throbber
Advanced Drug Delivery Reviews 48 (2001) 115–136
`
`www.elsevier.com/ locate/drugdeliv
`
`Chemical reactivity in solid-state pharmaceuticals: formulation
`implications
`*
`b
`, Wei Xu , Ann W. Newman
`
`c
`
`Stephen R. Byrn
`
`a ,
`
`aPurdue University,West Lafayette, IN 47907,USA
`bMerck and Co.,West Point, PA,USA
`cSSCI, Inc.,3065 Kent Ave,West Lafayette, IN 47906,USA
`
`Received 18 October 2000; accepted 21 December 2000
`
`Abstract
`
`Solid-state reactions that occur in drug substances and formulations include solid-state phase transformations, dehydration/
`desolvation, and chemical reactions. Chemical reactivity is the focus of this chapter. Of particular interest are cases where
`the drug-substance may be unstable or react with excipients in the formulation. Water absorption can enhance molecular
`mobility of solids and lead to solid-state reactivity. Mobility can be measured using various methods including glass
`transition (T ) measurements, solid-state NMR, and X-ray crystallography. Solid-state reactions of drug substances can
`g
`include oxidation, cyclization, hydrolysis, and deamidation. Oxidation studies of vitamin A, peptides (DL-Ala-DL-Met,
`N-formyl-Met-Leu-Phe methyl ester, and Met-enkaphalin acetate salt), and steroids (hydrocortisone and prednisolone
`derivatives) are discussed. Cyclization reactions of crystalline and amorphous angiotensin-converting enzyme (ACE)
`inhibitors (spirapril hydrochloride, quinapril hydrochloride, and moexipril) are presented which investigate mobility and
`chemical reactivity. Examples of drug-excipient interactions, such as transacylation, the Maillard browning reaction, and acid
`base reactions are discussed for a variety of compounds including aspirin, fluoxitine, and ibuprofen. Once solid-state
`reactions are understood in a pharmaceutical system, the necessary steps can be taken to prevent reactivity and improve the
`stability of drug substances and products.
`2001 Elsevier Science B.V. All rights reserved.
`
`Contents
`
`1. Introduction ............................................................................................................................................................................
`2. Solid-state reactions ................................................................................................................................................................
`3. Mobility .................................................................................................................................................................................
`4. Solid-state reactions of drug substances ....................................................................................................................................
`4.1. Oxidation reactions ..........................................................................................................................................................
`4.1.1. Vitamin A .............................................................................................................................................................
`4.1.2. Peptides.................................................................................................................................................................
`4.1.3. Steroids .................................................................................................................................................................
`4.2. Cyclization and hydrolysis ................................................................................................................................................
`4.3. Deamidation and hydrolysis ..............................................................................................................................................
`5. Drug excipient interactions in formulations ...............................................................................................................................
`
`116
`116
`117
`118
`118
`118
`119
`121
`125
`127
`129
`
`*Corresponding author. Tel.: 11-764-494-1460; fax: 11-765-494-6545.
`E-mail address: sbyrn@pharmacy.purdue.edu (S.R. Byrn).
`
`0169-409X/ 01/ $ – see front matter
`PII: S0169-409X( 01 ) 00102-8
`
`2001 Elsevier Science B.V. All rights reserved.
`
`Mylan Ex 1033, Page 1
`
`(cid:211)
`(cid:211)
`

`
`116
`
`S.R. Byrn et al. / Advanced Drug Delivery Reviews 48(2001)115–136
`
`5.1. Transacylation .................................................................................................................................................................
`5.2. Maillard reaction..............................................................................................................................................................
`5.3. Solid-state acid base reactions ...........................................................................................................................................
`6. Conclusions ............................................................................................................................................................................
`References ..................................................................................................................................................................................
`
`129
`130
`130
`134
`135
`
`1. Introduction
`
`Solid-state reactions can and do occur in drug
`substances and formulations. In formulations, solid-
`state reactions of the drug substance are of great
`interest. These can occur in cases where the drug
`substance is intrinsically chemically reactive or
`unstable. In such cases, the formulation can acceler-
`ate degradation in any or all of the following ways:
`
`• Acceleration due to interaction with excipients
`• Acceleration due to processing effects
`• Acceleration induced by excipients (but not in-
`volving chemical reactions with the excipient)
`
`Often, acceleration of reaction is due to the creation
`or presence of amorphous material. Thus, one of the
`best examples of
`these effects is the enhanced
`chemical reactivity of amorphous materials. In such
`cases processing or, possibly, simply interaction with
`the excipients can increase the amount of amorphous
`drug substance. This amorphous drug substance will
`then react due to its increased mobility and ability to
`interact with moisture.
`Direct reaction of excipients with the drug sub-
`stance can also occur, such as solid–solid acid base
`reactions. However, buffers or acids and bases are
`used to stabilize drug substances in formulations as
`well. In many cases these effects can be determined
`by mixing the drug substance with various excipi-
`ents. In other cases processing is needed to induce
`the reaction.
`The role of solid-state reactions in drug substances
`and formulations will be discussed here. Background
`information on solid-state reactions and mobility, as
`well as various examples of solid-state reactions in
`pharmaceutical applications will be presented.
`
`2. Solid-state reactions
`
`In order to understand more about the influence of
`formulations on solid state reactions it is worthwhile
`
`to review solid state reactions. Solid-state reactions
`in their broadest sense include solid-state phase
`transformations (polymorphic transformations), re-
`actions in which solvent of crystallization is lost or
`gained, and a broad range of solid state chemical
`reactions. Most of the emphasis of this chapter will
`be on chemical reactivity.
`It is necessary to establish criteria for solid state
`reactions in order
`to focus on true solid state
`reactions. This will avoid a liquid state reaction
`being identified as a solid state reaction. Morawetz
`suggested four criteria for determining whether a
`reaction is a true solid state reaction [1]. A fifth
`criterion can be added based on Paul and Curtin [2].
`A reaction occurs in the solid when:
`
`1. the liquid reaction does not occur or is much
`slower.
`2. pronounced differences are found in the reactivity
`of closely related compounds.
`3. different
`reaction products are formed in the
`liquid state.
`in different crystalline modi-
`4. the same reagent
`fications has different reactivity or leads to differ-
`ent reaction products.
`5. it occurs at a temperature below the eutectic point
`of a mixture of the starting material and products.
`
`the reaction is
`Once it has been established that
`occurring in the solid-state,
`the reaction can be
`understood in terms of a four step process [2]:
`
`1. Loosening of Molecules at the Reaction Site. It is
`reasonable to assume that molecular loosening
`achieves the mobility required to accomplish the
`next step.
`2. Molecular Change. This step is similar to the
`corresponding solution reaction where the bonds
`of the reactant are broken and the bonds of the
`product are formed.
`3. Solid Solution Formation. During the early stages
`of the reaction, a solid solution of the product in
`
`Mylan Ex 1033, Page 2
`
`

`
`S.R. Byrn et al. / Advanced Drug Delivery Reviews 48(2001)115–136
`
`117
`
`the site of
`is formed at
`the starting crystal
`reaction. However, after the concentration of the
`product reaches a certain point, the product will
`separate.
`4. Separation of Product. This step gives new
`crystals, either
`randomly oriented or with an
`orientation governed by the crystals of the starting
`material. This latter case is termed a topotactic
`reaction and will be discussed.
`
`Solid-state reactions begin at one or more nucleation
`sites and spread through the crystal. For desolvations
`and some thermal reactions, the reaction begins at a
`nucleation site and spreads through the crystal in a
`front that advances through the crystal.
`Nucleation sites (defects)
`for
`reaction are de-
`veloped during crystallization or can sometimes be
`produced by mechanical deformations such as prick-
`ing with a pin or cutting the crystal. Exposing the
`starting crystal
`to product crystals may produce
`nucleation sites. In other cases, nucleation is random
`and neither mechanical deformation nor exposure to
`product crystals nucleates the reaction. Obviously,
`this variability in nucleation and the random number
`of nucleation sites that are present in crystals can
`greatly complicate the kinetics of solid state re-
`actions.
`For solid gas reactions, diffusion of gas into the
`crystal itself requires molecular loosening. Thus, for
`solid–gas reactions simultaneous or sequential mo-
`lecular loosening and diffusion steps are involved.
`The chemical reaction is generally considered to
`follow the same mechanism as the solution reaction.
`As the reaction proceeds a solid solution of the
`product in the reactant crystal will be formed. After
`the limit of product solubility in the starting crystal
`lattice is reached, the product will either crystallize
`or continue to accumulate in an amorphous form.
`This step would not influence the apparent rate of the
`reaction as measured by the amount of product
`formed or by the disappearance of the reactant.
`However, if the rate is measured using the diffraction
`intensities of crystalline product, the rates may differ
`from those measured chemically.
`It is clear from the above discussion that the rates
`of solid state reactions depend on several factors,
`including nucleation and the molecular changes
`involved. It is important to realize that the crystal
`structure and crystal packing profoundly affect the
`
`molecular loosening and molecular change steps of a
`solid state reaction. The crystal packing determines
`the extent of molecular loosening required for the
`molecules to reorient sufficiently to undergo the
`required molecular changes. The crystal packing also
`determines the extent of molecular loosening re-
`quired for gases to diffuse to the reaction site in
`solid–gas reactions. Thus, determining the relevant
`crystal structures and investigating the molecular
`mobility in these structures can bring about a great
`deal of insight into solid state reactions.
`
`3. Mobility
`
`Understanding the mobility of groups in solids can
`lead to insight
`into the mobility of molecules in
`solution, the forces responsible for conformational
`interconversions, and the factors responsible for solid
`state reactions. It is clear that solid state degradations
`of pharmaceuticals are often related to molecular
`mobility [3–5]. In addition, Ahlneck and Zografi [6]
`have suggested that water absorption enhances the
`molecular mobility of pharmaceutical solids, perhaps
`explaining the enhanced chemical reactivity of these
`materials in the presence of water. Recent studies by
`Zografi’s laboratory and our laboratory confirm the
`relationship between molecular mobility and solid
`state reactivity as discussed in a later section.
`Further study of the relationship between molecu-
`lar mobility and solid state reactivity requires the
`development of new approaches to determining
`molecular mobility, especially in mixtures such as
`pharmaceutical dosage forms where single-crystal
`X-ray methods cannot be used.
`Solid state NMR offers several attractive ap-
`proaches to the study of the molecular mobility of
`solids. These include:
`
`1. Determination of the activation energies from T1
`relaxation of individual carbon atoms using vari-
`able temperature solid-state NMR.
`2. Study of processes which result in peak coalesc-
`ence of solid-state NMR resonances using vari-
`able temperature solid-state NMR.
`3. Use of interrupted decoupling to detect methylene
`and possible methine groups with unusual mobili-
`ty.
`
`Mylan Ex 1033, Page 3
`
`

`
`118
`
`S.R. Byrn et al. / Advanced Drug Delivery Reviews 48(2001)115–136
`
`4. Comparison of solid-state MAS spectra measured
`with and without cross polarization.
`
`X-ray crystallography offers another approach to the
`measurement of mobility in the solid-state. Trueb-
`lood and Dunitz have used variable temperature
`X-ray crystallography to measure the force constants
`for vibration of methyl groups [7–10].
`It is interesting to explore the possible relationship
`between studies of molecular mobility by solid-state
`NMR and X-ray crystallography. A priori there is no
`reason to expect that the force constants for methyl
`vibration (or any other motion) determined by X-ray
`crystallography and the activation energy for methyl
`rotation determined by solid state NMR would be
`correlated. Variable temperature solid state NMR is
`thought to measure activation energies for the rota-
`tion of methyl groups about their C axes. (However,
`3
`many of the barriers obtained for methyl rotation
`from solid state NMR are significantly lower than
`those expected based on molecular mechanics calcu-
`lations of methyl rotation barriers.) In contrast, X-ray
`crystallography measures the vibrational motion of
`the individual atoms. One would expect
`that vi-
`bration (as measured by X-ray crystallography) and
`rotation of methyl groups (as determined by solid-
`state NMR) in solids may not always be correlated.
`On the other hand,
`if the crystal packing in the
`vicinity of two methyl groups in the same molecule
`is different, one might expect that the methyl group
`that is not as tightly packed might exhibit a lower
`force constant for vibration and also a lower activa-
`tion energy for rotation. In our laboratory we have
`compared the relative barriers to methyl rotation of
`the multiple methyl groups
`in ibuprofen and
`phenacetin measured by solid state NMR to the force
`constants measured by variable temperature X-ray
`crystallography. These studies
`indicate that
`the
`methyl group with the greatest barrier to motion by
`solid state NMR also has the greatest force constant
`for vibration by X-ray crystallography.
`Water is known to enhance the mobility of amor-
`phous solids [6]. This is due to the plasticization that
`results when water is absorbed. In addition, water
`can be absorbed into amorphous regions in otherwise
`crystalline materials. Such solids that may contain a
`small or undetectable amount of amorphous material
`are expected to show enhanced reactivity. Once
`
`degradation begins in these small amorphous regions
`it can proceed throughout
`the crystal by forming
`eutectic melts of the products and reactants. Ahlneck
`and Zografi have pointed out that the effect of water
`is amplified in these cases because the small amounts
`of amorphous material can contain a relatively large
`amount of water; yet the total water content of the
`solid will be low [6]. Further studies of the mobility
`of solids in the presence and absence of water are
`needed in order to determine whether a predictive
`technique can be developed which will form the
`basis for selection of the most stable crystal form
`prior to stability studies.
`
`4. Solid-state reactions of drug substances
`
`The origin of solid-state reactions in a drug
`substance can be explained by various factors as
`discussed above. Solid-state reactions can include
`oxidation, cyclization, hydrolysis, and deamidation.
`Excipients present in formulated products may not be
`directly involved in the degradation mechanism, but
`may add parameters, such as water, which contribute
`to the solid-state reaction. Examples of degradation
`of drug substances alone and in formulations will be
`presented.
`
`4.1. Oxidation reactions
`
`4.1.1. Vitamin A
`Solid-state oxidation reactions have been known
`for many years. Early studies suggested that these
`reactions were free radical processes. For example,
`Finkel’shtein and co-workers postulated the mecha-
`nism for
`the reaction of beta carotene (Fig. 1)
`because 2,6 di-tert-butyl-4-methyl phenol (BHT) and
`other antioxidants inhibit this reaction and the rate
`depends on oxygen pressure and temperature [11].
`Crystalline esters of vitamin A (including the hemi
`succinate,
`the nicotinate, and the 3,4,5-trimethox-
`ybenzoate) decompose by both polymerization and
`oxidation pathways [12,13].Vitamin A exposed to air
`at room temperature for several years or heated at
`1008C for 5 h gave at least five ketones on TLC
`plates treated with 2,4-dinitrophenylhydrazine [14].
`We have found that vitamin A is a gummy yellow
`solid after 5 months of exposure to room light,
`
`Mylan Ex 1033, Page 4
`
`

`
`S.R. Byrn et al. / Advanced Drug Delivery Reviews 48(2001)115–136
`
`119
`
`cated because it produces a variety of products. Xu
`[18] has
`carried out
`a preliminary study of
`methionine oxidations by monitoring the disappear-
`ance of
`the peptide as well as monitoring the
`formation of the sulfoxide and sulfone products, as
`shown in Fig. 2.
`Three model peptides (Fig. 3) were investigated:
`DL-Ala-DL-Met (a zwitterionic dipeptide), N-formyl-
`Met-Leu-Phe methyl ester (a neutral tripeptide), and
`Met-enkaphalin acetate salt (a weakly acidic penta-
`peptide). These peptides were chosen for investiga-
`tion because both crystalline and amorphous material
`could be produced. To complete the study in a
`reasonable time, the oxidation of these peptides was
`accelerated by exposure of
`the peptides to UV
`radiation (254 nm) using a 15 W tube. Fig. 4 shows
`the results for
`the study of
`the degradation of
`crystalline and amorphous DL-Ala-DL-Met. Similar
`
`Fig. 2. Methionine oxidation reactions [18].
`
`Fig. 3. Model peptides DL-Ala-DL-Met (a zwitterionic dipeptide),
`N-formyl-Met-Leu-Phe methyl ester (a neutral
`tripeptide), and
`Met-enkaphalin acetate salt (a weakly acidic pentapeptide).
`
`Fig. 1. Autooxidation scheme of b-carotene [11].
`
`temperature, and air. Elemental analysis indicated
`that each molecule of vitamin A took up six oxygen
`atoms, and mass spectral studies indicated extensive
`degradation and the presence of many products.
`Diluents with antioxidant properties stabilize vita-
`min A palmitate in vitamin preparations
`[15].
`Aluminum salts of fatty acids such as stearic acid
`stabilize vitamin A, as does combination with gelatin
`and dextrin, which probably contain reducing sugars
`[16].
`An interesting correlation of the melting points of
`the vitamin A (retinoic acid) esters with their zero
`order rate of solid state decomposition has been
`observed [13]. As the melting point increased, the
`rate of decomposition decreased. The rates of de-
`composition of these esters in solution were virtually
`identical. These results were interpreted in terms of
`crystal lattice energy. It was argued that the higher
`melting esters had more crystal-lattice energy and
`thus were more stable to the solid–gas oxidation
`reaction. Thus, the higher the melting point the more
`efficient
`the packing and, conceivably,
`the less
`permeable the crystal is to reacting gas. A better
`measure of lattice energies in a series of compounds
`is based on the heat of sublimation [17].
`
`4.1.2. Peptides
`There is substantial interest in understanding more
`about
`the solid-state oxidation of methionine in
`peptides and proteins. One impediment to studying
`this oxidation reaction is that the process is compli-
`
`Mylan Ex 1033, Page 5
`
`

`
`120
`
`S.R. Byrn et al. / Advanced Drug Delivery Reviews 48(2001)115–136
`
`Fig. 4. UV induced degradation (left) and oxidation (right) of crystalline and amorphous DL-Ala-DL-Met at ambient conditions [18].
`
`studies of N-formyl-Met-Leu-Phe methyl ester, and
`Met-enkaphalin acetate salt were also carried out but
`are not shown. These studies show that amorphous
`material oxidizes and degrades faster than crystalline
`material in each of these cases.
`To rule out particle size as an explanation for the
`enhanced reactivity of the amorphous material, crys-
`talline DL-Ala-DL-Met was partitioned into different
`size fractions and each fraction was oxidized. Fig. 5
`shows the results of this study. It is clear from this
`
`figure that particle size does not affect the reactivity
`of the crystalline material. Regardless of the particle
`size the crystalline material is much less reactive
`compared to the amorphous material.
`This preliminary study shows that amorphous
`peptides are much more reactive towards oxygen
`than their crystalline counterparts and that the differ-
`ence in reactivity is probably not due to particle size.
`It
`is likely that
`the amorphous material
`is more
`reactive because it has more mobility.
`
`Fig. 5. The effect of particle size on the UV-induced Met oxidation of DL-Ala-DL-Met at ambient conditions [18].
`
`Mylan Ex 1033, Page 6
`
`

`
`S.R. Byrn et al. / Advanced Drug Delivery Reviews 48(2001)115–136
`
`121
`
`4.1.3. Steroids
`Hydrocortisone modifications have been exten-
`sively studied to investigate the stability of steroids.
`Hydrocortisone 21-tert-butylacetate (HTBA) yielded
`40% of
`the 11-keto product, cortisone 21-tert-
`butylacetate (CTBA), upon standing at room tem-
`perature for 1–2 years (Fig. 6) [19,20]. In contrast,
`other esters of this steroid, including the 21-ethyl
`ester, were completely resistant to air oxidation, even
`after 15 years at room temperature. The oxidation of
`hydrocortisone 21-tert butylacetate is accelerated by
`heat and greatly accelerated by free radical initiators
`and ultraviolet light.
`In our laboratory, HTBA has been obtained in five
`crystalline forms [21]. Forms I, II, and III were
`obtained from absolute ethanol solution either at
`
`Fig. 6. Oxidation of hydrocortisone 21-tert-butylacetate.
`
`Table 1
`Crystal forms of hydrocortisone 21-tert-butylacetate
`
`room temperature or in the refrigerator. X-ray pow-
`der diffraction studies indicated that
`they were
`different forms and elemental analysis showed that
`they contained varying amounts of ethanol within the
`crystals.
`The results of these studies are shown in Table 1.
`During crystallization from ethanol a mixture of
`Forms I, II, and III often appeared, but a pure single
`form could be obtained under certain conditions. A
`new form designated Form IV was produced when
`Forms I, II, and III were heated at 1208C. Forms I
`and II underwent desolvation and phase transforma-
`tion to Form IV, while Form III appeared to trans-
`form via a melt recrystallization process to Form IV.
`However, this observation needs to be substantiated
`by another method of analysis. Another form (Form
`V) was isolated from pyridine. All crystal forms,
`except Forms I and V, were inert to irradiation with
`ultraviolet light.
`Form I was oxidized from HTBA to cortisone
`21-tert-butylacetate (CTBA) under irradiation with
`ultraviolet light in air. A known weight and given
`size of crystals were put in vials and irradiated at
`308C. The extent of formation of CTBA was de-
`termined by integrating the C methyl NMR signal,
`18
`and the content of ethanol was measured by gas
`
`Form
`
`Crystal class
`Space group
`˚a (A)
`˚b (A)

`c (A)
`a
`b
`g
`Solvent of
`crystallization
`
`EtOH content
`a
`mp(8C)
`UV oxidation
`
`I
`
`hexagonal
`P6
`1
`17.485
`17.485
`15.376
`908
`908
`1208
`ethanol
`propanol
`n-amyl alcohol
`acetonitrile
`d
`0.9
`170–180
`reaction
`
`II
`
`monoclinic
`P2
`1
`12.440
`7.710
`14.724
`908
`88.78
`908
`ethanol
`
`1.0
`110–120
`no reaction
`
`b
`
`III
`
`triclinic

`P1
`23.0
`12.5
`29.0
`748
`1478
`748
`ethanol-
`tert-butanol
`
`–
`123–126
`no reaction
`
`c
`
`IV
`
`V
`
`unstable
`
`heat forms
`I, II, or III
`
`–
`234–238
`no reaction
`
`pyridine
`
`–
`
`reaction
`
`a The exact melting temperature may vary from one crystal to another.
`b Opaque at this temperature range, with final melting at 234–2388C.
`c After melting, the melt resolidified as the temperature was rising and finally remelted at 234–2388C.
`d When crystallized from ethanol, Form I contains ethanol; when crystallized from the other solvents, no solvent of crystallization is
`present [21].
`
`Mylan Ex 1033, Page 7
`
`

`
`122
`
`S.R. Byrn et al. / Advanced Drug Delivery Reviews 48(2001)115–136
`
`Table 2
`Percent desolvation and oxidation of crystalline hydrocortisone
`a
`21-tert-butylacetate 0.9 ethanolate upon exposure to UV light
`
`Days
`
`% EtOH lost
`
`% Cortisone formed
`
`1
`2
`3
`6
`10
`14
`21
`
`43.3
`75.6
`83.3
`88.9
`93.3
`95.6
`96.7
`
`20.0
`38.9
`50.0
`52.9
`56.3
`66.7
`71.4
`
`a
`
`From Ref. [21].
`
`chromatography. The percent desolvation and oxida-
`tion of HTBA are shown in Table 2. Although the
`loss of ethanol was faster than the oxidation, this
`desolvation–oxidation behavior is different from that
`of dihydrophenylalanine hydrate [22],
`in which
`nearly complete loss of water precedes the oxidation.
`In addition, the crystal opaqueness phenomenon of
`desolvation has not been found in crystal Form I.
`During desolvation, a significant change in appear-
`ance of Form I (or other similar crystals studied) can
`be observed only under polarized light. After a week,
`a circular pattern of decolorization moved from the
`ends toward the center, but the crystal remained clear
`under transmitted light even after 42 days. Crystallo-
`graphic studies show that desolvated Form I is still a
`single crystal with a reasonably good diffraction
`pattern.
`The crystal structure of Form I has been de-
`termined. The structure of this form is consistent
`
`with structures of other hydrocortisone derivatives
`[23]. It is proposed that the reactivity of HTBA Form
`I toward oxygen is due to the crystal packing, which
`allows penetration of oxygen down the axis of the
`helix of the crystal. It is further proposed that ethanol
`of crystallization is normally along this axis, and that
`the exit of the ethanol from the crystal further aids
`oxygen penetration. However,
`further crystallo-
`graphic studies are required to confirm these hypoth-
`eses.
`Our laboratory carried out a study of an additional
`series of hydrocortisone esters [4]. Table 3 summa-
`rizes the results of
`these studies and lists the
`crystallographic data, whether solvent of crystalliza-
`tion was located in the crystal and the type of crystal.
`Three types of crystal forms were defined as:
`
`1. Type A crystal forms are nonstoichiometric sol-
`vates that are reactive in the presence of oxygen.
`2. Type B crystal forms are stoichiometric solvates
`that are unreactive in the presence of oxygen.
`3. Type C crystal forms do not contain solvent and
`are also unreactive in the presence of oxygen.
`
`The structure of both crystal forms of hydrocortisone
`21-butanoate provide good examples of these esters.
`Crystals of Form I hydrocortisone 21-butanoate were
`crystallized from toluene and was found to belong to
`space group P2 2 2 [4]. Form II of hydrocortisone
`1 1 1
`21-butanoate was crystallized from 2-propanol and
`found to be hexagonal, belonging to the space group
`
`Table 3
`Summary of cell parameters and crystal data for hydrocortisone esters studied
`
`Ester
`
`Type
`
`a
`
`Solvent
`
`Space
`group
`
`a
`
`b
`
`c
`
`a
`
`b
`
`g
`
`propanoate Form I
`butanoate Form I
`butanoate Form II
`pentanoate Form I
`hexanoate Form I
`cyclopentylacetate
`Form I
`butanoate, 9a-fluoro
`Form I
`pentanoate, 9a-fluoro
`Form I
`
`C
`C
`A
`A
`A
`A
`
`A
`
`A
`
`IPA
`PhMe
`IPA
`EtOH
`EtOH
`EtOH
`
`EtOH
`
`EtOH
`
`1
`
`P2 2 2
`1 1 1
`P2 2 2
`1 1 1
`P6
`P6
`1
`P6
`P6
`
`1
`
`1
`
`16.807
`17.011
`17.704
`17.991
`
`16.810
`17.004
`17.736
`17.980
`
`15.117
`15.194
`15.035
`15.165
`
`P6
`
`1
`
`P6
`
`1
`
`16.832
`
`16.827
`
`15.135
`
`17.188
`
`17.193
`
`15.141
`
`z
`
`4
`4
`6
`6
`6
`6
`
`6
`
`6
`
`Final R
`
`Solvent
`located
`
`0.079
`0.084
`0.089
`0.075
`0.088
`0.098
`
`0.057
`
`0.069
`
`No
`No
`No
`Yes
`Yes
`Yes
`
`Yes
`
`Yes
`
`90
`90
`90
`90
`90
`90
`
`90
`
`90
`
`90
`90
`90
`90
`90
`90
`
`90
`
`90
`
`90
`90
`120
`120
`120
`120
`
`120
`
`120
`
`a Type A — reactive, nonstoichiometric solvate; type B — unreactive, stoichiometric solvate; type C — unreactive, unsolvated crystal
`form. All of these compounds give unique XRPD patterns [4].
`
`Mylan Ex 1033, Page 8
`
`

`
`S.R. Byrn et al. / Advanced Drug Delivery Reviews 48(2001)115–136
`
`123
`
`P6 . Hydrocortisone 21-butanoate also crystallized
`1
`in this hexagonal space group from methanol, etha-
`nol, 1-propanol, acetone, ethyl acetate, tert-butylace-
`tate, and DMF. Crystals formed from all of these
`solvents show identical X-ray powder diffraction
`patterns. These forms contain the steroid molecules
`in a helix with a tunnel down the center.
`Both forms of the 21-butanoate ester were exposed
`to UV light at 254 nm for 15 days and analyzed
`using both NMR and HPLC for oxidation to the
`corresponding cortisone 21-ester. Analysis showed
`that Form II (hexagonal form) oxidized, while Form
`I (orthorhombic form) did not oxidize. Under the
`same conditions, Form I (orthorhombic form) of
`hydrocortisone 21-propanoate (isostructural to Form
`I of the 21-butanoate ester) also did not oxidize.
`Moreover, all of the other hexagonal crystal forms
`listed in Table 3 oxidized under these conditions.
`Note that we prefer to classify these forms as Type A
`even though these forms are stoichiometric solvates
`containing a 1:2 ratio of solvent:hydrocortisone ester.
`This is because it is difficult to make the distinction
`between stoichiometric and nonstoichiometric sol-
`vates in hexagonal steroids.
`the maximum
`Initial experiments showed that
`amount of oxidation of powdered samples of Form II
`was about 30% regardless of their time of exposure
`to UV light. This is consistent with the reactivity of
`hydrocortisone tert-butylacetate. It is hypothesized
`that
`this happens because oxidation is limited to
`molecules near the surface of the solid. Form II
`(hexagonal) crystals were exposed to UV light and
`washed with chloroform. Both the washings and the
`
`washed crystals remaining were analyzed using
`HPLC. Oxidation product was found in the washings
`but not in the residual crystals, indicating that the
`inner parts of these crystals were not oxidized. This
`result further supports the hypothesis that oxidation
`is limited to molecules near the surface.
`This surface oxidation phenomenon seems to be
`quite odd for the hexagonal crystal form, especially
`with the presence of a tunnel through the center of
`the unit cell and the close proximity of the 11-b-OH
`to the inner surface of the tunnel. One hypothesis
`that may explain this result is that the tunnel is too
`small to allow oxygen to penetrate the crystal. To
`test this hypothesis, the cross sectional area of this
`tunnel was measured. However, simply measuring
`this area by measuring the tunnel area on a projec-
`tion perpendicular to the c-axis is not appropriate
`because of the helical nature of the crystal packing.
`In doing so, only a small area would be measured
`because of the overlapping of atoms from various
`molecules in the zigzag tunnel. In order to obtain a
`more accurate approximation to the true cross sec-
`tional area, the c-axis of the tunnel is cut into 3

`layers (Fig. 7). Each layer is 5.04 A deep (3 of the
`unit cell c dimension) along the c-axis and consists
`of different parts of each molecule.
`Hydrogen atoms lining the tunnel were used in
`calculated positions. The area was determined by
`delineating the boundaries of the tunnel and de-
`termining the area by the cut-and-weigh method. The
`cross-sectional area of each layer was determined to
`2˚
`be 23.9 A . This area is obviously large enough for
`molecular oxygen to pass in and out of the tunnel.
`
`Fig. 7. Cross sectional areas of the three layers of the tunnel down the c-axis of hydrocortisone 21-butanoate Form I [24].
`
`Mylan Ex 1033, Page 9
`
`

`
`124
`
`S.R. Byrn et al. / Advanced Drug Delivery Reviews 48(2001)115–136
`
`Therefore, a narrow tunnel is not the cause for the
`limited oxidation in this hexagonal crystal form.
`An alternative hypothesis is that the presence of
`solvent of crystallization in the tunnel of the hexa-
`gonal crystal form prevents internal oxidation. Crys-
`tallographic studies

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