throbber
Pharmaceutical Research, Vol. 8, No. 10, 1991
`
`Research Article
`
`Photoreactivity of LY277359 Maleate, a 5-Hydroxytryptamine3
`(5-HT3) Receptor Antagonist, in Solution
`
`Gerold L. Mosher" and Julianne McBee'
`
`Received November 8, 1990; accepted May 14, 1991
`
`Compound LY277359 maleate undergoes a photoinduced solvolysis reaction in water to generate the
`corresponding hydroxylated product and release chloride. Attempts to stabilize a parenteral formu-
`lation of the compound led to an investigation of possible reaction mechanisms. The data are consis-
`tent with a mechanism involving homolytic cleavage of the aryl—chloride bond followed by electron
`transfer to give an aryl cation intermediate. The cation thus formed reacts with surrounding nucleo-
`philes to give the substituted product. A kinetic expression for reaction rate was derived from the
`mechanism, and various components of the rate constant were evaluated experimentally. The reaction
`is slowed with the addition of chloride, presumably via a common ion effect (enhanced retroreaction).
`In the absence of added chloride, the reaction can be described kinetically by an initiation term. An
`inner filter effect is also observed, where increasing amounts of the hydroxylated product slow the
`reaction. Experimental data for observed rate constants as a function of starting concentration and
`light intensity are fit with good correlation to an equation describing the filter effect. Additional studies
`evaluated the effects of various structural features of the parent compound on the rate of the reaction
`in glass containers. It was determined that reactivity was dependent on two features: (1) the ortho
`positioning of the carboxyl and ether groups, which shifted an absorption band above the container
`cutoff; and (2) the para orientation of the chloro group to the ether, which is para activating in the
`photoexcited state.
`
`KEY WORDS: photosubstitution; aryl cation; inner filter; kinetics; common ion effect.
`
`INTRODUCTION
`
`Preformulation studies on LY277359 maleate (I), a po-
`tent and selective 5-HT3 receptor antagonist, were compli-
`cated when we discovered that aqueous solutions were un-
`stable in the presence of light. The use of amber vials slowed
`the reaction but did not impart sufficient stability for formu-
`lation. Analysis of the reaction mixture by high-performance
`liquid chromatography (HPLC)/mass spectrometry and nu-
`clear magnetic resonance (NMR) indicated that the only
`product formed was the substituted product (II) (Scheme I).
`The complete absence of the corresponding reduction prod-
`uct (III) was interesting, and this fact along with the need for
`a stable parenteral formulation of compound I led us into
`further studies to explore the nature of this reaction.
`In the study reported here, we examine the impact of
`chemical structure and reaction conditions on the photosub-
`stitution reaction and investigate possible reaction mecha-
`nisms. The overall goal was to understand the reaction better
`so that appropriate measures could be undertaken to stabi-
`lize a parenteral formulation of compound I.
`
`' Lilly Research Laboratories, Eli Lilly and Co., Indianapolis, In-
`diana 46285.
`2 To whom correspondence should be addressed.
`
`ss‘
`
`LY277359 maleate (I) X,
`
`Scheme I
`
`MATERIALS AND METHODS
`
`Materials
`
`Compounds I (endo-5-chloro-2,3-dihydro-2,2-dimethyl-
`N- (8-met h yl- 8 - a z ab icy clo [3 . 2 . llo c t - 3 - y1)- 7 - b e n -
`
`1215 (cid:9)
`
`0724-8741/91/1000-1215$06.50/0 (cid:9)
`
`1991 Plenum Publishing Corporation
`
`MYLAN Ex. 1013, Page 1
`
`

`
`1216 (cid:9)
`
`Mosher and McBee
`
`zofurancarboxamide (Z)-2-butendioate), II, IV, and VIII-X
`were used as received from Lilly Research Laboratories (1).
`Compounds V-VII, XI, and XII were obtained from Aldrich.
`HPLC-grade acetonitrile was used and all other chemicals
`were reagent grade. All solutions and buffers were prepared
`using water for injection, and all glassware used for drug
`solutions was silanized to prevent drug adsorption. Twenty-
`milliliter scintillation vials (Kimble), --280-nm cutoff, were
`used as reaction vessels in the light stability studies.
`
`UV Absorption Spectra
`
`A Beckman DU-70 spectrophotometer was used for UV
`absorption spectra using a 1-cm-path length cell and 60-nm/
`min scan speed. Solutions were scanned from 500 to 190 nm.
`The solvent of interest was used for calibration and back-
`ground correction.
`
`HPLC Analysis
`
`LY277359 maleate and its photoreaction products were
`separated isocratically on a DuPont Zorbax TMS column
`(4.6 mm x 25 cm) at ambient temperature. The mobile phase
`was acetonitrile:0.125 M (NH4)2SO4, pH 4.0 (28:72), deliv-
`ered at a 1.5-ml/min flow rate with a Spectra-Physics Model
`8810 pump. Injections were made using a Bio-Rad autosam-
`pler (Model AS-48) with a 20-0 fixed loop operated at 4°C.
`The sample chamber of the autosampler was covered with
`aluminum foil to protect the samples from light. Effluent was
`monitored on an Applied Biosystems Spectraflow UV ab-
`sorbance detector (Model 783A) at a wavelength of 235 nm
`with peak areas being determined on a Hewlett-Packard
`3396A integrator. Concentrations were calculated from ex-
`ternal standard curves generated from aqueous drug solu-
`tions prepared in silanized glassware.
`
`Kinetic Studies
`
`A clear, incandescent, 100-W bulb (General Electric)
`was used as the light source and a light chamber was con-
`structed to fix the distance the reaction vessels were placed
`from the bulb. The temperature in the chamber was con-
`trolled at 30 ± 1°C by air circulation. Solutions were pre-
`pared by weighing the compound of interest into volumetric
`flasks, solubilizing with the desired solvent, and decanting
`20 ml into the vessels. For studies with added halogens, KI,
`KBr, KCI, and NaCI were weighed directly into the vessels
`and solubilized with a 0.042 mM solution of compound I to
`make the solutions 0.154 M in halogen. Once prepared, all
`solutions were stored in the dark until needed. The studies
`were conducted by placing the capped vessels in the light
`chamber at a measured distance from the source and taking
`0.5-ml aliquots over time. Reversed-phase HPLC was used
`to assay the aliquots for compound concentration and mon-
`itor the appearance of reaction products. Reactions were
`followed for 6 hr. Unless otherwise noted, all reactions were
`conducted with the vessels placed 10 cm from the light
`source.
`Rate-concentration dependency studies were per-
`formed by varying the initial concentration of compound I.
`The effect of chloride ion was studied by monitoring the
`kinetics of the reaction in solutions that were 0.017, 0.077,
`0.154, and 0.308 M in chloride and 0.1 mM in compound I.
`Ionic strength effects were evaluated in a similar manner by
`using NaC104 in various concentrations up to 0.2 M.
`Reactions run in water/acetonitrile solutions were pre-
`pared by solubilizing compound I in a known amount of
`water and bringing to volume with acetonitrile. The solutions
`evaluated were 55.6, 41.7, 27.8, and 13.9 M in water and 0.1
`mM in compound I.
`Varying light intensity was accomplished by placing the
`reaction vessels 5, 10, or 15 cm from the filament of the bulb.
`The effect of compound II on the reaction kinetics of I
`was studied by mixing aqueous solutions (of known concen-
`trations) of I and II prior to light irradiation. Effects of other
`light absorbing organic species were studied by dissolving
`compound I in aqueous 1.1 mM solutions of p-aminobenzoic
`acid (XI) or p-nitrophenol (XII) before initiating the reaction.
`Control reactions were included in each study.
`
`RESULTS
`
`UV Absorbance
`
`The UV absorption spectra of compound I in water
`shows a secondary absorption band at 318 nm. The nature of
`this band was evaluated by comparing the UV absorbance
`spectrum in polar and nonpolar solvents. The direction of
`shift in the X. on changing solvents can be used as an
`indication of n-7r* versus Tr-Tr* transitions (2,3). Upon
`changing the solvent from n-hexane to water, a 12-nm red
`shift was observed in the Xmax. This and the magnitude of the
`molar absorptivity, 4800 cm -1 M - at 318 nm, are indicative
`of a Tr-Tr* (B-band) transition. The wavelength of maximal
`absorption for this band in water is given in Table I for the
`series of structurally related compounds evaluated in the
`present work.
`No appreciable change in molar absorptivity was noted
`when the solvent was changed to methanol, ethanol, n-pro-
`panol, or acetonitrile.
`
`Reaction Kinetics
`
`Representative data from light stability studies are
`shown in Fig. 1. The data demonstrate apparent first-order
`behavior as would be expected for a photochemical process
`in dilute solution. A least-squares regression analysis of the
`data gives the solid line, the slope of which is used to cal-
`culate an observed rate constant (kobs) value. These values
`are compared as the reaction conditions are varied.
`The effect of solvent polarity on the occurrence and rate
`of the substitution reaction is demonstrated rather dramati-
`cally in Table II. Due to possible variations in the reaction
`conditions, (lamp aging, different initial concentrations,
`etc.), a control reaction was run for each set of experiments.
`When the polarity of the medium was reduced by using a
`mixture of methanol in water, the kobs value was reduced by
`half. Two additional peaks appeared on the chromatograms
`from that study but were not isolated and identified. Their
`relative position on the chromatogram suggests the possibil-
`ities of a methoxy substitution and/or the reduction product.
`A similar reduction in kobs values was seen in the pres-
`ence of increasing amounts of acetonitrile in water. Irradi-
`
`MYLAN Ex. 1013, Page 2
`
`

`
`Photoreactivity of LY277359 Maleate in Solution (cid:9)
`
`1217
`
`Table I. Absorbance and Reactivity Data for the Compounds Investigated
`
`Compound (cid:9)
`
`Xmax (nm)
`
`Structure
`
`IV
`
`V
`
`VI
`
`VII (cid:9)
`
`VIII"
`
`IX
`
`X
`
`310b
`322'
`
`282
`
`280
`
`308
`
`300
`
`290
`
`301
`
`0.00446
`0.00692
`
`No reaction
`
`No reaction
`
`No reaction
`
`0.000958
`
`No reaction
`
`No reaction
`
`" Observed rate constant in min - ' from 0.043 mM solutions at 10 cm. kob, = 0.00708 min - for
`compound I.
`b Ionized carbonyl.
`Protonated carbonyl.
`d Maleate salt used.
`
`compound IX, or moving the chloro group to the four posi-
`tion, compound X, destroys reactivity. Likewise, removal of
`the 2,3-dihydrofuran ring, compound V, eliminates reactiv-
`ity. Compounds VI and VII, assessing the importance of an
`oxygen attached para to the halogen and possibly hydroxyl—
`carbonyl interactions, respectively were also nonreactive.
`However, some reactivity was observed with compound
`VIII, which is the opened-ring version of compound I.
`
`DISCUSSION
`
`Aromatic photosubstitution reactions have been under
`
`Table II. Effects of Solvent, Added Halogens, and pH on the Ob-
`served Rate Constant
`
`Solvent
`
`H2O (27.8 M)/CH3OH
`H2O (control)
`H2O (41.7 M)/CH3CN
`H2O (27.8 M)/CH3CN
`H2O (13.1 M)/CH3CN
`CH3CN
`CH3OH
`(CH3)20H
`(CH3)3OH
`
`NaCl (0.154 M)
`KCI (0.154 M)
`KBr (0.154 M)
`KI (0.154 M)
`
`H20, pH 3.0"
`pH 4.4"
`pH 11.0"
`
`[Compound I]
`(mM)
`
`/cabs
`(min - (cid:9) x 103)
`
`0.110
`
`If
`
`0.110
`
`If
`
`0.043
`
`0.043
`
`0.043
`
`0.043
`
`2.4
`4.1
`3.6
`2.4
`0.79
`No reaction
`
`3.9
`4.0
`9.6
`5.9
`
`4.7
`4.6
`5.0
`
`ating solutions of compound I in pure nonaqueous solvents
`of differing polarity gave no reaction.
`Also included in Table II are the results of reactions run
`in water of varying pH. No significant change in ko„ was
`observed.
`When excess chloride was added, the reaction was
`slowed; /cobs showed a dependency on the reciprocal of the
`chloride concentration. This correlation is illustrated in Fig.
`2. However, when halogen ions other than chloride were
`added, the rate of reaction was faster than when an equimo-
`lar amount of chloride was added (Table II), and additional
`products were generated. Although these products were not
`identified, the chromatographic results are in agreement with
`the proposal that iodide and bromide compete with chloride
`and water for substitution, generating the corresponding
`5-bromo and 5-iodo compounds.
`
`Structural Effects
`
`Structural effects on the rate of the solvolysis reaction
`of compound I are given in Table I. The reactivity of com-
`pound IV indicates no involvement by the amide-linked tro-
`pane ring in the reaction, but elimination of the carbonyl,
`
`0 100 200 300 400
`Time (min)
`
`In Concentration (M)
`
`Fig. 1. Representative data showing the first-order disappearance of
`compound I over time.
`
`" The starting pH of 4.4 was adjusted to pH 3.0 with 0.1 N HCl and
`to pH 11 with 1.0 N NaOH.
`
`MYLAN Ex. 1013, Page 3
`
`(cid:9)
`

`
`Mosher and McBee
`
`cies are the aryl cation, the radical anion, radical cation, and
`the o- complex. Of these, radical anions are probably not
`involved as they are known to be quenched very efficiently
`by oxygen (10), and the reaction proceeded quite well in our
`studies even though we made no attempt to deoxygenate the
`solutions. Although not totally eliminated, radical cations
`are unlikely because of the enhanced reactivity in the pres-
`ence of potassium iodide. Iodide reduces cation radicals
`very well and, in fact, is sometimes used for the iodometric
`assay of cation-radical salts (11). If the intermediate is a
`radical cation, the presence of iodide should decrease the
`rate.
`A r complex intermediate has been proposed as a pos-
`sible intermediate in the photosubstitution of halophenols
`and haloanisoles (12), but the solvent system used was di-
`oxane/water (95:5). The greater nucleophilicity of dioxane
`relative to water would greatly enhance its ability to interact
`in an SN2-type reaction. The kinetics of SN2 reactions with
`one reactant in excess are usually apparent first order as is
`seen for the present reaction. However, when the kinetic
`expression for the dependence of the reaction on added chlo-
`ride is derived from Scheme II using a steady-state assump-
`tion for the concentrations of the intermediates, the ob-
`served rate constant is independent of chloride concentra-
`tion (see Appendix). The kinetics of the present reaction do
`
`ArCI + OH (cid:9)
`
`OH] --0-ArOH + Cl
`
`Cl (cid:9)
`
`[ArCI Cl]
`Scheme II
`
`show dependence on added chloride concentration. A linear
`relationship is seen between chloride concentration and
`1/kobs (Fig. 2). This kinetic disagreement in the presence of
`added chloride, the expectation that SN2 (AR*) processes
`occur predominantly in less polar media (13), and the inde-
`pendence of reaction rate on pH argues against an SN2
`(AR*)-type reaction.
`The above evidence leads us to a mechanism involving
`an aryl cation intermediate, SO (AR*). This mechanistic
`scheme, proposed by Kropp et al. (14-17), involves initial
`homolytic cleavage of the carbon halogen bond followed by
`electron transfer within the resulting radical pair to generate
`an ion pair (Scheme III). The aryl cation can then interact, in
`
`+ C1
`
`electron
`transfer
`
`Ar + Cl
`
`H2O
`ArOH + H
`
`Scheme III
`
`our solvolysis reaction, with water to give the hydroxy-
`substituted product, compound II, or with chloride ion to
`regenerate compound I. The applicability of the mechanism
`has been expanded to cover a variety of alkyl halide photo-
`reactions (18,19). Perhaps even more germane to the present
`work is the use of the mechanism to explain photosolvolysis
`reactions in water (20,21). As explained by Moore and Pagni
`
`800 -
`
`600 -
`
`400 -
`
`200
`
`1/k obs (min)
`
`1218 (cid:9)
`
`
`0 (cid:9)
`0.4
`0.3 (cid:9)
`0.2 (cid:9)
`0.1 (cid:9)
`0 0 (cid:9)
`Concentration of Added Chloride (M)
`Fig. 2. Plot of 1/kobs versus the molar concentration of added chlo-
`ride. Each point represents data from one reaction vessel.
`
`investigation since the observation in the mid-1950s that UV
`irradiation of nitrophenyl phosphates gave a substitution
`product, the corresponding nitrophenol (4). The seemingly
`nucleophilic photosubstitution reaction generated consider-
`able interest because the meta isomer showed the most ef-
`ficient and clean reaction. This is in contrast to ground-state
`orientation rules for thermal reactions. By the late 1960s it
`was recognized that nucleophilic aromatic photosubstitution
`is a fairly general reaction (5,6). It can occur with a variety
`of cyclic, polycyclic, and heterocyclic aromatic systems,
`various solvents, and a whole range of leaving groups and
`reacting nucleophiles. As investigations continue, there is an
`increasing number of proposed mechanisms for the reac-
`tions, most of them dependent on the reaction conditions.
`
`Mechanistic Studies
`
`In general, irradiation of aryl halides produces products
`of nucleophilic substitution [Eq. (1)], reductive dehalogena-
`tion [Eq. (2)], and inter- and intramolecular arylation (7).
`
`by
`ArX —> Ar Y +
`
`hv
`ArX ---> ArH + SX (cid:9)
`SH
`
`(1)
`
`(2)
`
`Four general classes of mechanisms have been identi-
`fied for photostimulated aromatic substitution reactions (8)
`[Eq. (1)]. The classification is based on the type of interme-
`diate formed and uses an abbreviated notation for unimolec-
`ular (1), nucleophilic (N), substitution (S), via radicals (R).
`When radical anions or cations are formed, the reaction is
`termed SR N1 or SR,N1, respectively. If a primary photo-
`dissociation into an aromatic cation occurs the mechanism is
`denoted SN1 (AR*) and the classification of SN2 (AR*) is
`used for reactions involving the formation of cr complexes.
`Reductive dehalogenation, Eq. (2), provides yet another
`mechanism to be considered and usually involves homolytic
`cleavage of the aryl—halogen bond. The aryl radical interme-
`diate thus formed can abstract a hydrogen atom from appro-
`priate hydrogen donor solvents to give the reduced product
`(9). The absence of a reduction product in the current reac-
`tion does not necessarily eliminate the possibility of an aryl
`radical as the reactive intermediate but does suggest that
`alternative pathways are more probable.
`The effect of solvent polarity on the rate of reaction is
`consistent with the appearance of a polar or charged inter-
`mediate in the rate-determining step. Possible charged spe-
`
`MYLAN Ex. 1013, Page 4
`
`(cid:9)
`

`
`Photoreactivity of LY277359 Maleate in Solution (cid:9)
`
`1219
`
`(21), the absence of the reduction product is due to the in-
`ability of the radical intermediates to abstract a hydrogen
`atom efficiently from water. Therefore the two radicals ei-
`ther recombine or undergo electron transfer to give the ions.
`The large dielectric constant of water will facilitate the elec-
`tron transfer and favor the formation of ionic species.
`The aryl cation formed is highly reactive and will react
`rather indiscriminately with surrounding nucleophiles (usu-
`ally water). The ability of added nucleophiles to react with
`the cation is dependent upon the stability of the cation and
`the concentration of the nucleophile. For the highly reactive
`and unstable phenyl cation, nucleophile concentrations must
`approach that of the solvent for reaction to occur (22). The
`reactivity of compound I shows a modest inverse depen-
`dence on added chloride as a competing nucleophile (see
`below), the dependency agreeing kinetically with Scheme
`III. This is not unexpected as the cation formed, relative to
`the phenyl cation, should be somewhat stabilized by the
`ether oxygen. On changing the pH from 3 to 11, no effect on
`reaction rate was observed, further supporting a reactive
`cationic intermediate. In spite of the nucleophilicity of -OH,
`its concentration at pH 11 cannot compete with the avail-
`ability of water. There are several examples in the literature
`of photohydrolyses that are independent of hydroxide ion
`concentration over a large pH range (23).
`
`Kinetic Description
`
`The proposed mechanism illustrated in Scheme III can
`be described by Scheme IV if it is assumed that the electron
`transfer occurs very fast and the concentrations of water and
`chloride are invariant. If a steady-state assumption is made
`
`k,
`ki (cid:9)
`ArC174--4:Ar ±-0--ArOH
`k2
`Scheme IV
`
`for [Art], then the loss of ArCI can be described by Eq. (3):
`
`kik3
`d[ArCI] (cid:9)
`
`[ArCI]
`-
`
`dt - k2 + k3
`
`(3)
`
`and the relationship between the rate constants from Scheme
`IV and /cobs is
`
`kobs k2k-Fik3k3
`
`(4)
`
`In comparing Scheme IV to Scheme III, it is seen that pho-
`tochemical initiation, containing terms for intensity and
`quantum efficiency, is incorporated into k1. A chloride con-
`centration term is included in k2 and can be described as k2
`k'2[C1-], and k3 incorporates a water concentration term,
`k3 = k'3[H20]. No ionic strength effect was observed over
`the range evaluated (0-0.2 Al).
`The various components of the overall rate constant
`were evaluated experimentally by altering the reaction con-
`ditions and examining the resulting observed rate constants.
`By monitoring the reaction in the presence of added chloride
`ion, alterations in k2 were imposed. Transforming Eq. (4) by
`
`substituting the chloride terms for k2 and rewriting in the
`reciprocal form gives Eq. (5):
`
`1 (cid:9)
`kobs (cid:9)
`
`k' 2[01 (cid:9)
`+
`kik3
`
`1
`k1
`
`(5)
`
`This equation shows a linear relationship between 1/kobs and
`[Cl-]. The inverse of the y intercept is equal to k1 , and this
`value along with the slope can be used to solve for the ratio
`of k3 to k'2. This plot, given in Fig. 2, shows good linearity
`over the chloride range studied (r = 0.93). Regression anal-
`ysis gives a value of 0.00407 for k1 and a relationship of k'2
`= 5.22 k3. Thus at chloride concentrations around 0.2 M, the
`k2 (k2 = k'2[C1- ]) and k3 terms are approximately equal and
`at chloride concentrations of 0.02 M or less, the k3 term
`dominates and Eq. (4) reduces to /cobs = k1.
`In the strictest sense, Eq. (3) is not valid at zero added
`chloride since the chloride concentration changes as the re-
`action progresses. However, as a worse-case scenario, if the
`reaction is followed to completion in the absence of added
`chloride, the maximum amount of chloride ion produced in
`our studies would be equal to the initial concentration of
`compound I (approximately 0.1 mM). This chloride concen-
`tration is easily within the range discussed above where kobs
`= k, and the observed rate is independent of the amount of
`chloride present or being produced.
`To evaluate the k3 component (k3 = k'3[1-120]), the wa-
`ter concentration was varied by using a mixed solvent sys-
`tem with decreasing amounts of water in acetonitrile. The
`results are given in Table II and indicate that kobs decreases
`as water concentration decreases. In retrospect, this study
`should not have given us any information since zero added
`chloride causes the k3 term to drop out and therefore our
`model is insensitive to this effect. In addition, since the aryl
`cation that is formed during the reaction is quite reactive, the
`nucleophilicity of acetonitrile may allow the formation of the
`nitrilium ion, as has been seen for anthracene (24). It seems
`reasonable that the decrease in reaction rate that we are
`seeing is due to the decreased polarity of the solvent which
`decreases the ability of the electron transfer to occur and the
`ionized species to form (17,18,25). Thus under these condi-
`tions, the assumptions made in deriving Scheme IV fail.
`Evaluation of the k1 term, containing intensity and
`quantum yield, using the ko„,, relationship [Eq. (4)], began
`with a determination of the rate constants for the reaction as
`light intensity was varied. These results are given in Fig. 3
`for a range of starting concentrations. As the initial concen-
`tration of compound I increases, kobs decreases in a nonlin-
`ear fashion. Varying the light intensity by changing the dis-
`tance the reaction vessel is placed from the light source gives
`similar concentration effects but causes the curve to shift to
`higher kobs values with increasing intensity. Attempts to lin-
`earize the data by plotting kobs against the reciprocal of the
`concentration (26) were unsuccessful, indicating changes in
`the quantum yield, 1, and/or the intensity with concentra-
`tion. Since the reaction was conducted at concentrations of
`compound I where not all of the incident light was absorbed,
`alterations in the intensity term are expected, and can be
`corrected by using Beer's law. Among the many other
`causes for variations in 1 and/or intensity is the potential for
`the product, II, to absorb light competitively (inner-filter
`
`MYLAN Ex. 1013, Page 5
`
`(cid:9)
`(cid:9)
`(cid:9)
`

`
`1220 (cid:9)
`
`Mosher and McBee
`
`factors of geometry, etc., which are unique to the apparatus
`being used. It should therefore be replaced with the more
`general term n as suggested by Moore (26). By making this
`substitution and one for the Beer's law expression for ab-
`sorbance into the equation, a working form is generated.
`
`kobs
`
`kn(1 (cid:9)
`
`e-2.303 b Co)
`
`Co
`
`(9)
`
`The data from Fig. 3 were fit to Eq. (9) with the non-
`linear regression function provided in the SAS software,
`while floating the quantities ((rn.) and (e b). Initial estimates
`were obtained from a Taylor's series expansion of Eq. (9)
`which gave a final result linear in concentration describing
`the limiting slope at low Co values. The results of the SAS
`regression are given in Table III and are used to generate the
`solid lines in Fig. 3.
`The good fit of Eq. (9) to our experimental data indi-
`cates that the changes in /cobs with concentration can be ex-
`plained in terms of an inner-filter effect. This is further sub-
`stantiated in Table IV, where the addition of compound II
`and other organic species, compounds XI and XII, known to
`absorb light in the same wavelength range, are seen to slow
`the reaction.
`
`Structure-Reactivity Relationships
`
`Spectroscopic analysis of the compounds in Table I pro-
`vides some rationale for the observed reactivity. All of the
`compounds show a 7r-ir* band peaking in the wavelength
`range of 280-325 nm. Compounds V and VI, having func-
`tional groups either meta or para to the chloro substituent,
`show an absorption band peaking around 280 nm. This wave-
`length is below our reaction vessel cutoff and these com-
`pounds would not be expected to show reactivity. The com-
`pounds which have a carbonyl and an ether or hydroxyl
`oxygen in an ortho configuration show a red shift in this peak
`(VII, 308 nm; VIII, 300 nm), presumably due to hydrogen
`bonding. These absorption bands are barely in the spectrum
`transmitted by the vessel and reactivity would still be ques-
`tionable. The fact that VIII (lower kmax) has some reactivity
`and VII does not implies that factors other than vessel cutoff
`are involved. One possibility is the need for the alkoxy-
`group activation of the para position (31,32) not just the
`ortho/para direction (33) as has been postulated for hydroxy
`groups. The lack of reactivity with X (Xmax = 301 nm),
`where the chloro is in the 4 position, further suggests that
`para activation is of considerable importance.
`Constraining the ether oxygen into a ring further shifts
`the knax as illustrated by compounds I (320 nm) and IV (322)
`(protonated carbonyl) and gives full reactivity. However,
`allowing a charge on the carbonyl or removing it (IX),
`
`Table III. Parameters Generated from SAS Regression Analysis
`
`Distance from
`light source (cm)
`
`5
`10
`15
`
`orn X 106 (cid:9)
`
`Eb x 10-4
`
`2.55
`0.509
`0.362
`
`1.13
`1.18
`0.89
`
`0.07
`
`0.06
`
`0.05
`
`0.04
`
`0.03
`
`0.02
`
`0.01
`
`k obs (1/min)
`
`0.00
`0.00 (cid:9)
`
`0.05 (cid:9)
`
`0.10 (cid:9)
`
`0.15 (cid:9)
`
`0.20
`
`0.25
`
`Initial Concentration (mM)
`Fig. 3. Plot of lc„,„ versus the initial molar concentration of I for
`reaction vessels at different distances from the light source. Each
`point represents results from one reaction vessel. The overlaid lines
`are a fit of Eq. (9) to the data. (ID) 5 cm; (0) 10 cm; and (A) 15 cm.
`
`effect) (27). Since a UV scan of compound II shows it to
`have an absorbance spectrum almost identical to that of
`compound I, competitive light absorption by the product
`was highly probable.
`In a light irradiation experiment, the inner filter effect
`can be described as the absorption of incident radiation by a
`species other than the intended primary absorber (28). In this
`experiment, the secondary absorber proves to be the prod-
`uct, II, of the reaction. Zimmerman et al. (29) were the first
`to consider the case where the reactant and product have
`similar spectral characteristics. The concept was later ex-
`panded by Bunce (30) to cover the general case where the
`product and reactant are different spectrally. For our studies
`we use a form intended for irradiation studies at an isosbestic
`point (30).
`
`CO (cid:9)
`In Cf (cid:9)
`
`(mo(1 (cid:9)
`
`e-2.303 Abs) (cid:9)
`
`Co
`
`(6)
`
`In Eq. (6), Co and Cf are initial and final concentrations of
`reactant respectively, (Dr is the quantum yield for reactant
`disappearance, and /0 is the accumulated photon dose. The
`term (1 - e-2.303 Abs) is a correction factor for the amount
`of light actually absorbed. Differentiation with respect to
`time gives Eq. (7), where I is now an intensity term, i.e.,
`photon dose per unit time.
`dC (cid:9) OA) _ e-2.303 Abs)
`-
`Co
`dt
`
`(7)
`
`Although this equation describes reactions run using mono-
`chromatic light, we assume for demonstrative purposes that
`it will hold for the summation of wavelengths initiating our
`reaction. This is a reasonable assumption since the reactant
`and product have almost identical UV absorption spectra.
`Comparing Eq. (7) to Eqs. (3) and (4) gives Eq. (8),
`where the observed rate constant is seen to be a function of
`the initial concentration of compound I used and the absor-
`bance of the solution.
`
`kobs
`
`(Dri(1 (cid:9)
`
`e-2.303 Abs)
`
`Co
`
`(8)
`
`It should be noted at this time that the I term contains
`
`MYLAN Ex. 1013, Page 6
`
`(cid:9)
`(cid:9)
`(cid:9)
`(cid:9)
`(cid:9)
`

`
`Photoreactivity of LY277359 Maleate in Solution (cid:9)
`
`1221
`
`Table IV. Effect of Adding UV Absorbing Species on the Observed Rate of the Substitution Reaction
`
`Species
`added
`
`II
`II
`II
`Control
`XI
`XII
`Control
`
`[Compound I]
`(mM)
`
`[Species]
`(mM)
`
`Distance from
`light source (cm)
`
`lcol„
`(min' x 103)
`
`0.11
`
`If
`
`0.11
`0.017
`
`0.11
`0.0733
`0.0367
`
`1.12
`1.06
`
`10
`
`10
`22
`
`2.84
`2.78
`4.12
`4.68
`2.20
`0.351
`5.30
`
`k, (cid:9)
`k3
`ArCI + OH - 7.1:1"--[ArCI • • OH - ] --0-ArOH + Cl -
`k2
`
`+ CI -.7.7"--[ArCI •• Cl -
`4
`
`Scheme II*
`
`d[ArCI]
` = -ki[01-1-][ArCl] + k2[ArC1 • • OH -]
`dt
`
`- k5[C1-][ArCl] + k6[ArC1 •• Cl -] (cid:9)
`
`(10)
`
`likewise,
`
`d[ArCI • • CI
`dt (cid:9)
`
`= k5[C1-][ArCl] - k6[ArCI • • Cl-]
`
`even with the oxygen tied into the furan ring, reduces or (cid:9)
`eliminates reactivity in our system as the Xmax is shifted back (cid:9)
`toward the vessel cutoff. (cid:9)
`In conclusion, the photoinduced solvolysis reaction of
`compound I probably involves homolytic cleavage of the (cid:9)
`aryl-chloride bond followed by electron transfer to give an (cid:9)
`aryl cation intermediate. This highly reactive species then (cid:9)
`reacts almost indiscriminately with surrounding nucleophiles
`to give the corresponding substitution product; in water this
`is the hydroxylated species. Greatest reactivity is seen when ment is as follows. The differential equation describing the
`the parent molecule contains a substituent oriented para to loss of ArCI with time is
`the chloro group which is para activating in the photoexcited
`state and, of course, an absorption band appearing at wave- (cid:9)
`lengths above the vessel cutoff. For the series of compounds
`investigated, this occurs with an ether-linked oxygen para to (cid:9)
`the chloro group (preferably constrained into a furan ring)
`and an uncharged carbonyl attached ortho to the ether. (cid:9)
`The reaction is slowed with the addition of sodium chlo- (cid:9)
`ride attributed to a common ion effect (enhanced retroreac- (cid:9)
`tion) (34,35). In the absence of added chloride, the reaction
`can be described kinetically by an initiation term. An inner-
`filter effect is seen with this reaction where the appearance and
`of II slows the reaction by absorbing a portion of the incident
`radiation. The addition of II, or other species known to ab- (cid:9)
`sorb light in the wavelength range of 370-280 nm, slows the (cid:9)
`reaction via the same mechanism. Unfortunately, the use of
`(12)
`this technique to formulate stable parenteral formulations of
`I is complicated by the requirements that excipients need to Invoking the steady-state assumption for [ArC1 • • CI -1 and
`be nontoxic and pharmacologically inactive and to c

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