Cmax with T1/2 > 600 min, Pncrit is approximately 2, whereas for
BE, as the permeability of a drug becomes higher, a faster dis- drugs with T1/2 less than approximately 600 min, Pncrit is larger
solution is required (Table 3). This might be one of the reasons than 2. When Pn < Pncrit , Pn1 = Pn2 is required to prove the
for that some of the BCS class I drugs failed to show Cmax BE BE.
even when they complied with the rapid dissolution criteria.77
Several low-permeability drugs show clinical BE even when
Proofs for EIP and EIE Conditions
the dissolution profiles significantly differed among the formu-
lations, for example, cimetidine,78 fexofenadine,79 and so on. The strategies to prove the EIP and EIE conditions are summa-
The Cmax of a high-permeability and short half-life drug can be rized in Table 4 (conditions A–G). The EIP and EIE conditions
sensitive to the dissolution profile, for example, ibuprofen.80 discussed above are for in vivo, but not for in vitro. Therefore,
In the Do > 1 cases, the Dncrit plane shifts lower, suggesting a justification is required when using in vitro data to prove in
that AUC and Cmax become less sensitive to the dissolution vivo conditions, for example, IVIVC and a safe margin.

DOI 10.1002/jps.24391 Sugano and Terada, JOURNAL OF PHARMACEUTICAL SCIENCES 104:2777–2788, 2015
2784 REVIEW

Figure 4. The relationship among Dncrit , Pn, and T1/2 . (a) Dncrit plane satisfying both AUC BE and Cmax BE. (b) Cross-section at T1/2 =
60 min. The AUC and Cmax lines satisfy the equivalence of AUC and Cmax , respectively. The Dcrit values for Pn > 2 (Fa > 0.85–0.90) and
T1/2 < 10 h become higher than that for the Pn < 2 cases.

Table 3. Dissolution Criteria for Biowaiver

Dncrit T85% (min)a

Pn T1/2 AUC Cmax BEb Dncrit × 3c FaCS Based Current BCS-BWS

0.2 60 4.7 2.5 4.7 14 28 15d


0.2 600 4.7 4.3 4.7 14 28 15d
2 60 3.0 2.9 3.0 9.0 45 30
2 600 3.0 2.6 3.0 9.0 45 30
20 60 1.7 9.2 9.2 28 14 30
20 600 1.7 2.3 2.3 6.8 59 30

a
85% dissolution time.
b
Take the larger Dncrit .
c
SMF = 3 was used as an example.
d
EMA and WHO guidelines.

A. Dn > Dncrit r The constituents of the intestinal fluid in vivo such as bile
micelles usually increase the solubility of a drug.
r The excipients increase the solubility of a drug.
The conditions of a compendium in vitro dissolution test r The equilibrium solubility of a drug is smaller than the
could be dissimilar with those in vivo dissolution. Therefore,
supersaturated solubility.
an appropriate safe margin should be applied when using an
in vitro dissolution test as a surrogate for in vivo dissolution.
Usually, the dissolution rate criterion is described by the time Therefore, the equilibrium solubility of an API in plain
to reach 85% dissolution in an in vitro dissolution test (in vitro buffers should be smaller than the apparent solubility of a drug
T85% ). In vivo Dncrit can be converted to in vitro T85% as: in vivo from a drug product. Consequently, in vivo Do < Docrit
can be proved by using in vitro data (in vivo Do < in vitro Do <
Docrit ).
Tabs ln (1 − 0.85)
In vitro T85% = − (25)
Dnctir × SMF
C. Pn > Pncrit

where SMF is the safe margin factor. For AUC BE of a low-


permeability drug, Dncrit > 5 is required. When applying In the T1/2 more than approximately 600 min range, Fa >
SMF = 3, in vitro T85% becomes approximately 30 min (Ta- 0.85–0.90 (Pn > 2) would be appropriate to justify biowaiver
ble 3). For a high-permeability drug with T1/2 = 60 min, in vitro (Fig. 5). However, in the T1/2 less than approximately 600 min
T85% should be set 15 min for Cmax BE (Table 3). range, it does not guarantee Cmax BE. When Pn < Pncrit , a proof
for Pn1 = Pn2 in vivo is required [see sections (F) or (G) below].
When using a clinical Fa value for the permeability criterion,
B. Do < Docrit no safety margin would be required. However, when an in vitro
permeability assay is used as a surrogate, an appropriate safe
The following three reasons would justify the use of the in margin would be required.
vitro equilibrium solubility of an API in plain buffers to prove
Do < Docrit of a drug product in vivo. D. Dn1 = Dn2

Sugano and Terada, JOURNAL OF PHARMACEUTICAL SCIENCES 104:2777–2788, 2015 DOI 10.1002/jps.24391
REVIEW 2785

10 E. Do1 = Do2

9 The apparent solubility of a drug can be increased by the


excipients in the drug product. Therefore, the solubility of an
8 API cannot be used to prove Do1 = Do2 of drug products. A
Pn crit for 80% C max or AUC

dissolution test with a sink condition cannot be used to prove


7
Do1 = Do2 . A nonsink dissolution test with biorelevant media
6 may be used to prove Do1 = Do2 , especially for supersaturable
(b) (a1) APIs and formulations.52
5
UWL line F. PUWL1 = PUWL2
4

3 The permeability of a drug can be affected by the excipients


(c2) (a2) in the drug product. However, it is highly likely that most of the
2 excipients have little or no effect on PUWL (diffusion through a
water layer). It would be difficult to judge the rate-limiting step
1 AUC = 80% line in permeability from the in vivo Fa data, as Fa > 0.98 (Pn >
(c1) (Fa= 0.8) 4) is required to be UWL limited. In an in vitro assay such as
0 Caco-2, the apparent permeability (Papp ) should be more than
60 600 6000 50 × 10−6 cm/s to ensure UWL-limited permeation. Therefore,
EliminationT1/2 (min) a sufficient agitation is necessary for an in vitro assay.

Figure 5. The relationship between Pncrit and T1/2 . When Pn is above G. Pep1 = Pep2
the Pncrit line [(a1) and (a2) areas], Pn satisfies EIE for both AUC
and Cmax . Fa > 0.85–0.90 data can be used in the T1/2 more than
approximately 600 min range. When Pn is below the Pncrit line, EIP In contract to PUWL , it is well known that Pep can be altered
should be proved. PUWL1 = PUWL2 is required in the (b) area and Pep1 = by some excipients,84 especially for transporter and paracel-
Pep2 is required in the (c1) and (c2) areas. Fa < 0.8 can be used to prove lular substrates. Once these excipients are excluded from the
that the permeability is limited by the epithelial membrane, whereas formulation, Pep1 = Pep2 could be anticipated.
the other approach is required to diagnose the rate-limiting step for the
Fa > 0.8 cases [the (b) and (c2) areas]. Summary for Biowaiver Strategy
All in all, the discussion based on the Fa equation and FaCS
suggested that a drug product with a high permeability (Pn >
An IVIVC strategy can be applied to prove Dn1 = Dn2 .82 In 2), low solubility to dose ratio (Do < 0.3), and long elimination
IVIVC, in vitro dissolution profiles have to be proved to cor- T1/2 (>10 h) would be most suitable for biowaiver with the data
relate with those in vivo. When Dn << Pn/Do, Fa becomes of in vitro T85% < 30 min, in vitro Do < 0.3 and Fa > 0.85–0.9
sensitive only to Dn and insensitive to Do and Pn. As Pn and (Table 5). Therefore, the discussion based on FaCS basically
Do become EIE in this case, the last piece to prove the Fa supports the biowaiver for BCS Class I drugs except for the
congruent conditions is Dn1 = Dn2 (so that EIP of Dn). Usu- short elimination T1/2 cases. In the FaCS-based discussion, Pn
ally, for a controlled-release formulation, high-solubility/ high- plays two distinct roles: (1) the determination of dissolution
permeability drugs (Pn > 2 and Do < 1, therefore, Pn/Do > 2) is criteria (with T1/2 ), and (2) the proof of Pn equivalence. The
selected to ensure a good absorption from the whole GI tract.83 FaCS-based discussion would provide some hints about what
By applying the 80% dominancy rule, Dn < 0.4 is sufficient for should be carried out to improve biowaiver strategy and in vitro
Fa to be DRL. Dn = 0.4 corresponds to ca. 50% release at 5 h. tests in the future.85

Table 4. Equivalent-in-Effect and EIP Conditions Suggested by FaCS

Conditions Measures Criteria

Equivalent-in-effect (EIE)
(A) Dn > Dncrit In vitro dissolution test In vitro Dn > Dncrit × SMFa
(B) Do < 0.3 In vitro solubility test In vitro Do = Dose/(Sblank × V) < 0.3
(C) Pn > Pncrit Clinical Fa data Fa > Fa (Pncrit )b
In vitro permeability test Pn (from in vitro Papp )c > Pncrit × SMF
Equivalence of independent parameter (EIP)
(D) Dn1 = Dn2 IVIVC IVIVC criteria
(E) Do1 = Do2 Not known Not known
(F) PUWL1 = PUWL1 In vitro permeability PUWL d < VE × fu × Papp × SMF
(G) Pep1 = Pep2 Excipient restriction No excipient affecting Pep

a
SMF depends on the predictability of an in vitro test.
b
Fa = 0.85 and 0.90 corresponds to Pn = 1.9 and 2.3, respectively.
c
Pn calculated from Papp by Eqs. 4 and 14.
d
PUWL estimated from Deff (Ref. 36).

DOI 10.1002/jps.24391 Sugano and Terada, JOURNAL OF PHARMACEUTICAL SCIENCES 104:2777–2788, 2015
2786 REVIEW

Table 5. Biowaiver Strategy

Type of Equivalencea

Dn Do Pn FaCS BCSb Comments

EIE (A) EIE (B) EIE (C) PL-Uc + PL-Ed I Most suitable for biowaiver
EIE (A) EIE (B) EIP (F) PL-Ue I Likely to be PUWL1 = PUWL2
EIE (A) EIE (B) EIP (G) PL-Ef III Excipient restriction for Pep1 = Pep2
EIE (A) EIP (E) EIP (F) SL-U II Challenge to prove Do1 = Do2
EIE (A) EIP (E) EIP (G) SL-E IV Least suitable for biowaiver
EIP (D) EIEg EIEg DRL N/A IVIVC to prove Dn1 = Dn2

a
See test and Table 4 for the annotation of each alphabet. The color filling indicates the levels of challenge to prove the BE conditions; green, OK for biowaiver;
yellow, almost OK; orange, needs more scientific evidence; red, significant challenge required.
b
BCS classes approximately corresponding to each biowaiver strategy.
c
Area (a1) in Figure 5.
d
Area (a2) in Figure 5.
e
Area (b) in Figure 5.
f
Areas (c1) and (c2) in Figure 5.
g
Derived from Dn << Pn/Do.
N/A, not applicable.

CONCLUSION 8. Dressman JB, Fleisher D. 1986. Mixing-tank model for predicting


dissolution rate control or oral absorption. J Pharm Sci 75(2):109–
The theoretical investigations discussed here can significantly 116.
contribute to a better understanding of oral drug absorption. 9. Oh DM, Curl RL, Amidon GL. 1993. Estimating the fraction dose
They can provide guidance for API form selection, formulation absorbed from suspensions of poorly soluble compounds in humans: A
design, and biowaiver strategy. Although it is likely that com- mathematical model. Pharm Res 10(2):264–270.
putational simulations will be used more widely in the future, 10. Yu LX, Lipka E, Crison JR, Amidon GL. 1996. Transport ap-
analytical solutions and classification systems will continue to proaches to the biopharmaceutical design of oral drug delivery systems:
provide a way to understand the fundamentals of oral drug Prediction of intestinal absorption. Adv Drug Deliv Rev 19(3):359–
absorption. 376.
11. Yu LX. 1999. An integrated model for determining causes of poor
oral drug absorption. Pharm Res 16(12):1883–1887.
12. Wu C-Y, Benet LZ. 2005. Predicting drug disposition via application
ACKNOWLEDGMENTS of BCS: Transport/absorption/elimination interplay and development of
a biopharmaceutics drug disposition classification system. Pharm Res
The authors greatly appreciate Prof. Gordon Amidon, Prof. 22(1):11–23.
Shinji Yamashita, Dr. Ryusuke Takano, and Ms. Asami Ono 13. Sugano K. 2009. Introduction to computational oral absorption sim-
for their kind suggestions and discussions. The authors also ulation. Expert Opin Drug Metab Toxicol 5(3):259–293.
greatly appreciate the reviewers for their kind suggestions to 14. Dokoumetzidis A, Macheras P. 2006. A century of dissolution re-
improve the manuscript. search: From Noyes and Whitney to the biopharmaceutics classification
system. Int J Pharm 321(1–2):1–11.
15. Sugano K, Okazaki A, Sugimoto S, Tavornvipas S, Omura A, Mano
T. 2007. Solubility and dissolution profile assessment in drug discovery.
Drug Metab Pharmacokinet 22(4):225–254.
REFERENCES
16. Artursson P, Karlsson J. 1991. Correlation between oral absorption
1. Sugano K. 2012. Biopharmaceutics modeling and simulations: The- in humans and apparent drug permeability coefficients in human in-
ory, practice, methods, and applications. New Jersey: John Wiley & testinal epithelial Caco2 cells. Biochem Biophys Res Commun 175:880–
Sons, Inc. 885.
2. Dressman JB, Amidon GL, Fleisher D. 1985. Absorption potential: 17. Hidalgo IJ, Raub TJ, Borchardt RT. 1989. Characterization of
Estimating the fraction absorbed for orally administered compounds. the human colon carcinoma cell line (Caco-2) as a model sys-
J Pharm Sci 74(5):588–589. tem for intestinal epithelial permeability. Gastroenterology 96:736–
3. Johnson KC, Swindell AC. 1996. Guidance in the setting of drug 749.
particle size specifications to minimize variability in absorption. Pharm 18. Irvine JD, Takahashi L, Lockhart K, Cheong J, Tolan JW, Selick
Res 13(12):1795–1798. HE, Grove JR. 1999. MDCK (Madin–Darby canine kidney) cells: A
4. Amidon GL, Lennernas H, Shah VP, Crison JR. 1995. A theoretical tool for membrane permeability screening. J Pharm Sci 88(1):28–
basis for a biopharmaceutic drug classification: The correlation of in 33.
vitro drug product dissolution and in vivo bioavailability. Pharm Res 19. Kansy M, Senner F, Gubernator K. 1998. Physicochemical high
12(3):413–420. throughput screening: Parallel artificial membrane permeation as-
5. Butler JM, Dressman JB. 2010. The developability classification sys- say in the description of passive absorption processes. J Med Chem
tem: Application of biopharmaceutics concepts to formulation develop- 41(7):1007–1010.
ment. J Pharm Sci 99(12):4940–4954. 20. Wohnsland F, Faller B. 2001. High-throughput permeability pH
6. Young D, Devane JG, Butler J. 1997. In vitro–in vivo correlations. profile and high-throughput alkane/water log P with artificial mem-
Springer. branes. J Med Chem 44(6):923–930.
7. Noyes AA, Whitney WR. 1897. The rate of solution of solid substances 21. Sugano K, Hamada H, Machida M, Ushio H. 2001. High throughput
in their own solutions. J Am Chem Soc 19(12):930–934. prediction of oral absorption: Improvement of the composition of the

Sugano and Terada, JOURNAL OF PHARMACEUTICAL SCIENCES 104:2777–2788, 2015 DOI 10.1002/jps.24391
REVIEW 2787

lipid solution used in parallel artificial membrane permeation assay. J peaks in the plasma concentration–time profile. Pharm Res 10(6):879–
Biomol Screen 6(3):189–196. 883.
22. Sugano K. 2009. Fraction of dose absorbed calculation: Comparison 44. Kawai Y, Fujii Y, Tabata F, Ito J, Metsugi Y, Kameda A, Akimoto
between analytical solution based on one compartment steady state K, Takahashi M. 2011. Profiling and trend analysis of food effects on
approximation and dynamic seven compartment model. CBI J 9:75– oral drug absorption considering micelle interaction and solubilization
93. by bile micelles. Drug Metab Pharmacokinet 26(2):180–191.
23. Sugano K. 2011. Fraction of a dose absorbed estimation for struc- 45. Galia E, Nicolaides E, Horter D, Lobenberg R, Reppas C, Dressman
turally diverse low solubility compounds. Int J Pharm 405(1–2):79–89. JB. 1998. Evaluation of various dissolution media for predicting in vivo
24. Yu LX, Amidon GL. 1999. A compartmental absorption and transit performance of class I and II drugs. Pharm Res 15(5):698–705.
model for estimating oral drug absorption. Int J Pharm 186(2):119–125. 46. Singh BN. 2005. A quantitative approach to probe the dependence
25. Sawamoto T, Haruta S, Kurosaki Y, Higaki K, Kimura T. 1997. and correlation of food-effect with aqueous solubility, dose/solubility ra-
Prediction of the plasma concentration profiles of orally administered tio, and partition coefficient (log P) for orally active drugs administered
drugs in rats on the basis of gastrointestinal transit kinetics and ab- as immediate-release formulations. Drug Dev Res 65(2):55–75.
sorbability. J Pharm Pharmacol 49(4):450–457. 47. Ahmed IS, Aboul-Einien MH, Mohamed OH, Farid SF. 2008. Rel-
26. http://www.simcyp.com/. ative bioavailability of griseofulvin lyophilized dry emulsion tablet vs.
27. Sugano K. 2010. Computational oral absorption simulation of free immediate release tablet: A single-dose, randomized, open-label, six-
base drugs. Int J Pharm 398:73–82. period, crossover study in healthy adult volunteers in the fasted and
28. Atkins P. 2003. Galileo’s finger: The ten great ideas of science. fed states. Eur J Pharm Sci 35(3):219–225.
Oxford Univ Press, UK, pp 392. 48. Sunesen VH, Vedelsdal R, Kristensen HG, Christrup L, Muellertz
29. Rodgers T, Leahy D, Rowland M. 2005. Physiologically based phar- A. 2005. Effect of liquid volume and food intake on the absolute bioavail-
macokinetic modeling 1: Predicting the tissue distribution of moderate- ability of danazol, a poorly soluble drug. Eur J Pharm Sci 24(4):297–
to-strong bases. J Pharm Sci 94(6):1259–1276. 303.
30. Rodgers T, Rowland M. 2006. Physiologically based pharmacoki- 49. Rolan PE, Mercer AJ, Weatherley BC, Holdich T, Meire H, Peck RW,
netic modelling 2: Predicting the tissue distribution of acids, very weak Ridout G, Posner J. 1994. Examination of some factors responsible for a
bases, neutrals and zwitterions. J Pharm Sci 95(6):1238–1257. food-induced increase in absorption of atovaquone. Br J Clin Pharmacol
31. Takano R, Sugano K, Higashida A, Hayashi Y, Machida M, Aso 37(1):13–20.
Y, Yamashita S. 2006. Oral absorption of poorly water-soluble drugs: 50. Gardner CR, Walsh CT, Almarsson O. 2004. Drugs as materi-
Computer simulation of fraction absorbed in humans from a miniscale als: Valuing physical form in drug discovery. Nat Rev Drug Discov
dissolution test. Pharm Res 23(6):1144–1156. 3(11):926–934.
32. Lennernäs H, Aarons L, Augustijns P, Beato S, Bolger M, Box K, 51. Fiese EF. 2003. General pharmaceutics—The new physical phar-
Brewster M, Butler J, Dressman J, Holm R. 2014. Oral biopharma- macy. J Pharm Sci 92(7):1331–1342.
ceutics tools—Time for a new initiative—An introduction to the IMI 52. Takano R, Takata N, Saitoh R, Furumoto K, Higo S, Hayashi
project OrBiTo. Eur J Pharm Sci 57:292–299. Y, Machida M, Aso Y, Yamashita S. 2010. Quantitative analysis of
33. Yu LX, Amidon GL, Polli JE, Zhao H, Mehta MU, Conner DP, Shah the effect of supersaturation on in vivo drug absorption. Mol Pharm
VP, Lesko LJ, Chen M-L, Lee et al. 2002. Biopharmaceutics classifica- 7(5):1431–1440.
tion system: The scientific basis for biowaiver extensions. Pharm Res 53. Shiraki K, Takata N, Takano R, Hayashi Y, Terada K. 2008. Dissolu-
19(7):921–925. tion improvement and the mechanism of the improvement from cocrys-
34. Chen ML, Amidon GL, Benet LZ, Lennernas H, Yu LX. 2011. The tallization of poorly water-soluble compounds. Pharm Res 25(11):2581–
BCS, BDDCS, and regulatory guidances. Pharm Res 28(7):1774–1778. 2592.
35. Ding X, Rose JP, Van Gelder J. 2012. Developability assessment of 54. Shefter E, Higuchi T. 1963. Dissolution behavior of crystalline sol-
clinical drug products with maximum absorbable doses. Int J Pharm vated and nonsolvated forms of some pharmaceuticals. J Pharm Sci
427(2):260–269. 52(8):781–791.
36. Takano R, Furumoto K, Shiraki K, Takata N, Hayashi Y, Aso Y, 55. Kobayashi Y, Ito S, Itai S, Yamamoto K. 2000. Physicochemical
Yamashita S. 2008. Rate-limiting steps of oral absorption for poorly properties and bioavailability of carbamazepine polymorphs and dihy-
water-soluble drugs in dogs; prediction from a miniscale dissolution drate. Int J Pharm 193(2):137–146.
test and a physiologically-based computer simulation. Pharm Res 56. Friesen DT, Shanker R, Crew M, Smithey DT, Curatolo
25(10):2334–2344. WJ, Nightingale JA. 2008. Hydroxypropyl methylcellulose acetate
37. Sugano K, Kataoka M, Mathews CdC, Yamashita S. 2010. Predic- succinate-based spray-dried dispersions: An overview. Mol Pharm
tion of food effect by bile micelles on oral drug absorption considering 5(6):1003–1019.
free fraction in intestinal fluid. Eur J Pharm Sci 40:118–124. 57. Anby MU, Williams HD, McIntosh M, Benameur H, Edwards GA,
38. Amidon GE, Higuchi WI, Ho NFH. 1982. Theoretical and experi- Pouton CW, Porter CJ. 2012. Lipid digestion as a trigger for supersat-
mental studies of transport of micelle-solubilized solutes. J Pharm Sci uration: Evaluation of the impact of supersaturation stabilization on
71(1):77–84. the in vitro and in vivo performance of self-emulsifying drug delivery
39. Sugano K. 2009. Estimation of effective intestinal membrane per- systems. Mol Pharm 9(7):2063–2079.
meability considering bile micelle solubilisation. Int J Pharm 368(1– 58. Gao P, Rush BD, Pfund WP, Huang T, Bauer JM, Morozowich W,
2):116–122. Kuo MS, Hageman MJ. 2003. Development of a supersaturable SEDDS
40. Sugano K. 2009. Oral absorption simulation for low solubility com- (S-SEDDS) formulation of paclitaxel with improved oral bioavailability.
pounds. Chem Biodiversity 6(11):2014–2029. J Pharm Sci 92(12):2386–2398.
41. Ivermectine interview form. Accessed on February 11, 2015, at: 59. Wils P, Warnery A, Phung-Ba V, Legrain S, Scherman D. 1994.
http://www.info.pmda.go.jp/go/pack/6429008F1020 2 03/. High lipophilicity decreases drug transport across intestinal epithelial
42. Yamaguchi T, Ikeda C, Sekine Y. 1986. Intestinal absorption of a cells. J Pharmacol Exp Ther 269(2):654–658.
b-adrenergic blocking agent nadolol. I. Comparison of absorption be- 60. Krishna G, Chen K-J, Lin C-C, Nomeir AA. 2001. Permeability of
havior of nadolol with those of other b-blocking agents in rats. Chem lipophilic compounds in drug discovery using in-vitro human absorp-
Pharm Bull 34(8):3362–3369. tion model, Caco-2. Int J Pharm 222(1):77–89.
43. Lennernaes H, Regaardh CG. 1993. Evidence for an interaction 61. Yalkowsky SH, Valvani SC. 1980. Solubility and partitioning.
between the b-blocker pafenolol and bile salts in the intestinal lu- I: Solubility of nonelectrolytes in water. J Pharm Sci 69(8):912–
men of the rat leading to dose-dependent oral absorption and double 922.

DOI 10.1002/jps.24391 Sugano and Terada, JOURNAL OF PHARMACEUTICAL SCIENCES 104:2777–2788, 2015
2788 REVIEW

62. Diamond JM, Katz Y. 1974. Interpretation of nonelectrolyte parti- 75. Kortejärvi H, Urtti A, Yliperttula M. 2007. Pharmacokinetic simu-
tion coefficients between dimyristoyl lecithin and water. J Membr Biol lation of biowaiver criteria: The effects of gastric emptying, dissolution,
17(2):121–154. absorption and elimination rates. Eur J Pharm Sci 30(2):155–166.
63. Brewster ME, Noppe M, Peeters J, Loftsson T. 2007. Effect of the 76. Tsume Y, Amidon GL. 2010. The biowaiver extension for BCS class
unstirred water layer on permeability enhancement by hydrophilic cy- III drugs: The effect of dissolution rate on the bioequivalence of BCS
clodextrins. Int J Pharm 342(1–2):250–253. class III immediate-release drugs predicted by computer simulation.
64. Liversidge GG, Cundy KC. 1995. Particle size reduction for im- Mol Pharm 7(4):1235–1243.
provement of oral bioavailability of hydrophobic drugs: I. Absolute oral 77. Ramirez E, Laosa O, Guerra P, Duque B, Mosquera B, Borobia AM,
bioavailability of nanocrystalline danazol in beagle dogs. Int J Pharm Lei SH, Carcas AJ, Frias J, Velicky M, Bradley DF, Tam KY, Dryfe RA.
125(1):91–97. 2010. Acceptability and characteristics of 124 human bioequivalence
65. Jinno J-I, Kamada N, Miyake M, Yamada K, Mukai T, Odomi M, studies with active substances classified according to the Biopharma-
Toguchi H, Liversidge GG, Higaki K, Kimura T. 2006. Effect of particle ceutic Classification System. In situ artificial membrane permeation
size reduction on dissolution and oral absorption of a poorly water- assay under hydrodynamic control: Permeability-pH profiles of war-
soluble drug, cilostazol, in beagle dogs. J Control Release 111(1–2):56– farin and verapamil. Br J Clin Pharmacol 70(5):694–702.
64. 78. Jantratid E, Prakongpan S, Amidon GL, Dressman JB. 2006.
66. Sugano K. 2010. Possible reduction of effective thickness of intesti- Feasibility of biowaiver extension to biopharmaceutics classifica-
nal unstirred water layer by particle drifting effect. Int J Pharm 387(1– tion system class III drug products. Clin Pharmacokinet 45(4):385–
2):103–109. 399.
67. Norris DA, Puri N, Sinko PJ. 1998. The effect of physical barriers 79. Ono A, Sugano K. 2014. Application of the BCS biowaiver ap-
and properties on the oral absorption of particulates. Adv Drug Deliv proach to assessing bioequivalence of orally disintegrating tablets
Rev 34(2,3):135–154. with immediate release formulations. Eur J Pharm Sci 64:37–
68. Doyle-McCullough M, Smyth SH, Moyes SM, Carr KE. 2007. Fac- 43.
tors influencing intestinal microparticle uptake in vivo. Int J Pharm 80. Bramlage P, Goldis A. 2008. Bioequivalence study of three ibupro-
335(1–2):79–89. fen formulations after single dose administration in healthy volunteers.
69. Garcı́a-Arieta A, Gordon J. 2012. Bioequivalence requirements in BMC Pharmacol 8(1):18.
the European Union: Critical discussion. AAPS J 14(4):738–748. 81. Dahlgren D, Roos C, Sjögren E, Lennernäs H. 2015. Direct in vivo
70. Hansen MB. 2002. Small intestinal manometry. Physiol Res human intestinal permeability (Peff ) determined with different clinical
51(6):541–556. perfusion and intubation methods. J Pharm Sci.[Epub ahead of print]
71. Yamashita S, Kataoka M, Higashino H, Sakuma S, Sakamoto T, 82. Extended release oral dosage forms: Development, eval-
Uchimaru H, Tsukikawa H, Shiramoto M, Uchiyama H, Tachiki H. uation, and application of in vitro/in vivo correlations.
2013. Measurement of drug concentration in the stomach after intra- Accessed on January 23, 2015, at: http://www.fda.gov/
gastric administration of drug solution to healthy volunteers: Anal- downloads/drugs/guidancecomplianceregulatoryinformation/
ysis of intragastric fluid dynamics and drug absorption. Pharm Res guidances/ucm070239.pdf.
30(4):951–958. 83. Thombre AG. 2005. Assessment of the feasibility of oral controlled
72. Davis SS, Hardy JG, Fara JW. 1986. Transit of pharmaceutical release in an exploratory development setting. Drug Discov Today
dosage forms through the small intestine. Gut 27(8):886–892. 10(17):1159–1166.
73. Yazdanian M, Briggs K, Jankovsky C, Hawi A. 2004. The “high 84. Rege BD, Yu LX, Hussain AS, Polli JE. 2001. Effect of common
solubility” definition of the current FDA guidance on biopharmaceutical excipients on Caco-2 transport of low-permeability drugs. J Pharm Sci
classification system may be too strict for acidic drugs. Pharm Res 90(11):1776–1786.
21(2):293–299. 85. Tsume Y, Mudie DM, Langguth P, Amidon GE, Amidon GL. 2014.
74. Rinaki E, Dokoumetzidis A, Valsami G, Macheras P. 2004. Identifi- The biopharmaceutics classification system: Subclasses for in vivo pre-
cation of biowaivers among class II drugs: Theoretical justification and dictive dissolution (IPD) methodology and IVIVC. Eur J Pharm Sci
practical examples. Pharm Res 21(9):1567–1572. 57:152–163.

Sugano and Terada, JOURNAL OF PHARMACEUTICAL SCIENCES 104:2777–2788, 2015 DOI 10.1002/jps.24391

You might also like