throbber
+
`
`+
`
`October 1996
`Volume 85, Number 10
`
`REVIEW ARTICLE
`
`Pharmaceutical Applications of Cyclodextrins. 1. Drug Solubilization and
`Stabilization
`
`THORSTEINN LOFTSSON*X AND MARCUS E. BREWSTER†
`Received December 29, 1995, from the *Department of Pharmacy, University of Iceland, P.O. Box 7210, IS-127 Reykjavik, Iceland, and
`†Pharmos Corporation, Two Innovation Drive, Alachua, FL 32615 .
`Final revised manuscript received March 1, 1996 .
`Accepted
`for publication March 19, 1996X.
`
`Abstract 0 Cyclodextrins are cyclic oligosaccharides which have recently
`been recognized as useful pharmaceutical excipients. The molecular
`structure of these glucose derivatives, which approximates a truncated
`cone or torus, generates a hydrophilic exterior surface and a nonpolar
`cavity interior. As such, cyclodextrins can interact with appropriately sized
`molecules to result
`in the formation of
`inclusion complexes. These
`noncovalent complexes offer a variety of physicochemical advantages
`over the unmanipulated drugs including the possibility for increased water
`solubility and solution stability. Further, chemical modification to the parent
`cyclodextrin can result in an increase in the extent of drug complexation
`and interaction.
`In this short review, the effects of substitution on various
`cyclodextrin properties and the forces involved in the drug- cyclodextrin
`complex formation are discussed. Some general observations are made
`predicting drug solubilization by cyclodextrins.
`In addition, methods which
`are useful in the optimization of complexation efficacy are reviewed. Finally,
`the stabilizing/destabilizing effects of cyclodextrins on chemically labile
`drugs are evaluated.
`
`Introduction
`Although cyclodextrins are frequently regarded as a new
`group of pharmaceutical excipients, they have been known
`for over 100 years.1 The foundations of cyclodextrin chemistry
`were laid down in the first part of this century2,3 and the first
`patent on cyclodextrins and their complexes was registered
`in 1953.4 However, until 1970 only small amounts of cyclo-
`dextrins could be produced and high production costs pre-
`vented their widespread usage in pharmaceutical formula-
`tions. Recent biotechnological advancements have resulted
`in dramatic improvements in cyclodextrin production, which
`has lowered their production costs. This has led to the
`availability of highly purified cyclodextrins and cyclodextrin
`derivatives which are well suited as pharmaceutical excipi-
`
`X Abstract published in Advance ACS Abstracts, May 1, 1996.
`
`ents. These carbohydrates are mainly used to increase the
`aqueous solubility, stability, and bioavailability of drugs, but
`they can also, for example, be used to convert liquid drugs
`into microcrystalline powders, prevent drug-drug or drug-
`additive interactions, reduce gastrointestinal or ocular irrita-
`tion, and reduce or eliminate unpleasant taste and smell.
`The following is a short review of the effects of cyclodextrins
`on the solubility and stability of drugs in aqueous solutions
`with emphasis on the more recent developments. For further
`information on cyclodextrins and their physicochemical prop-
`erties the reader is referred to several excellent books and
`reviews published in recent years.5-13
`
`Structure and Physicochemical Properties
`Cyclodextrins are cyclic (R-1,4)-linked oligosaccharides of
`R-D-glucopyranose containing a relatively hydrophobic central
`cavity and hydrophilic outer surface. Owing to lack of free
`rotation about the bonds connecting the glucopyranose units,
`the cyclodextrins are not perfectly cylindrical molecules but
`are toroidal or cone shaped. Based on this architecture, the
`primary hydroxyl groups are located on the narrow side of
`the torus while the secondary hydroxyl groups are located on
`the wider edge (Figure 1). The most common cyclodextrins
`are R-cyclodextrin, (cid:226)-cyclodextrin, and (cid:231)-cyclodextrin, which
`consist of six, seven, and eight glucopyranose units, respec-
`tively. While it is thought that, due to steric factors, cyclo-
`dextrins having fewer than six glucopyranose units cannot
`exist, cyclodextrins containing nine, ten, eleven, twelve, and
`thirteen glucopyranose units, which are designated (cid:228)-, (cid:15)-, œ-,
`Ł-, and ı-cyclodextrin, respectively, have been reported.14,15
`Of these large-ring cyclodextrins only (cid:228)-cyclodextrin has been
`well characterized.16,17 Chemical and physical properties of
`the four most common cyclodextrins are given in Table 1. The
`melting points of R-, (cid:226)-, and (cid:231)-cyclodextrin are between 240
`and 265 °C, consistent with their stable crystal
`lattice
`structure.18
`The parent cyclodextrins, in particular (cid:226)-cyclodextrin, have
`limited aqueous solubility, and their complex formation with
`
`© 1996, American Chemical Society and
`American Pharmaceutical Association
`
`S0022-3549(95)00534-X CCC: $12.00
`
`Journal of Pharmaceutical Sciences / 1017
`Vol. 85, No. 10, October 1996
`
`APOTEX EX1056
`
`Page 1
`
`

`
`+
`
`+
`
`a
`
`b
`
`Figure 1s(a) The chemical structure and (b) the toroidal shape of the (cid:226)-cyclodextrin molecule.
`Table 1sSome Characteristics of r-, (cid:226)-, (cid:231)-, and (cid:228)-Cyclodextrina
`R
`(cid:226)
`(cid:231)
`
`(cid:228)
`
`6
`972
`4.7- 5.3
`14.5
`
`7
`1135
`6.0- 6.5
`1.85
`
`8
`1297
`7.5- 8.3
`23.2
`
`9
`1459
`10.3- 11.2
`8.19
`
`No. of glucopyranose units
`Molecular weight
`Central cavity diameter (Å)
`Water solubility at 25 (cid:176) C (g/100 mL)
`a Modified from refs 5 and 17.
`lipophilic drugs, and other compounds with limited aqueous
`solubility, frequently gives rise to B-type phase-solubility
`diagrams as defined by Higuchi.19 That is, addition of these
`unmodified cyclodextrins to aqueous drug solutions or drug
`suspensions often results in precipitation of solid drug-
`cyclodextrin complexes. The aqueous solubility of the parent
`cyclodextrins is much lower than that of comparable acyclic
`saccharides, and this could partly be due to relatively strong
`binding of the cyclodextrin molecules in the crystal state (i.e.,
`relatively high crystal lattice energy). In addition, (cid:226)- and
`(cid:228)-cyclodextrin form intramolecular hydrogen bonds between
`secondary OH groups, which detracts from hydrogen bond
`formation with surrounding water molecules, resulting in less
`negative heats of hydration.5,17 Thus, intramolecular hydro-
`gen bonding can result in relatively unfavorable enthalpies
`of solution and low aqueous solubilities. Substitution of any
`of the hydrogen bond forming hydroxyl groups, even by
`hydrophobic moieties such as methoxy and ethoxy functions,
`will result in a dramatic increase in water solubility.5 For
`example, the aqueous solubility of (cid:226)-cyclodextrin is only 1.85%
`(w/v) at room temperature but increases with increasing
`degree of methylation. The highest solubility is obtained
`when two-thirds of the hydroxyl groups (i.e., 14 of 21) are
`methylated, but then falls upon more complete alkylation.
`That is, the permethylated derivative has a solubility that is
`lower than that of, e.g., heptakis(2,6-O-dimethyl)-(cid:226)-cyclodex-
`trin but that is still considerably higher than that of unsub-
`stituted (cid:226)-cyclodextrin.7 Other common cyclodextrin deriva-
`tives are formed by other types of alkylation or hydroxy-
`alkylation of the hydroxyl groups.5,20 The main reason for the
`solubility enhancement in these derivatives is that chemical
`manipulation frequently transforms the crystalline cyclodex-
`trins into amorphous mixtures of isomeric derivatives. For
`example, (2-hydroxypropyl)-(cid:226)-cyclodextrin is obtained by treat-
`ing a base-solubilized solution of (cid:226)-cyclodextrin with propylene
`oxide, resulting in an isomeric system that has an aqueous
`solubility well in excess of 60% (w/v).21 The number of isomers
`generated based on random substitution is very large. Sta-
`tistically, for example, there are about 130 000 possible
`heptakis(2-O-(hydroxypropyl))-(cid:226)-cyclodextrin derivatives, and
`given that introduction of the 2-hydroxypropyl function also
`
`1018 / Journal of Pharmaceutical Sciences
`Vol. 85, No. 10, October 1996
`
`introduced an optically active center, the number of total
`isomers, i.e., geometrical and optical, is even much greater.
`In reality, the chemical alkylation of cyclodextrins is not
`totally random, based on relative reactivities of the hydroxy
`functions in the molecule. The secondary OH groups on the
`cyclodextrin molecule (i.e., OH-2 and OH-3 on the glucopy-
`ranose units) are somewhat more acidic than the primary OH
`group (i.e., OH-6). Thus, alkylation of OH-6, the least
`sterically crowded functionality, is favored in strong basic
`solutions while alkylation of OH-2, the most acidic of the
`hydroxyl groups but also the most hindered, is favored in a
`weak basic solution.22 Thus, some degree of regioselectivity
`is possible. Both the molar substitution, i.e., the average
`number of alkyl or hydroxyalkyl groups that have been reacted
`with one glucopyranose unit, and the location of the alkyl or
`hydroxyalkyl groups on the cyclodextrin molecule will affect
`the physicochemical properties of the derivatives including
`their ability to form drug complexes.23 However, theoretical
`studies have shown that the alkylation and hydroxyalkylation
`of the cyclodextrins should not introduce significant steric
`hindrance.24 Some of the commercially available cyclodextrins
`are listed in Table 2.
`
`Cyclodextrin Complexes
`The central cavity of the cyclodextrin molecule is lined with
`skeletal carbons and ethereal oxygens of the glucose residues.
`It is therefore lipophilic. The polarity of the cavity has been
`estimated to be similar to that of aqueous ethanolic solution.5
`It provides a lipophilic microenvironment into which suitably
`sized drug molecules may enter and be included. No covalent
`bonds are formed or broken during drug-cyclodextrin complex
`formation, and in aqueous solutions, the complexes are readily
`dissociated. Free drug molecules are in equilibrium with the
`molecules bound within the cyclodextrin cavity. Measure-
`ments of stability or equilibrium constants (Kc) or the dis-
`sociation constants (Kd) of the drug-cyclodextrin complexes
`are important since this is an index of changes in physico-
`chemical properties of a compound upon inclusion. Most
`methods for determining the K values are based on titrating
`changes in the physicochemical properties of the guest
`molecule, i.e., the drug molecule, with the cyclodextrin and
`then analyzing the concentration dependencies. Additive
`properties that can be titrated in this way to provide informa-
`tion on the K values include25 aqueous solubility,19,26-28
`chemical reactivity,10,29,30 molar absorptivity and other optical
`properties (CD, ORD),31-34 phase solubility measurements,35
`NMR chemical shifts,23,36 pH-metric methods,37 calorimetric
`titration,38 freezing point depression,39 and LC chromato-
`
`Page 2
`
`

`
`+
`
`+
`
`Table 2sSome Currently Available Cyclodextrins Obtained by
`Substitution of the OH Groups Located on the Edge of the Cyclodextrin
`Ringa
`
`R
`
`Methyl
`
`Butyl
`
`2-Hydroxypropyl
`
`Acetyl
`
`Succinyl
`
`Glucosyl
`Maltosyl
`
`Carboxymethyl ether
`
`Phosphate ester
`
`Simple polymers
`Carboxymethyl
`
`Cyclodextrin Derivatives
`
`(cid:226)
`
`Alkylated:
`
`Methyl
`Ethyl
`Butyl
`
`Hydroxylalkylated:
`Hydroxyethyl
`2-Hydroxypropyl
`2-Hydroxybutyl
`Esterified:
`
`Acetyl
`Propionyl
`Butyryl
`Succinyl
`Benzoyl
`Palmityl
`Toluenesulfonyl
`Esterified and Alkylated:
`Acetyl methyl
`Acetyl butyl
`Branched:
`Glucosyl
`Maltosyl
`
`Ionic:
`Carboxymethyl ether
`Carboxymethyl ethyl
`Phosphate ester
`3-Trimethylammonium-2-
`hydroxypropyl ether
`Sulfobutyl ether
`Polymerized:
`Simple polymers
`Carboxymethyl
`
`(cid:231)
`
`Methyl
`
`Butyl
`Pentyl
`
`Hydroxyethyl
`2-Hydroxypropyl
`
`Acetyl
`
`Succinyl
`
`Glucosyl
`Maltosyl
`
`Carboxymethyl ether
`
`Phosphate ester
`
`Simple polymers
`Carboxymethyl
`
`the
`a Since both the number of substitutes and their location will affect
`physicochemical properties of the cyclodextrin molecules, such as their aqueous
`solubility and complexing abilities, each derivative listed should be regarded as a
`group of closely related cyclodextrin derivatives.
`
`graphic retention times.40 While it is possible to use both
`guest or host changes to generate equilibrium constants, guest
`properties are usually most easily assessed. Connors has
`evaluated the population characteristics of cyclodextrin com-
`plex stabilities in aqueous solution.41
`The thermodynamic parameters, i.e., the standard free
`energy change (¢G), the standard enthalpy change (¢H), and
`the standard entropy change (¢S), can be obtained from the
`temperature dependence of the stability constant of the
`cyclodextrin complex.42 The thermodynamic parameters for
`several series of drugs and other compounds have been
`determined and analyzed.43-45 The thermodynamic param-
`eters of several other drugs are listed in Table 3. The complex
`formation is almost always associated with a relatively large
`negative ¢H and a ¢S that can be either positive or negative.
`Also, complex formation is largely independent of the chemical
`properties of the guest (i.e., drug) molecules. The association
`of binding constants with substrate polarizability suggests
`that van der Waals forces are important in complex forma-
`tion.50 Hydrophobic interactions are associated with a slightly
`positive ¢H and a large positive ¢S; therefore, classical
`hydrophobic interactions are entropy driven, suggesting that
`they are not involved with cyclodextrin complexation since,
`as indicated, these are enthalpically driven processes. Fur-
`thermore, for a series of guests there tends to be a linear
`relationship between enthalpy and entropy, with increasing
`
`7
`7
`
`7
`7
`2
`4
`5
`6
`2
`3
`4
`6
`5
`8
`9
`1
`
`(cid:226)-CD
`
`Diazepam (pKa 3.3)
`
`(cid:226)-CD
`
`Hydrochlorothiazide
`(pKa 8.8 and 10.4)
`
`Hydrocortisone
`Phenytoin, un-ionized
`Phenytoin, ionized
`Naproxen
`Adenine arabinoside
`Adenosine
`Ibuprofen (pKa 5.2)
`
`(cid:226)-CD
`(cid:226)-CD
`(cid:226)-CD
`(cid:226)-CD
`
`Table 3sStandard Enthalpy Change (¢H) and Standard Entropy Change
`(¢S) for Several Drug- Cyclodextrin Complexes
`pH ¢H(kJ/mol) ¢S(J/(mol K)) Ref
`Cyclodextrina
`Drug
`- 32
`- 70
`HP-R-CD
`49
`- 38
`- 67
`(cid:226)-CD
`46
`- 21
`- 21
`46
`- 13
`31
`18
`- 28
`- 64
`32
`- 21
`- 53
`32
`- 29
`47
`15
`- 32
`47
`4
`- 29
`47
`3
`- 17
`47
`34
`- 0.2
`47
`70
`- 3.3
`47
`69
`- 17
`47
`22
`- 18
`47
`19
`- 40
`47
`62
`- 39
`47
`59
`- 42
`47
`70
`- 68
`- 166
`HP-(cid:226)-CD
`48
`Acetylsalicylic acid
`- 18
`- 26
`HP-(cid:226)-CD
`49
`Acetazolamide
`- 71
`- 151
`17(cid:226)-Estradiol
`HP-(cid:226)-CD
`49
`- 20
`- 6
`HP-(cid:226)-CD
`49
`Hydrocortisone
`- 55
`- 127
`HP-(cid:226)-CD
`48
`1
`Methyl acetylsalicylate
`- 63
`- 144
`HP-(cid:226)-CD
`48
`1
`Methyl salicylate
`- 57
`- 134
`M/DM-(cid:226)-CD
`48
`1
`Acetylsalicylic acid
`- 20
`- 28
`M/DM-(cid:226)-CD Methyl acetylsalicylate
`48
`1
`- 28
`- 56
`HP-(cid:231)-CD
`48
`1
`Acetylsalicylic acid
`- 75
`- 194
`HP-(cid:231)-CD
`48
`1
`Methyl acetylsalicylate
`- 73
`- 176
`HP-(cid:231)-CD
`48
`1
`Methyl salicylate
`a HP-R-CD:
`(2-hydroxypropyl)-R-cyclodextrin . (cid:226)-CD: (cid:226)-cyclodextrin. HP-(cid:226)-
`(2-hydroxypropyl)-(cid:226)-cyclodextrin. M/DM-(cid:226)-CD: mixture of maltosyl- and
`CD:
`dimaltosyl-(cid:226)-cyclodextrin (3:7). HP-(cid:231)-CD:
`(2-hydroxypropyl)-(cid:231)-cyclodextrin.
`enthalpy related to less negative entropy values.43-45,48 This
`effect, termed compensation, is often correlated with water
`acting as a driving force in complex formation. The main
`driving force for complex formation could, therefore, be the
`release of enthalpy-rich water from the cyclodextrin cavity.47
`The water molecules located inside the cavity cannot satisfy
`their hydrogen-bonding potentials; therefore, they are of
`higher enthalpy.51 The energy of the system is lowered when
`these enthalpy-rich water molecules are replaced by suitable
`guest molecules which are less polar than water. Other
`mechanisms that are thought to be involved with complex
`formation have been identified in the case of R-cyclodextrin.
`In this instance, release of ring strain is thought to be involved
`with the driving force for compound-cyclodextrin interaction.
`Hydrated R-cyclodextrin is associated with an internal hy-
`drogen bond to an included water molecule which perturbs
`the cyclic structure of the macrocycle. Elimination of the
`included water and the associated hydrogen bond is related
`to a significant release of steric strain decreasing the system
`enthalpy.52
`In addition, “nonclassical hydrophobic effects”
`have been invoked to explain complexation. These nonclas-
`sical hydrophobic effects are a composite force in which the
`classic hydrophobic effects (characterized by large positive ¢S)
`and van der Waals effects (characterized by negative ¢H and
`negative ¢S) are operating in the same system. Using
`adamantanecarboxylates as probes, R-, (cid:226)-, and (cid:231)-cyclodextrins
`were examined.53 In the case of R-cyclodextrin, experimental
`data indicated small changes in ¢H and ¢S consistent with
`little interaction between the bulky probe and the small cavity.
`In the case of (cid:226)-cyclodextrin, a deep and snug-fitting complex
`was formed leading to a large negative ¢H and a near zero
`¢S. Finally, complexation with (cid:231)-cyclodextrin demonstrated
`near zero ¢H values and large positive ¢S values consistent
`with a classical hydrophobic interaction. Evidently, the cavity
`size of (cid:231)-cyclodextrin was too large to provide for a significant
`
`Journal of Pharmaceutical Sciences / 1019
`Vol. 85, No. 10, October 1996
`
`Page 3
`
`

`
`+
`
`+
`
`Table 4sEffect of Poly(vinylpyrrolidone) Concentration on the Value of
`the Apparent Stability Constant (Kc) of Some
`Drug-
`(2-Hydroxypropyl)-(cid:226)-cyclodextrin (1:1) Complexes at Room
`Temperature (20- 23 (cid:176) C)a
`
`PVP (% w/v)
`
`Acetazolamide
`
`0.00
`0.10
`0.25
`0.50
`
`86.2
`95.4
`97.0
`96.2
`
`Kc (M-1)
`Hydrocortisone
`
`1010
`1450
`c
`1190
`
`17(cid:226)-Estradiol
`
`52900
`58800
`78200
`80400
`
`a From ref 49. b Poly(vinylpyrrolidone). c Not determined.
`
`contribution by van der Waals-type interactions. These
`various explanations show that there is no simple construct
`to describe the driving force for complexation. Although
`release of enthalpy-rich water molecules from the cyclodextrin
`cavity is probably an important driving force for drug-
`cyclodextrin complex formation, other forces may be impor-
`tant. These forces include van der Waals interactions,34,54
`hydrogen bonding,55,56 hydrophobic interactions,34,57 release
`of ring strain in the cyclodextrin molecule,56 and changes in
`solvent-surface tensions.58
`Methods of preparing drug-cyclodextrin complexes have
`been reviewed.25
`In the solution phase, the procedure is
`generally as follows: an excess amount of the drug is added
`to an aqueous cyclodextrin solution, and the suspension is
`agitated for up to 1 week at the desired temperature. The
`suspension is then filtered or centrifuged to form a clear drug-
`cyclodextrin complex solution. For preparation of solid for-
`mulations of the drug-cyclodextrin complex, the water is
`removed from the aqueous drug-cyclodextrin complex solu-
`tion by evaporation or sublimation. It is sometimes possible
`to shorten this process by formation of supersaturated solu-
`tions through sonication followed by precipitation at the
`desired temperature. In some cases, the efficiency of com-
`plexation is not very high, and therefore, relatively large
`amounts of cyclodextrins must be used to complex small
`amounts of drug. To add to this difficulty, vehicle additives,
`osmolality modifiers, and pH adjustments commonly used in
`drug formulations, such as sodium chloride, buffer salts,
`surfactants, preservatives, and organic solvents, very often
`reduce the efficiency. For example, in aqueous solutions,
`ethanol and propylene glycol at low concentrations have been
`shown to reduce the cyclodextrin complexation of testosterone
`and ibuprofen by acting as competing guest molecules while
`at higher concentrations they can reduce complexation through
`a manipulation of solvent dielectric constant.48,59 Likewise,
`non-ionic surfactants have been shown to reduce cyclodextrin
`complexation of diazepam60 and preservatives to reduce the
`cyclodextrin complexation of various steroids.61 On the other
`hand, additives such as ethanol can promote complex forma-
`tion in the solid or semisolid state.62 Un-ionized drugs usually
`form a more stable cyclodextrin complex than their ionic
`counterparts; thus, the complexation efficiency of basic drugs
`can be enhanced by addition of ammonia to the aqueous
`complexation media.
`For example, solubilization of pancratistatin with (hydroxy-
`propyl)-cyclodextrins was optimized upon addition of am-
`monium hydroxide.63 Freeze-drying of the solutions removed
`ammonia, resulting in ammonia-free solid complex prepara-
`tions which dissolved rapidly to form clear supersaturated
`pancratistatin solutions. The resulting solutions were stable
`for a few hours, time sufficient for potential use in parenteral
`preparations. Finally, enhanced complexation can be obtained
`by formation of ternary complexes (or cocomplexes) between
`a drug molecule, a cyclodextrin molecule, and a third compo-
`nent. For instance, addition of a small amount of various
`
`1020 / Journal of Pharmaceutical Sciences
`Vol. 85, No. 10, October 1996
`
`water-soluble polymers to an aqueous complexation medium,
`followed by heating of the medium in an autoclave, can
`significantly increase the apparent stability constant of the
`drug-cyclodextrin complex (Table 4).49,64,65 A somewhat
`similar effect has been obtained through formation of drug-
`hydroxy acid-cyclodextrin ternary complexes or salts with
`basic drugs.66-68
`
`Drug Solubilization
`
`The most common pharmaceutical application of cyclodex-
`trins is to enhance drug solubility in aqueous solutions. Some
`of the reports generated on this topic have been reviewed,5-9
`and additional data is available from the individual cyclodex-
`trin manufacturers. The solubilizing effects of various cyclo-
`dextrins on three different drugs are listed in Table 5.
`Although prediction of compound solubilization by cyclodex-
`trins continues to be highly empirical, various historical
`observations permit several general statements. First, the
`lower the aqueous solubility of the pure drug, the greater the
`relative solubility enhancement obtained through cyclodextrin
`complexation. Drugs that possess aqueous solubility in the
`micromole/liter range generally demonstrate much greater
`enhancement than drugs possessing solubility in the micro-
`mole/liter range or higher. In Table 5, the enhancement factor,
`i.e., the solubility in the aqueous cyclodextrin solution divided
`by the solubility in pure water, for paclitaxel, for example, is
`much larger than the enhancement factors for hydrocortisone
`and pancratistatin. A similar observation was made when
`the solubilizing effect of (2-hydroxypropyl)-(cid:226)-cyclodextrin on
`53 different drugs was investigated.9 Second, cyclodextrin
`derivatives of lower molar substitution are better solubilizers
`than the same type of derivatives of higher molar substitution.
`In Table 5, both randomly methylated (cid:226)- and (cid:231)-cyclodextrins
`with molar substitution 0.6 provide for better solubilization
`than the same type of randomly methylated cyclodextrins with
`molar substitution 1.8. With the exception of R-cyclodextrin,
`permethylated derivatives (of (cid:226)- and (cid:231)-cyclodextrin) possess
`a lower complexing potential (lower Kc value) than the parent
`cyclodextrins.23 Of the commercially available materials, the
`methylated cyclodextrins with relatively low molar substitu-
`tion appear to be the most powerful solubilizers. The chain
`length of the alkyl group, on the other hand, appears to be of
`less importance.24,70 Third, charged cyclodextrins can be
`powerful solubilizers, but their solubilizing effect appears to
`depend on the relative proximity of the charge to the cyclo-
`dextrin cavity. The farther away the charge is located, the
`better the complexing abilities. For example, (2-hydroxy-3-
`(trimethylammonio)propyl)-(cid:226)- and -(cid:231)-cyclodextrin possess ex-
`cellent solubilizing effects while (cid:226)-cyclodextrin sulfate has a
`relatively low complexation potential (Table 5). Sulfobutyl
`ether (cid:226)-cyclodextrin, where the anion has been moved away
`from the cavity by a butyl ether spacer group, is an excellent
`solubilizer.71 (Carboxymethyl)-(cid:226)-cyclodextrin is another in-
`teresting anionic cyclodextrin derivative.72 Compared to
`neutral cyclodextrins, enhanced complexation is frequently
`observed when the drug and cyclodextrin molecules have
`opposite charge but decreased complexation is observed if they
`carry same type of charge. For example,
`(2-hydroxy-3-
`(trimethylammonio)propyl)-(cid:226)-cyclodextrin is an excellent solu-
`bilizer for many acidic drugs capable of forming anions.
`Another finding is that while many ionizable drugs are able
`to form cyclodextrin complexes, the stability constant of the
`complex is much larger for the un-ionized than for the ionized
`form. For example, both the un-ionized and the cationic (i.e.,
`the protonated) form of chlorpromazine give rise to 1:1
`complexes with (cid:226)-cyclodextrin but the stability constant for
`the un-ionized form is 4 times larger than for the cationic
`
`Page 4
`
`

`
`+
`
`+
`
`Table 5sSolubility of Drugs in Different Cyclodextrin Solutions at Room Temperature
`
`Drug
`
`Cyclodextrina
`
`Concnb (% w/v)
`
`Solubility (mM)
`
`Enhancementc Factor
`
`Ref
`
`Hydrocortisone (MW 362)
`
`Paclitaxel (Taxol, MW 854)d
`
`Pancratistatin (MW 325)
`
`None
`Glucosyl-R-CD
`Maltosyl-R-CD
`HP-(cid:226)-CD MS 0.6
`HE-(cid:226)-CD
`RM-(cid:226)-CD MS 0.6
`RM-(cid:226)-CD MS 1.8
`HTMAP-(cid:226)-CD MS 0.5
`CM-(cid:226)-CD MS 0.6
`Glucosyl-(cid:226)-CD
`Maltosyl-(cid:226)-CD
`RM-(cid:231)-CD MS 0.6
`RM-(cid:231)-CD MS 1.8
`None
`(cid:226)-CD
`Dimaltosyl-(cid:226)-CD
`HE-(cid:226)-CD
`HP-(cid:226)-CD
`DM-(cid:226)-CD
`(cid:231)-CD
`HP-(cid:231)-CD
`None
`HTMAP-(cid:226)-CD MS 1.4
`S-(cid:226)-CD Na-salt MS 2.3
`CM-(cid:226)-CD Na-salt MS 0.6
`HP-(cid:226)-CD MS 0.5
`Maltosyl-(cid:226)-CD MS 0.14
`DM-(cid:226)-CD MS 2.0
`HE-(cid:226)-CD
`(cid:231)-CD
`HTMAP-(cid:231)-CD MS 0.3
`HP-(cid:231)-CD MS 0.7
`TM-(cid:231)-CD MS 3.0
`
`10
`10
`10
`10
`10
`10
`10
`10
`10
`10
`10
`10
`
`1.5
`50
`50
`50
`50
`15
`50
`
`10
`10
`10
`10
`10
`10
`10
`10
`10
`10
`10
`
`0.993
`7.45
`11.3
`33.7
`48.3
`72.2
`50.8
`30.3
`44.6
`46.9
`28.7
`58.8
`38.6
`4 · 10-4
`0.005
`0.115
`0.914
`0.856
`39.6
`0.020
`0.080
`0.16
`0.86
`0.28
`0.83
`1.0
`0.95
`1.2
`0.83
`0.80
`0.49
`0.83
`0.49
`
`7.50
`11.4
`33.9
`48.6
`72.7
`51.2
`30.1
`44.9
`47.2
`28.9
`55.2
`38.9
`
`13
`288
`2285
`2140
`99.000
`50
`200
`
`5.4
`1.8
`5.2
`6.3
`5.9
`7.5
`5.2
`5.0
`3.1
`5.2
`3.1
`
`49
`49
`49
`49
`49
`27
`27
`27
`27
`49
`49
`27
`27
`69
`69
`69
`69
`69
`69
`69
`69
`63
`63
`63
`63
`63
`63
`63
`63
`63
`63
`63
`63
`
`randomly methylated (cid:226)-cyclodextrin.
`(hydroxyethyl)-(cid:226)-cyclodextrin. RM-(cid:226)-CD:
`(2-hydroxypropyl)-(cid:226)-cyclodextrin. HE-(cid:226)-CD:
`a(cid:226)-CD: (cid:226)-cyclodextrin. HP-(cid:226)-CD:
`HTMAP-(cid:226)-CD: (2-hydroxy-3-(trimethylammonio)propyl)-(cid:226)-cyclodextrin. CM-(cid:226)-CD: (carboxymethyl)-(cid:226)-cyclodextrin. Glucosyl-(cid:226)-CD: glucosyl-(cid:226)-cyclodextrin. Maltosyl-
`(cid:226)-CD: maltosyl-(cid:226)-cyclodextrin. DM-(cid:226)-CD: 2,6-O-dimethyl-(cid:226)-cyclodextrin. S-(cid:226)-CD: (cid:226)-cyclodextrin sulfate. (cid:231)-CD: (cid:231)-cyclodextrin. RM-(cid:231)-CD:
`randomly methylated
`(cid:231)-cyclodextrin. HP-(cid:231)-CD: (2-hydroxypropyl)-(cid:231)-cyclodextrin. HTMAP-(cid:231)-CD: (2-hydroxy-3-(trimethylammonio)propyl)-(cid:231)-cyclodextrin. TM-(cid:231)-CD:
`trimethyl (cid:231)-cyclodextrin.
`MS: molar substitution (i.e., the average number of OH groups on each glucose repeat unit that have been substituted). Na salt: sodium salt. b Concentration of the
`aqueous cyclodextrin solution. c The solubility in the aqueous cyclodextrin solution divided by the solubility in water. d pH 7.4
`
`form.37 The Kc for the phenytoin-(cid:226)-cyclodextrin complex is
`over 3 times larger for the un-ionized form than for the anionic
`form.46 However, it is frequently possible to enhance cyclo-
`dextrin solubilization of ionizable drugs by appropriate pH
`adjustments. Thus, the solubilizing effects of both (2-hydrox-
`ypropyl)-(cid:226)-cyclodextrin and dimethyl-(cid:226)-cyclodextrin on dihy-
`droergotamine mesylate have been found to increase with
`decreasing pH (i.e., formation of the cationic form). Both the
`saturation solubility and the slopes of the phase-solubility
`diagrams increase with decreasing pH.73 Similar results have
`been reported for the complexation of phenytoin with (cid:226)-cy-
`clodextrin46 and for the complexation of indomethacin,74
`prazepam, acetazolamide, and sulfamethoxazole75 with (2-
`hydroxypropyl)-(cid:226)-cyclodextrin.
`As mentioned before, it is also possible to enhance com-
`plexation and, thus, the solubilizing effect of cyclodextrins by
`addition of polymers or hydroxy acids to the cyclodextrin
`solutions. It has been shown that polymers, such as water-
`soluble cellulose derivatives and other rheological agents, can
`form complexes with cyclodextrins and that such complexes
`possess physicochemical properties different from those of
`individual cyclodextrin molecules.49,76 In aqueous solutions
`water-soluble polymers increase the solubilizing effect of
`cyclodextrins on various hydrophobic drugs by increasing the
`apparent stability constants of the drug-cyclodextrin com-
`plexes. For example, the solubilizing effect of 10% (w/v) (2-
`hydroxypropyl)-(cid:226)-cyclodextrin solution on a series of drugs and
`other compounds was increased from 12 to 129% when 0.25%
`
`(w/v) poly(vinylpyrrolidone) was added to the aqueous cyclo-
`dextrin solution.49 Water-soluble polymers are also capable
`of increasing aqueous solubilities of the parent cyclodextrins
`without decreasing their complexing abilities, thus making
`them more feasible as pharmaceutical excipients. Likewise,
`addition of hydroxy acids, such as citric, malic, or tartaric acid,
`can enhance the solubilizing effect of cyclodextrins through
`formation of super complexes or salts.67
`It is frequently
`possible to obtain even larger solubilization enhancement by
`applying several methods simultaneously. For instance,
`prazepam is a benzodiazepine with a pKa of about 3.
`(2-
`Hydroxypropyl)-(cid:226)-cyclodextrin has a solubilizing effect on both
`the un-ionized and the ionized form of the drug, and as
`expected, hydroxypropyl methylcellulose has a synergistic
`effect on the solubilization. However, the synergistic effect
`was more pronounced for the ionized form (Figure 2).75
`Finally, pharmaceutical formulations should contain as small
`an amount of cyclodextrin as possible since excess cyclodextrin
`can reduce, e.g., drug bioavailability and preservative efficacy.
`Drug solubility should be determined in the final formulation
`and under normal production conditions to determine if too
`much, or too little, cyclodextrin is being used.
`
`Effect on Drug Stability
`The effects of cyclodextrins on the chemical stability of
`drugs is another useful property of these excipients and has
`been extensively examined in the literature.10 Cyclodextrin
`
`Journal of Pharmaceutical Sciences / 1021
`Vol. 85, No. 10, October 1996
`
`Page 5
`
`

`
`+
`
`+
`
`observed first-order rate constant (kobs) is the weight average
`of the two rate constants:
`
`d[D]t
`dt
`
`) -kobs[D]t
`
`and
`
`kobs ) koff + kc(1 - ff)
`
`where ff is the fraction of free drug in solution, or
`
`ff )
`
`1
`1 + Kc[CD]
`
`A Lineweaver-Burk type of equation25 can be obtained by
`further manipulations of the above equations:
`
`1
`ko - kobs
`
`)
`
`1
`Kc(ko - kc)
`
`1
`[CD]
`
`+ 1
`ko - kc
`
`A plot of 1/(ko - kobs) versus 1/[CD] will give rise to a straight
`line (i.e. if the assumption of a 1:1 complex is correct) with a
`y-intercept equal to 1/(ko - kc) and a slope equal to 1/Kc(ko -
`kc), from which the values of kc and Kc can be derived. At low
`concentration most drug-cyclodextrin complexes are of 1:1
`stoichiometry. Even complexes which are of higher order
`stoichiometry at high cyclodextrin and/or drug concentration
`form 1:1 complexes at lower concentration. The stoichiometry
`(i.e., the guest:host molar ratio) will, however, affect the
`stabilizing/destabilizing effect of the complexation.77-80 Thus,
`at relatively high concentrations the antiallergic drug, tra-
`nilast, forms a 2:1 (guest:host) complex with (cid:231)-cyclodextrin
`which accelerates the drug degradation (dimerization) by
`approximately 5500-fold. With increasing (cid:231)-cyclodextrin con-
`centrations, 1:1 and 1:2 complexes are formed resulting in a
`decreased rate of dimerization.77 The rate of dimerization
`was, for example, 19 300 times slower within the 1:2 complex
`than outside it. Similar observations were made when the
`degradation rate of some pilocarpine prodrugs was studied
`in aqueous (2-hydroxypropyl)-(cid:226)-cyclodextrin solutions.80 As
`mentioned before, the enthalpy of the system decreases during
`complex formation, resulting in increased complexation (i.e.,
`increased Kc value) when the temperature is lowered. T

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