throbber
Pharmaceutical Research, Vol. 26, No. 6, June 2009 (# 2009)
`DOI: 10.1007/s11095-008-9822-x
`
`Expert Review
`
`The Science of USP 1 and 2 Dissolution: Present Challenges
`and Future Relevance
`
`Vivian Gray,1,7 Gregg Kelly,2 Min Xia,3 Chris Butler,4 Saji Thomas,5 and Stephen Mayock6
`
`Received September 4, 2008; accepted December 24, 2008; published online January 23, 2009
`Abstract. Since its inception, the dissolution test has come under increasing levels of scrutiny regarding its
`relevance, especially to the correlation of results to levels of drug in blood. The technique is discussed,
`limited to solid oral dosage forms, beginning with the scientific origins of the dissolution test, followed by
`a discussion of the roles of dissolution in product development, consistent batch manufacture (QC
`release), and stability testing. The ultimate role of dissolution testing, “to have the results correlated to in
`vivo results or in vivo in vitro correlation,” is reviewed. The recent debate on mechanical calibration
`versus performance testing using USP calibrator tablets is presented, followed by a discussion of
`variability and hydrodynamics of USP Apparatus 1 and Apparatus 2. Finally, the future of dissolution
`testing is discussed in terms of new initiatives in the industry such as quality by design (QbD), process
`analytical technology (PAT), and design of experiments (DOE).
`in vitro–in vivo correlation; quality by design;
`
`KEY WORDS: biorelevant methods; dissolution;
`variability.
`
`INTRODUCTION
`
`This paper explores the advantages and disadvantages of
`the current methodology in light of recent challenges. While
`acknowledging its limitations, a case is made that the current
`dissolution test for drug product performance has value.
`The scope of this paper includes information on current
`issues, but it is not a tutorial on dissolution testing. The focus
`is on USP Apparatus 1 (baskets) and 2 (paddles) because
`these two systems constitute the bulk of dissolution testing in
`the pharmaceutical industry (1).
`The paper is organized by contemporary dissolution topics.
`Presented first is a description of the current challenges the paper
`will address. The challenges generally are divided into two
`classes, biorelevance and variability. Challenges covered by each
`subheading are discussed, followed by a brief section on the
`origin of the method procedure governed by United States
`Pharmacopeia chapter on Dissolution <711> (1). The intent is
`to demonstrate the scientific basis of current industry practice.
`Then, a review of dissolution by application exposes both the
`
`1 V. A. Gray Consulting, Inc., 9 Yorkridge Trail, Hockessin, Delaware
`19707, USA.
`2 Analytical R & D, Pfizer Global R & D, Groton, Connecticut, USA.
`3 Analytical Development, Vertex Pharmaceuticals, Cambridge, Mas-
`sachusetts, USA.
`4 Validation and CAPA, Ortho Clinical Diagnostics, Rochester, New
`York, USA.
`5 QC Lab Operations, Par Pharmaceutical, Spring Valley, New York,
`USA.
`6 Analytical Services, Catalent Pharma Solutions, Research Triangle
`Park, San Diego, North Carolina, USA.
`7 To whom correspondence should be addressed. (e-mail: vagray@rcn.
`com)
`
`value and limitations of the technique as an analytical tool.
`Application to formulation development, quality control, and in
`vitro–in vivo correlations (IVIVC) is covered. Next, variability
`inherent to dissolution testing is explored in the context of the
`challenges. A discussion is presented on calibration, including use
`of physical measurements and calibrator tablets, plus error
`associated with experimental conditions or analyst technique. It
`should be noted that recently the calibrator tablets were renamed
`by USP as Performance Verification Standards; however, since
`this is a new development, we have kept the term calibrator
`tablets throughout this paper. Hydrodynamics is the final section
`under dissolution method variability. The future of dissolution
`testing is discussed in sections on process analytical testing (PAT),
`design of experiments (DOE), and quality by design (QbD).
`Finally, the utility and future of the technique are summarized.
`
`Challenges
`
`Both biorelevance and technique variability are used to
`challenge the validity of dissolution testing. The basis for each
`challenge is presented below.
`The most significant challenge for many dissolution
`methods used as a nominal performance measure stems from
`the lack of biorelevance. Scientists have stated that develop-
`ing a dissolution method and setting associated specifications
`that are not linked to in vivo performance may limit the value
`of testing (2–7). It is not difficult to see that the vortex in the
`current design of USP apparatus is not the same as in a
`churning stomach. The majority of dissolution testing is
`carried out in a simple salt medium at a particular pH. The
`gastrointestinal lumen is significantly different, containing a
`plethora of biomolecules and salts in a changing pH
`
`1289
`
`0724-8741/09/0600-1289/0 # 2009 Springer Science + Business Media, LLC
`
`ENDO - Ex. 2028
`Amneal v. Endo
`IPR2014-00360
`
`

`

`1290
`
`Gray et al.
`
`environment. A lack of a biorelevant (physiologically based)
`dissolution system and specification often leads to data that
`are disconnected from in vivo results (2,3,7,8). Examples are
`cited where the dissolution method is either overly or not
`sufficiently discriminating (2,7,9). Few cases have been found
`where the method is appropriately discriminating (10). A
`dissolution method that is developed solely as a quality
`control tool for manufacturing is much less desirable than one
`that has bearing on patient safety or efficacy. If measurements
`have no bearing on the pharmacokinetic impact, then testing
`is not controlling the most important aspect of performance
`(2,3,7,8). Calls for better method development with biorele-
`vant specifications are the result (2,3,6,7).
`Variability associated with dissolution testing is another
`area receiving a great deal of attention. Many studies
`demonstrate the source and extent of variability (2–4,8,11–
`14). These sources can be divided into four subsets. The first is
`the physical or mechanical setup of the test. Tolerances allowed
`in operating the apparatus are defined by the USP (1). The
`definitions are designed to allow the apparatus to function with
`acceptable method variability, but even when operating within
`these limits, different dissolution profiles for the same drug
`product may result. Other physical factors are not controlled
`by the USP description but have an effect. Among the
`parameters in this class are shaft or basket wobble, vessel/shaft
`tilt, shaft centering, shaft height in vessel, and rotational speed
`(3,13). Vessel roundness, surface uniformity, or other hydrody-
`into this class and impact results (3–
`namic effects fall
`5,11,14,15). Even small changes in basket mesh size seem to
`have an influence on results (14). Another class of variability
`arises from operational differences. Parameters in this group
`are incidental vibration, the extent of degassing, inconsistent
`tablet placement in vessels, and inconsistent use of clips or
`sinkers (3,11–13). The third class of variability comes indirectly
`from performance differences in calibrator tablets that are real
`(8,16,17), operator induced, or from excipient deposition (18).
`As the name implies, calibrator tablets are used to verify
`overall system precision to qualify apparatus and control
`system variability. However, different disintegration mecha-
`nisms between calibrator and sample tablets are cited as a
`source of variability (3). Proposed remedies for calibrator
`tablet variability are mechanical calibration (3,13,19), project-
`specific manufacturer calibrator tablets possessing similar
`processing and mechanistic disintegration qualities (3,6), or
`non-USP apparatus (4,5,20). The fourth source of variability
`comes from manufacturing and is due to lot-to-lot or tablet-to-
`tablet processing or handling differences of the drug product
`(3,6,16). It includes particle size distribution and polymorph
`changes during drug substance manufacture. Changes in
`excipient characteristics are known to impact results (7). The
`variability from this cause is independent of the method but is
`reflected in the results. Sorting out the origin among all the
`potential sources of variability can be problematic.
`
`Scientific Origins of Dissolution
`
`Scientific Origins
`
`determining that the dosage would dissolve. To this day,
`dissolution is the only test that indicates if a dosage form will
`dissolve in the patient. The disintegration test was the first
`test designed to do this, but it has obvious limitations.
`Although a tablet or capsule can disintegrate into smaller
`particles,
`if it does not dissolve,
`it is not available to be
`absorbed in the small intestine.
`Dissolution, as a general dosage performance test, was
`primarily linked to changes in the drug product formulation
`and the critical process parameters that can affect dissolution.
`During the process validation of
`tablet manufacturing,
`dissolution testing is performed on tablets at
`the target
`hardness and at the high and low extremes.
`Dissolution is still a critical test to determine the effects of
`aging of the product on stability. Changes in tablet hardness,
`moisture, or other excipient changes can affect dissolution.
`Capsule cross-linking can have a significant effect on dissolution
`of samples on stability. In many respects, this continues to be the
`most compelling reason to have an effective dissolution test for
`testing a solid oral dosage product.
`Some of the basic aspects of the dissolution test have
`their origins in general conditions in the human body. The
`test is conducted at 37°C. The paddle or basket rotation is
`designed to produce reproducible hydrodynamics that can be
`consistent from lab to lab. The real physical purpose of the
`agitation is to remove the drug-saturated layer of dissolution
`from around the dosage and replace it with fresh medium
`without causing a significant physical change in the dosage.
`The use of a 900-mL volume was determined in order to be
`enough to establish sink conditions (at least three times
`saturation) for most active pharmaceutical ingredients. Dis-
`solution media were developed to mimic the pH of the gastro-
`intestinal tract. At one time, simulated intestinal fluid had a
`pH of 7.4. This was changed to a pH of 6.8 in the mid-90s,
`because it was determined that this more closely represents
`the intestinal pH (21).
`
`Dissolution Testing within the USP
`
`The basic dissolution test in USP chapter <711> Disso-
`lution describes the apparatus, the dissolution procedure, and
`product specifications. The old chapter <724> described
`Apparatus 3 through 7, while Apparatus 1 and 2 were
`described in chapter <711>. The newer editions of the USP
`have now combined Apparatus 1 through 4 in chapter <711>.
`The dissolution procedures have been harmonized in the
`pharmacopeias internationally, although there are some
`sections that remain unique to each pharmacopeia. The
`USP chapter <1088> describes the procedure for in vitro–in
`vivo evaluation of dosage forms, and chapter <1092> presents
`the development and validation of the dissolution procedure.
`The dissolution test has evolved over time and will continue
`to be improved as it is called upon to give more data that are
`relevant to dosage performance in the patient.
`
`The FDA and Dissolution Testing in History
`
`Initially, the dissolution test was used primarily as a
`formulation development tool and as a quality control test for
`
`The FDA has placed much importance on the dissolution
`test and reviews the USP monograph dissolution tests for
`consistency with the dissolution conditions in the approved
`
`

`

`USP 1 and 2 Dissolution: Present Challenges and Future Relevance
`
`1291
`
`product’s New Drug Application. Most solid oral dosage
`forms are required to have a dissolution test, and it is not
`uncommon to have a drug recall due to a failed dissolution
`test. Part of the approval process of an NDA solid oral
`dosage is the FDA evaluation of the dissolution method.
`Members of the FDA helped to develop the Biopharma-
`ceutical Classification System (BCS). This has led to the
`development of guidances related to dissolution testing that
`are available on the FDA website.
`
`Utility and Basic Goals of the Test—Formulation Development
`
`Formulators consider the dissolution test to be a very
`powerful tool. The test can be used to show the dependence
`of dissolution rate on the presence and concentration of
`certain excipients and on manufacturing variables. There is
`abundant literature on the use of dissolution as a comparative
`test that can show a formulation change. A select few will be
`highlighted in this section of the paper.
`As early as 1976, Khan and Rooke (22) described the effect
`of disintegration type upon the relationship between compres-
`sion force and dissolution efficiency. The discussion compared
`three common disintegrants, sodium carboxymethyl cellulose,
`sodium starch glycolate, and a cation-exchange resin to then less
`commonly known excipients as insoluble sodium carboxymethyl
`cellulose, casein formaldehyde, calcium carboxymethyl cellu-
`lose, and a cross-linked polyvinylpyrrolidone. They concluded
`that the disintegrant type has a pronounced effect on the
`dissolution rate. In this work, the paddle was used at 50 rpm with
`water as the medium. The study also cited a 1963 publication of
`Levy et al. (23) that correlated an increase in compression force
`with starch-containing formulations.
`Chowhan and Palagyi (24) explored the issues of
`hardness and the effect of the dissolution rate. This study
`showed how hardness was increased by partial moisture loss
`in compressed tablets. Several
`factors were investigated
`including type and percentage of excipient, water solubility,
`hygroscopicity of excipients or drug, and the influence of
`frequently used binders. They concluded that since the
`dissolution is related to moisture content of the granulation
`and the hardness of the tablets at the time of compression, the
`dissolution specification would ensure that the tablets meet the
`moisture and hardness requirements. It was recommended
`that the moisture content of the granulation and initial
`hardness be used as in-process controls. In this work, the
`paddle was used at 120 rpm with water or 7.4 phosphate buffer
`as the medium.
`In 1981 Taborsky-Urdinola et al. (25) published a paper
`that won an APHA Research award. The importance of the
`paper was the proof that packaging type and storage conditions
`in multiple and unit dose containers markedly affect the
`dissolution results of model Prednisone tablets. A conclusion
`was that relabeling repackaged tablets with the expiration date
`of the original container was invalid. The dissolution conditions
`were paddle at 50 rpm using water as the medium.
`Chowhan and Chi (26) continued his research and in
`1985 described the role of lubricants and their effect on
`dissolution results. Two lubricants, magnesium stearate and
`sodium stearyl fumarate, were compared under identical mixing
`conditions to determine drug–excipient interactions. The con-
`clusions were that sodium stearyl fumarate did not exhibit drug–
`
`excipient interactions, whereas magnesium stearate did exhibit
`significant drug–excipient interactions that adversely affected
`the disintegration time and dissolution rate.
`Changes in surface area and dissolution rate were illumi-
`nated by Sunada et al. (27) in 1989. The changes in surface area
`during the dissolution process were measured, and the relation-
`ship between the surface producing rate constant and the initial
`particle size of sieved samples was estimated. There was also the
`simulation of the dissolution process based on the changes in
`surface area and the surface producing rate constant. The
`paddle speed was 250 rpm in water.
`In the 1990s, the use of dissolution as an indicator of
`aging began. Chowhan (28) discussed the complexity of aging
`as related to selected factors other than the packaging and
`storage conditions. Factors such as the hygroscopicity of the
`superdisintegrants; method of disintegrant incorporation; gran-
`ulation moisture content; effect of high or low humidity on the
`type of disintegrant (e.g., dibasic calcium phosphate dihydrate
`and tribasic calcium phosphate); effect of the use of lactose,
`dextrose, or MMC; and gelatin shell cross-linking were all
`evaluated. It was concluded that guidelines calling for acceler-
`ated conditions could give a good indication of aging issues.
`Babu and Pandit (29) described how stability of gliben-
`clamide was enhanced by complexation with β-cyclodextrin.
`The dissolution rate was employed as an indicator of aging
`using the paddle at 100 rpm and pH 7.4 phosphate buffer.
`Dissolution rate was one of the important parameters
`measured when differentiating forms I and II (R, S) of
`propranolol hydrochloride. Bartolomei et al. (30) showed that
`dissolution rates of the two polymorphs were different using
`the paddle at 50 rpm with 0.1 N hydrochloric acid medium;
`the test was run at 20°C and 37°C.
`The effects of temperature and humidity on the physical
`properties of piroxicam tablets were shown by Sarisuta et al.
`(31). The tablets containing various fillers (lactose or
`mannitol) were studied after storage for 12 weeks at 40°C
`and 52% relative humidity (RH) and 40°C and 96% RH. The
`physical properties of
`the tablets were measured every
`2 weeks. Dissolution was measured using the paddle at
`50 rpm with simulated gastric fluid as the medium. The
`dissolution rate decreased from week to week, regardless of
`the filler used. It was explained that the decrease in dissolution
`was due to moisture sorption by the tablet ingredients, which
`led to the formation of a saturated solution of water-soluble
`substances. Consequently, crystal growth and swelling of
`polymeric material occurred. This yielded a continuous
`structure of larger crystals so the exposed surface area was
`significantly reduced, hence the dissolution rate decreased.
`In 1999 Rohrs et al. (32) showed the effect of croscar-
`mellose sodium disintegrant on delavirdine mesylate. In the
`presence of high humidity, the water presumably acted as a
`reaction medium and a plasticizer for croscarmellose sodium,
`facilitating protonation of the carbonyl sites on the disintegrant.
`The important finding is that this reaction could very well occur
`with any acid salt of a free base. A change in inter-particle
`bonding can explain the reduction in tablet deaggregation
`during dissolution. The dissolution was performed using the
`paddle at 50 rpm, and the medium was 0.05 M phosphate buffer
`at pH 6 with 0.6% sodium dodecylsulfate surfactant.
`The effect of powder substrate composition on the
`dissolution rate of methyclothiazide liquisolid compacts was
`
`

`

`1292
`
`Gray et al.
`
`illustrated by Spireas et al. (33). Dissolution rates were
`increased by optimizing carrier-to-coating ratios in methyclo-
`thiazide liquisolid tablets containing a 5% w/w drug solution
`in polyethylene glycol 400 with difference excipient ratios.
`The dissolution conditions used were the paddle at 50 rpm
`and a medium of 0.1 N hydrochloric acid.
`The influence of excipients, especially binders, on the
`dissolution rate of paracetamol tablet formulations was shown
`by Abebayo et al. (34). The effect of binders, namely
`breadfruit and cocoyam starch mucilage binders, was related
`to their surface tension and viscosity. The dissolution test
`used the basket at 50 rpm and pH 5.8 phosphate buffer.
`Numerous articles on the subject of extended-release
`formulation information show that there is variability in the
`dissolution rate with the change of matrix ingredients and
`ratios (35–38). In the references cited, as in many cases, the
`paddle was used in the dissolution test.
`In the last 30–40 years, the dissolution test with USP
`paddle and basket apparatus has been used extensively to
`provide information to the formulators regarding critical
`process variables. Only a limited amount of the literature is
`shown here as the literature is full of examples of in vitro
`release testing used to determine change in the formulation
`or manufacturing process. The power of the dissolution test is
`undisputed in assisting product development
`from early
`phases to monitoring stability.
`
`QC Testing for Batch Manufacturing Consistency
`and Specification Setting
`
`Product Batch Release
`
`The value of in vitro dissolution testing as a quality
`control tool is demonstrated by its long history of regulatory
`acceptance. Dissolution testing has been included in the USP
`since 1970 and continues to be an important test today as
`evidenced by the large number of monographs that include
`dissolution requirements (over 600 as of 2006) (39). This
`points to an important benefit for drug manufacturers—
`dissolution testing fulfills a regulatory requirement.
`Although the primary purpose of the dissolution test
`specification is to distinguish between acceptable and unaccept-
`able batches, it is also used as a measure of batch-to-batch
`consistency of the manufacturing process. In this case, the
`method may be developed to be sensitive to manufacturing
`variables determined to influence drug release (40).
`
`Stability and Shelf Life
`
`Dissolution testing is also the primary method used to
`demonstrate stability of drug product performance through-
`out its shelf life. Although not specified by name in the
`guidance, dissolution testing fulfills the ICH Q1A (R2)
`requirement that stability studies include testing of drug
`influence “product quality, safety and/or
`attributes that
`efficacy” and that are susceptible to change over time (41).
`Dissolution has proven to be a valuable tool to indicate
`changes in such characteristics as crystallinity (42), glass
`transition temperature and pore structure of polymeric
`excipients (43), polymorphism (44), gelatin capsule cross-
`
`linking (45), and moisture content (32). This information can
`be used to make informed decisions on selection of formula-
`tion, manufacturing process, and packaging.
`
`Setting Specifications, Establishing Product History,
`Post-Approval Manufacturing Changes
`
`The dissolution test plays an important role in setting
`drug product specifications. The dissolution specification
`includes the specific dissolution procedure as well as accep-
`tance criteria;
`it is intended to show that manufactured
`product is bioequivalent to pivotal clinical lots and confirm
`it was manufactured within acceptable values of critical
`manufacturing variables. Conformance to the acceptance
`criteria can be used to determine stability of the drug and to
`justify waiver of additional clinical studies following certain
`post-approval changes (46–48). Following approval, dissolu-
`tion data for manufactured lots form a product history from
`which the “true” capability and variability of the process can
`be derived. This information may be used as justification for
`revised acceptance criteria (49).
`Recently, the dissolution test has been criticized for not
`being predictive of bioavailability because methods do not
`mimic GI conditions closely enough (45). The lack of
`predictivity is not necessarily a limitation of the test, but
`may result from inappropriate selection of acceptance criteria
`or specific analytical conditions. In some applications, an
`overly sensitive dissolution test is desirable. For example, the
`FDA guidance on dissolution testing of immediate-release
`(IR) solid oral drugs (40) includes a procedure for manufac-
`turing bioequivalent product
`lots with different in vitro
`dissolution to identify and establish an acceptable range for
`critical manufacturing variables. As for the question of
`biorelevance, because the goal of the dissolution procedure
`is to establish equivalence with acceptable clinical lots, the
`procedure need only be predictive of bioavailability. For this
`purpose, mimicking the gastrointestinal tract is not relevant.
`Ensuring that the procedure is predictive should be addressed
`during a rational method development
`following QbD
`principles.
`
`Tests for Similarity and Difference
`
`Because the comparison of dissolution profiles is used to
`evaluate the effects of formulation changes, the stability of
`product performance over time, and lot-to-lot manufacturing
`consistency and to demonstrate bioequivalence, it is important
`to understand the strengths and weaknesses of the various
`methods used for comparing them. Coming up with an
`objective data-based means of deciding if dissolution profiles
`are similar or different is a challenge. Because a dissolution
`profile is a plot of cumulative percent drug released (i.e., each
`data point is dependent on the previous data point) versus
`time, the underlying assumption of data independence is
`violated, precluding the use of statistical tests of difference
`(50). The use of exploratory data analysis methods, such as
`overlapping confidence intervals at individual time points as a
`test of similarity, becomes problematic when they overlap at
`some, but not all, of the time points.
`
`

`

`USP 1 and 2 Dissolution: Present Challenges and Future Relevance
`
`1293
`
`Mathematical comparison methods such as f1 and f2
`(51,52) utilize differences between average reference and test
`profiles at each sampling interval to provide a single number
`with which to quantify the similarity or difference between
`them. The mathematical comparator f2 (similarity factor)
`(51), which is recommended by the FDA (40,46–48), has the
`advantage of being easy to calculate. However, this technique
`is sensitive to the number of dissolution time points after the
`plateau is reached (52) and does not account for vessel-to-
`vessel variability, and there does not appear to be a well-
`defined basis for the “sameness” threshold of f2=50 (50). More
`importantly, because f2 is a sample statistic and is not based on
`a known population, the probability of type I (rejecting similar
`profiles as dissimilar) and type II (accepting dissimilar profiles
`as being similar) error is unknown (50,53,54). A modification
`to this method, in which f2 is calculated for each individual
`dosage unit, has been used to allow the inter-vessel variability
`to be expressed (54). The use of bootstrapping to simulate
`confidence intervals has been used with highly variable data to
`avoid making false conclusions (52,55).
`Model-dependent methods involve fitting the reference
`dissolution profile data to a mathematical function (known
`physical curve); similarity of a test profile is evaluated in
`terms of difference between the mean model parameters of
`the reference and test curves (56). Models that have been
`used include zero-order (54,56), first-order (54,57), Hixson–
`Crowell (54,57), Higuchi (54,57), quadratic (57), Weibull
`(54,56–58), Gompertz (57,58), Probit (58), exponential (58),
`and logistic (57,58). These methods have the advantage of
`taking into account variance and covariance of the data sets,
`and sampling time points for the reference and test profiles
`do not have to be the same. However, it is not always possible
`to find a model that adequately fits the data. Selection of an
`inappropriate model curve can yield misleading results,
`resulting in incorrect conclusions, so it is important to run a
`lack-of-fit test on the reference data prior to comparing
`model parameters (59).
`Statistical multivariate methods using multivariate
`ANOVA have also been used (60,61). These do take into
`account variability and correlation structure of cumulative
`percent-released-versus-time data. An advantage is that they
`can be used to make estimates of type I and type II errors.
`
`In Vitro and In Vivo Relationships and Bioequivalence
`Challenges in Dissolution Method Development
`
`IVIVCs were introduced as the desire of both industry
`and regulatory agency to reduce development time, cost, and
`regulatory burden (62). Recognizing that dissolution rate,
`aqueous solubility, and gastrointestinal permeability are the
`key parameters that control the rate and extent of drug
`absorption, Amidon et al. (63) proposed a Biopharmaceutics
`Classification Scheme (BCS) in 1995. Later, FDA classified
`drug substances into four groups: class I—high solubility, high
`permeability; class II—low solubility, high permeability; class
`III—high solubility, low permeability; class IV—low solubil-
`ity, low permeability (48,64). For rapidly dissolving class I
`drugs, because of their high solubility and high permeability
`characteristics, the in vivo dissolution is not the rate-limiting
`step, so IVIVC may not be possible (63–65). In addition,
`since gastric emptying is the key factor in determining the
`
`plasma profile, if the excipients in the drug product alter the
`gastric-emptying rate, bioinequivalent products will be the
`result. For class II drugs, on other hand, dissolution may be
`the limiting step of the drug absorption, therefore, an IVIVC
`may be expected (21,62). More research is needed to develop
`and validate in vitro dissolution methods for class II drugs so
`that they can be used to predict in vivo dissolution (66). For
`class III drugs, permeability is the limiting step of
`the
`absorption, and a limited IVIVC may be expected, and
`finally, for class IV drugs, IVIVC is difficult. The drug will
`have both limited dissolution and permeability so it will be
`difficult, at best, to develop a dissolution model unless the
`permeability is borderline low.
`Currently, there are four levels of IVIVC defined in
`FDA guidances (62,67–72). Level A correlation is a point-to-
`point relationship between in vitro dissolution and the in vivo
`pharmacokinetic data (73). It
`is generally linear and is
`reviewed as a predictive and preferred approach (62,73,74).
`In the case of a level A correlation, in vitro dissolution data
`can serve as a surrogate for in vivo performance. For a class I
`drug, IVIVC is generally not likely, but when formulated as
`an extended-release product and the solubility and perme-
`ability of the drug is site-independent, a level A correlation is
`expected (75,76). For a class II drug formulated as an
`extended-release product, and the solubility and permeability
`of the drug are site-independent, a level A correlation is also
`likely; however, if the permeability is site-dependent, IVIVC
`is unlikely (75). Level B correlation applies the principles of
`statistical moment analysis. It compares the mean in vitro
`dissolution time to either the mean residence time or the
`mean in vivo dissolution time (77,78), so it does not reflect
`the actual in vivo plasma concentration curve. Therefore, level
`B correlation alone cannot support biowaivers. A level C
`correlation represents a single-point relationship between a
`dissolution parameter (e.g., t50%, t90%) and a pharmacokinetic
`parameter (e.g., AUC, Tmax, Cmax). This correlation does not
`reflect the entire plasma-concentration–time curve or dissolu-
`tion profile (62,79); therefore,
`it is considered the lowest
`correlation level. However, Level C correlation can provide
`useful information in early formulation development. For a class
`I drug, if the permeability is site-dependent, a level C correlation
`is expected (76). A multiple level C correlation compares one or
`more pharmacokinetic parameters of interest (e.g., Cmax, AUC)
`to the amount of drug dissolved at several time points of the
`dissolution profile. This level of correlation may support a
`biowaiver if the correlation has been established over the entire
`dissolution profile with one or more pharmacokinetic parame-
`ters of interest. If a multiple level C correlation is possible, then
`it is likely that a level A correlation is possible as well, and the
`latter is the preferred correlation. In addition, level D correla-
`tion, which is a qualitative a rank-order correlation, has been
`described in an FDA guidance (62). This correlation can be
`useful
`in drug development but cannot support regulatory
`application.
`Compared with immediate-release (IR) products (80,81),
`more attention has been given to the application of IVIVC
`for controlled-release oral dosage formulations (82–85),
`where formulation technology controls the release rate, thus
`drug release is the rate-limiting factor in the absorption
`process. For BCS class II drugs with immediate-release (IR)
`formulations, because of the intricacy of gastric emptying as
`
`

`

`1294
`
`Gray et al.
`
`well as the low resolution of plasma data at early time points
`(0–3 h), a meaningful level A correlation seems unlikely and
`few publications have been made so far. However, Lue et al.
`(86) and Buch et al. (87) recently applied biorelevant
`dissolution media (BDM) in the investigation of the IVIVC of
`class II, immediate-release (IR) compounds. The application of
`IVIVC to non-oral products, such as parenteral depots or
`injectable dosage forms, has also been investigated (72,88–93).
`It is worthwhile to mention that for IR drugs, in vitro–in vivo
`re

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket