ArticlePDF Available

High Solubility Piperazine Salts of the Nonsteroidal Anti-Inflammatory Drug (NSAID) Meclofenamic Acid

Authors:

Abstract and Figures

Meclofenamic acid (MFA) is the most potent antiinflammatory drug among the fenamic acids. We report (1) two cocrystals of MFA with isonicotinamide (INA) and 4,4′-bipyridine (BPY); (2) polymorphs of MFA and piperazine (PPZ) 1:1 salt (orthorhombic P212121 O and monoclinic P21/c M), MFA−PPZ− H2O 1:1:1 salt hydrate, MFA−PPZ 2:1 salt; and (3) MFA and 2- aminopyridine (2-APY) 1:1 salt, MFA and 4-aminopyridine (4-APY) 1:1:1 salt hydrate. Sublimation of MFA gave single crystals for X-ray diffraction which provided good quality data for refinement and all atomic coordinates. The cocrystal and salt structures are assembled via neutral O−H···O, O−H···N, N−H···O, N−H···N, and ionic O− H···O−, N+−H···O− hydrogen bonds. The disorder of the methyl group in the MFA crystal structure is absent in the cocrystal and salt structures, which contain different conformers (A or B) of methyl group orientation. The solubility of MFA−INA (1:1) and MFA−BPY (1:0.5) cocrystals is 2.9 and 7.6 times higher than that of MFA at 37 °C in 50% EtOH−water. Interestingly, MFA−PPZ-M 1:1 salt and its 1:1:1 hydrate are 2724- and 1334-fold more soluble than MFA. Both of these salts transformed in 50% EtOH−water slurry at 37 °C to MFA−PPZ 2:1 salt after 24 h, which in turn transformed to MFA after another 24 h of slurry stirring. Remarkably, the dissolution rate of MFA−PPZ-M (1:1) salt in water is just slightly lower than that of the marketed sodium meclofenamate.
Content may be subject to copyright.
High Solubility Piperazine Salts of the Nonsteroidal
Anti-Inflammatory Drug (NSAID) Meclofenamic Acid
Palash Sanphui, Geetha Bolla, and Ashwini Nangia*
School of Chemistry, University of Hyderabad, Prof. C. R. Rao Road, Gachibowli, Central University P.O., Hyderabad 500046, India
*
SSupporting Information
ABSTRACT: Meclofenamic acid (MFA) is the most potent anti-
inflammatory drug among the fenamic acids. We report (1) two
cocrystals of MFA with isonicotinamide (INA) and 4,4-bipyridine
(BPY); (2) polymorphs of MFA and piperazine (PPZ) 1:1 salt
(orthorhombic P212121O and monoclinic P21/cM), MFAPPZ
H2O 1:1:1 salt hydrate, MFAPPZ 2:1 salt; and (3) MFA and 2-
aminopyridine (2-APY) 1:1 salt, MFA and 4-aminopyridine (4-APY)
1:1:1 salt hydrate. Sublimation of MFA gave single crystals for X-ray
diffraction which provided good quality data for refinement and all
atomic coordinates. The cocrystal and salt structures are assembled via
neutral OH···O, OH···N, NH···O, NH···N, and ionic O
H···O,N
+H···Ohydrogen bonds. The disorder of the methyl group in the MFA crystal structure is absent in the cocrystal
and salt structures, which contain different conformers (A or B) of methyl group orientation. The solubility of MFAINA (1:1)
and MFABPY (1:0.5) cocrystals is 2.9 and 7.6 times higher than that of MFA at 37 °C in 50% EtOHwater. Interestingly,
MFAPPZ-M 1:1 salt and its 1:1:1 hydrate are 2724- and 1334-fold more soluble than MFA. Both of these salts transformed in
50% EtOHwater slurry at 37 °C to MFAPPZ 2:1 salt after 24 h, which in turn transformed to MFA after another 24 h
of slurry stirring. Remarkably, the dissolution rate of MFAPPZ-M (1:1) salt in water is just slightly lower than that of the
marketed sodium meclofenamate.
INTRODUCTION
About 40% of new molecular entities coming out of the drug
discovery pipeline will never advance in the development chain
because of biopharmaceutical issues, such as poor aqueous
solubility, low dissolution rate, low permeability, and first-pass
metabolism in the liver. An enhancement of drug solubility for
therapeutic agents can improve their bioavailability. Identifying
the optimum solid form of an active pharmaceutical ingredient
(API) is always desirable for clinical use.
1
Over 80% of all drugs
are marketed as tablets, and oral administration is the most
preferred drug delivery route. Different solid-state forms, such
as polymorphs, amorphous, cocrystals, salts, and their hydrates,
can be crystallized to improve physicochemical properties of
drug substances. Crystal engineering is particularly well suited
to cocrystallization of drugs with safe coformers.
2
The advan-
tage of cocrystals and salts for improving physical properties
such as solubility, bioavailability, stability, etc.
3
is that the struc-
ture of the drug molecule is unchanged but the modification
is at the supramolecular level (intermolecular interactions,
hydrogen bonding, molecular packing). A practical advantage is
that the extent of solubility enhancement for cocrystals (420
times) is an order of magnitude higher than that for poly-
morphs (23 times). Pharmaceutical salts of course are the
most preferred formulation for solubility enhancement (100
1000 times).
4
On the downside, salts are more likely to form
hydrates (often a drawback in terms of stability) compared to
cocrystals. The ΔpKarule, wherein ΔpKa=pKa(conjugate acid
of base) pKa(acid), is a useful guide to know beforehand if
an acidbase complex will give a neutral cocrystal (ΔpKa<3)
or an ionic salt (ΔpKa> 3). A more practical cutoff for organic
salts is ΔpKa< 0 for cocrystal, ΔpKa> 3 for salts, and the range
0<ΔpKa< 3 being an unpredictable zone wherein multiple
proton states could be observed.
5
Meclofenamic acid (MFA)
6
(Figure 1) is a nonsteroidal anti-
inflammatory, antipyretic, analgesic drug used in the treatment
of postoperative and traumatic inflammation and swelling
by the inhibition of prostaglandin biosynthesis pathway.
6a
It is
a selective cyclo-oxygenase-2 (COX-2) inhibitor.
6b
Similar to
Received: January 2, 2012
Revised: February 17, 2012
Published: February 24, 2012
Figure 1. Two conformers (A and B) of meclofenamic acid (MFA)
present in crystal structures.
Article
pubs.acs.org/crystal
© 2012 American Chemical Society 2023 dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 20232036
diclofenac, MFA is also used as a KCNQ2/Q3 potassium
channel opener, depressor of cortical neuron activity, and
exhibits anticonvulsant activity.
6e
MFA also has a transient
effect on platelet aggregation, but unlike aspirin, it does not
cause bleeding.
6h
It is a BCS class II drug of low solubility
(30 mg/L) and high permeability (log Pow = 5).
7
Novel gel and
cream formulations of MFA provide maximal topical activity to
the drug.
6e
Exposure of dilute solutions of MFA to visible
or UV light resulted in fairly rapid decomposition.
6f
Because of
its low aqueous solubility
7
and thus bioavailability, the sodium
salt of MFA (solubility > 250 g/L)
7b
is sold commercially as
100 mg capsules under brand names Meclomen, Melvon,
Movens, and Arquel. Fábián et al.
8
recently published cocrystals
of flufenamic acid, niflumic acid, tolfenamic acid, and
mefenamic acid with nicotinamide, but they mentioned some
difficulty with MFA. Our results on cocrystals and salts of MFA
demonstrate that organic salts of APIs can exhibit comparable
dissolution rates to traditional metal ion salts. The novel
cocrystals and salts of MFA with isonicotinamide (INA), 4,4-
bipyridine, piperazine, aminopyridine, etc. were prepared by
solvent-assisted grinding, and their solubility was measured in a
USP dissolution tester. Crystal structures, phase transforma-
tions, and solubility of MFA cocrystals and salts are discussed.
RESULTS AND DISCUSSION
The sodium salt of MFA was obtained from Sigma-Aldrich,
India, and converted to the free acid with aq. HCl. The Na salt
of MFA is readily soluble in water. Vijayan et al.
9
(1981)
published the first crystal structure of MFA but it had a high R-
factor of 0.135. The same authors reported crystal structure of
the 1:1 complex of MFA with choline hydrate and ethanol-
amine.
10
The disorder in the methyl group orientation of MFA
(Figure 1) persisted at low temperature. We have now obtained
good quality single crystals of MFA by sublimation and col-
lected data at 298 and 100 K. There are no reports on crystal
structures of MFA and its salt/cocrystal in the interim period.
Crystallographic parameters for all crystal structures are listed
in Table 1. ORTEP diagrams are displayed in Figure S1,
Supporting Information.
CRYSTAL STRUCTURE DESCRIPTION
Meclofenamic Acid (MFA). MFA consists of two aryl
moieties, the N-2,6-dichloro-3-methylphenyl group and the N-
2-benzoic acid group. The aryl rings are twisted almost in a
perpendicular orientation of an 82°dihedral angle between the
two ring planes. This twist relieves steric congestion of ortho-
substituted phenyls at the secondary amine. MFA was crys-
tallized by slow sublimation at 190200 °C over 45hto
afford diffraction quality single crystals which was solved in the
triclinic space group P1̅with one molecule in the asymmetric
unit. Similar to other fenamic acids, for example, mefenamic
acid and tolfenamic acid,
11
two MFA molecules form a centro-
symmetric carboxylic acid dimer of OH···O hydrogen bond
(O···O, 2.632(4) Å) in the R22(8) ring motif.
12
An intra-
molecular NH···O hydrogen bond (N···O, 2.679(4) Å) of
R11(6) ring motif (Figure 2) rigidifies the molecule. Hydrogen
bond parameters in crystal structures are listed in Table 2.
The 3-methyl group is disordered over two positions with un-
equal site occupancy factor (sof conformer B= 0.58(1) and
A= 0.42(1); see Figure 1) at 298 K (RT structure). X-ray reflec-
tions for MFA crystal were collected at 100 K (LT structure)
in an attempt to resolve the disorder issue (R-factor 0.110).
The disorder of the Me group persisted, now with sof of
conformer B= 0.577(15) and A= 0.423(15). The elongated
thermal ellipsoid of Cl2 (chlorine atom 2) perpendicular to the
aromatic plane in the RT structure could be modeled as Cl2A
and Cl2B (chlorine 2A and 2B) with sof of 0.55(3) and 0.45(3)
in the LT structure. The RT structure of MFA is described in
this paper because it has better refinement parameters and
lower R-factor (0.083).
Meclofenamic AcidIsonicotinamide Cocrystal (1:1,
MFAINA). MFAINA (1:1) cocrystal was prepared by solid-
state grinding, and the resulting solid was crystallized from aceto-
nitrile to afford the single crystal X-ray structure in monoclinic
space group P21/c, which contained conformer A of MFA. The
carboxylic acid dimer in the reference drug crystal structure is
replaced by carboxylic acidpyridine heterosynthon (O···N,
2.657(4) Å, OH···N, 174°) as the main bimolecular R22(7)
ring motif (Figure 3a). Different synthons present in the novel
solid phases are displayed in Scheme 1. INA molecules aggre-
gate via the carboxamide dimer (N···O, 2.926(4) Å, NH···O,
169°) at the secondary level. Such 4-molecule supra-
molecular units are connected via CCl···Ointeraction(3.2)
in a ladder motif along the b-axis (Figure 3b).
Meclofenamic Acid4,4-Bipyiridine Cocrystal (1:0.5,
MFABPY). MFABPY (1:0.5) cocrystal (crystallized from
acetonitrile) was solved in monoclinic space group P21/cand
contains MFA conformer A. The acidpyridine heterosynthon
(O···N, 2.668(4) Å, OH···N, 178°;R
22(7) ring motif) is
supported by an auxiliary CH···O hydrogen bond (C···O,
3.194(6) Å) between the pyridine ortho CH to the carbonyl
of carboxylic acid (Figure 4a). The trimolecular units in the
cocrystal make a stepladder motif along the c-axis (Figure 4b).
MFABPY (1:1) is a cocrystal based on CO bond distances
of 1.215(8) Å and 1.316(7) Å in the carboxylic acid group and
CNC bond angle of 115.7(5)°in the pyridine ring.
Piperazinium Meclofenamate Salt Polymorphs (1:1,
MFAPPZ-M and MFAPPZ-O). Piperazinium meclofena-
mate salt (1:1) crystallized as monoclinic (P21/c) and ortho-
rhombic (P212121) polymorphs concomitantly from acetonitrile
solvent. There is one molecule each of meclofenamate anion
(conformer B) and piperazinium cation in the asymmetric unit.
Compared to the cocrystal, proton transfer occurred in the salt
structure from carboxylic acid to the NH base of piperazine.
In the monoclinic structure (MFAPPZ-M), each piperazine
molecule forms four hydrogen bonds with two meclofenamates
and two piperazinium cations through ionic N+H···O
(N···O, 2.677(4) Å) and neutral NH···N(N···N, 2.851(4) Å)
and NH···O(N···O, 2.949(4) Å) hydrogen bonds (Figure 5a).
The carboxylate CO bond distances are 1.236(4) Å and
1.271(4) Å and piperazinium CNC is 112.0(3)°suggest-
ing an ionized species. The CO bond distances difference
(ΔCO) is less than 0.1 Å compared to the neutral species.
The packing of meclofenamate anions (conformer A) and
piperazinium cations is similar in the orthorhombic poly-
morph (MFAPPZ-O). Each piperazine molecule forms
four hydrogen bonds with two meclofenamate anions and
two piperazinium cations through N+H···O(2.698(3) Å),
NH···N (2.858 (3) Å), and NH···O (2.960 (3) Å) hy-
drogen bonds (Figure 5b). CO bond distances (1.226(6) Å,
1.263(5) Å) and CNC angle (111.8(3)°) are consistent
with a salt species.
5
In both polymorphs, six piperazine cations
are sandwiched between three meclofenamate anions along the
c-axis. A minor difference between the two polymorphs is
the orientation of the methyl group in MFA, that is, conformer
Crystal Growth & Design Article
dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 202320362024
A or B. Piperazine is a small cyclic diamine that is pharma-
ceutically acceptable and has anthelmintic activity.
13
Piperazine
(pKa= 9.72) can act both as a neutral and a cationic co-
former to give a cocrystal or salt product with an acidic API.
The Cambridge Structural Database (CSD ver. 5.32, 2010,
November 2011 update)
14
contains about 20 organic cocrystals
and 25 salts or salt hydrates of piperazine (Table S1, Supporting
Information). Surprisingly, there is only one piperazinium
Table 1. Crystallographic Parameters of MFA and Its Cocrystals and Salts
MFA-100K MFA-RT MFAINA (1:1) MFABPY (1:0.5) MFAPPZ-M (1:1)
empirical formula C14H10Cl2NO2C14H10Cl2NO2C14H11Cl2NO2·C6H6N2OC
14H11Cl2·0.5(C10H8N2)C
14H10Cl2NO2·C4H11N2
formula weight 295.13 295.13 418.27 374.23 382.28
crystal system triclinic triclinic monoclinic monoclinic monoclinic
space group P1̅P1̅P21/cP21/cP21/c
T(K) 100 298 298 298 298
a(Å) 8.5209(11) 8.566(3) 7.5327(13) 7.4180(6) 16.3805(13)
b(Å) 8.8775(11) 8.968(3) 31.313(5) 8.2144(6) 8.3717(5)
c(Å) 9.1986(12) 9.378(3) 9.601(3) 28.770(2) 14.4609(11)
α(°) 104.265(2) 103.090(6) 90 90 90
β(°) 103.337(2) 103.194(6) 121.439(16) 97.099(7) 102.347(8)
γ(°) 91.569(2) 92.538(6) 90 90 90
V3) 653.56(14) 679.7(4) 1932.2(7) 1739.6(2) 1934.8(3)
Dcalcd (g cm3) 1.500 1.442 1.438 1.429 1.312
μ(mm1) 0.492 0.473 0.363 0.388 0.351
θrange 2.3625.89 2.4725.77 2.5929.03 2.7529.10 2.7426.31
Z224 4 4
range h9to+9 9to+9 9to+9 8to+8 20 to +20
range k10 to +10 10 to +10 39 to +39 9to+9 10 to +10
range l10 to +10 10 to +10 11 to +12 31 to +31 18 to +18
reflections collected 5863 6239 9616 8621 8452
total reflections 2153 2227 3944 2487 3956
observed reflections 1954 1698 1932 1726 1574
R1[I>2σ(I)] 0.1106 0.0833 0.0567 0.0892 0.0641
wR2(all) 0.2193 0.1910 0.1547 0.2088 0.1850
goodness-of-fit 1.209 1.122 0.915 1.127 0.933
diffractometer SMART BRUKER SMART BRUKER Oxford CCD Oxford CCD Oxford CCD
MFAPPZ- O (1:1) MFAPPZ (2:1) MFAPPZH2O (1:1:1) MFA2-APY (1:1) MFA4-APYH2O (1:1:1)
empirical formula C14H10Cl2NO2·C4H11N22(C14H10Cl2NO2)·C4H12N2C14H10Cl2NO2·C4H11N2·H2OC
14H10Cl2NO2·C5H7N2C14H10Cl2NO2·C5H7N2·H2O
formula weight 382.28 678.42 400.29 390.26 408.27
crystal system orthothombic triclinic monoclinic monoclinic monoclinic
space group P212121P1̅P21/cP21/nP21/c
T(K) 298 298 298 298 298
a(Å) 7.929(3) 7.9109(3) 16.137(4) 14.675(4) 16.227(4)
b(Å) 8.168(3) 11.1289(4) 8.4889(15) 7.1318(15) 8.4770(15)
c(Å) 28.381(15) 18.1180(7) 15.596(3) 18.605(6) 15.686(4)
α(°) 90 81.943(3) 90 90 90
β(°) 90 80.424(4) 114.94(3) 110.73(3) 116.18(3)
γ(°) 90 88.809(3) 90 90 90
V3) 1838.1(13) 1557.36(11) 1937.1(7) 1821.2(8) 1936.2(7)
Dcalcd (g cm3) 1.381 1.447 1.373 1.423 1.401
μ(mm1) 0.370 0.425 0.358 0.375 0.360
θrange 2.8626.31 2.7426.31 2.7724.71 2.7729.22 2.7826.37
Z42 4 4 4
range h9to+9 9to+9 18 to +18 18 to +18 20 to +20
range k10 to +10 13 to +13 9to+9 8to+8 10 to +10
range l35 to +35 22 to +22 18 to +18 23 to +23 19 to +19
reflections
collected 5203 12514 6934 7324 8218
total reflections 3366 6346 3283 3721 3958
observed
reflections 1540 3882 1942 2081 2641
R1[I>2σ(I)] 0.0597 0.0687 0.0868 0.0459 0.0609
wR2(all) 0.0663 0.1770 0.2540 0. 1228 0.1608
goodness-of-fit 0.865 1.030 1.046 0.919 1.038
diffractometer Oxford CCD Oxford CCD Oxford CCD Oxford CCD Oxford CCD
Crystal Growth & Design Article
dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 202320362025
monocation (Refcode CUKVOU) listed with a carboxylate
counterion up to the recent update of the CSD; there are
73 refcodes for piperazinium dication (Table S2, Supporting
Information). We report three piperazinium monocations with
meclofenamate anion in MFAPPZ-M, MFAPPZ-O, and
MFAPPZH2O crystal structures. Furthermore, there are no
examples of piperazine salts which are polymorphic. MFAPPZ
(1:1) is the first example of piperazine salt polymorphs with 3D
coordinates determined. Crystal density (1.312, 1.381 g cm3),
packing fraction (64.2, 68.0%, calculated in Platon), and lattice
energy (176.93, 186.39 kcal mol1, calculated using Com-
pass force field in Cerius2)
15
of monoclinic and orthorhombic
polymorphs suggest that the orthorhombic form should be
Table 2. Hydrogen Bonds in Crystal Structures (Neutron-Normalized Distances)
crystal forms interaction H···A (Å) D···A (Å) DH···A(°) symmetry code
MFA N1H1··· O1 1.95 2.679 (4) 126 intramolecular
O2H2··· O1 1.66 2.632 (4) 171 1 x,2y,z
MFAINA N1H1··· O1 1.83 2.653(4) 137 intramolecular
O2H2··· N3 1.68 2.657(4) 174 x,y,1+z
N2H2A··· O3 1.93 2.926(4) 169 1 x,y,1z
MFABPY N1H1··· O1 1.84 2.640(4) 133 intramolecular
O2H2··· N2 1.69 2.668(4) 178 x,1+y,z
C15H15···O1 2.41 3.194(4) 129 x,1+y,z
MFAPPZ-M N1H1··· O1 1.75 2.604(4) 140 intramolecular
N2H2··· O1 1.68 2.677(4) 171 x, 1/2 y,1/2 + z
N2H2··· O2 2.53 3.231(4) 126 x,1/2 y,1/2 + z
N2H2A··· N3 1.89 2.851(4) 159 x,1/2 + y,1/2 z
N3H3A··· O2 1.99 2.949(4) 157 x, 3/2 y, 1/2 + z
MFAPPZ-O N1H1··· O1 1.79 2.619(3) 137 intramolecular
N2H2··· O1 1.71 2.698(3) 164 1 x, 1/2 + y, 1/2 z
N2H2··· O2 2.48 3.246(3) 133 1 x, 1/2 + y, 1/2 z
N2H2A··· N3 1.85 2.858(3) 178 1/2 + x, 1/2 y,z
N3H3A··· O2 1.97 2.960(3) 165 x, 1/2 + y, 1/2 z
C6H6··· Cl1 2.71 3.201(4) 107 intramolecular
MFAPPZ (2:1) N1H1··· O1 2.05 2.682 (3) 119 intramolecular
N2H2··· O3 1.99 2.642 (3) 120 intramolecular
N3H3A··· O1 1.70 2.690(3) 164 1 x,1y,1z
N3H3B··· O4 1.74 2.731(3) 167 1 x,y,1z
N4H4A··· O2 1.77 2.750(3) 162 1+x,y,z
N4H4B··· O3 1.72 2.695(3) 162 1 x,1y,1z
C29H29B···O3 2.49 3.318(3) 132 1 x,1y,1z
C32H32B···Cl3 2.67 3.446(3) 128 1 x,1y,1z
MFAPPZH2ON1H1···O1 1.91 2.699(3) 133 intramolecular
N3H1A···N2 1.99 2.939(4) 155 1 x, 1/2 + y, 1/2 z
N3H1B···O1 1.90 2.842(4) 154 x, 1/2 y,1/2 + z
N2H2···O3 2.10 3.029(4) 152 1 x,1/2+ y, 1/2 z
O3H3A···O1 1.87 2.796(4) 155 x,y,z
O3H3B···O2 1.72 2.692(4) 170 1 x,y,1z
C16H16B···O2 2.43 3.488(4) 167 1 x,y,1z
C17H17B···O3 2.35 3.356(4) 155 1 x, 1/2 + y, 1/2 z
C3H3···Cl1 2.67 3.607(4) 145 x,1/2 y, 1/2 + z
MFA2-APY N1H1···O1 1.88 2.631(3) 128 intramolecular
N2H2···O1 1.64 2.642(3) 175 1 x,y,z
N3H3A···O2 1.84 2.841(3) 173 1 x,y,z
N3H3B···O2 1.92 2.897(3) 164 1/2 + x, 1/2 y, 1/2 + z
MFA4-APYH2ON1H1···O2 1.79 2.603(4) 135 intramolecular
N2H2···O3 1.92 2.837(4) 149 x, 1/2 y, 1/2 + z
N3H3A···O1 2.07 3.017(4) 156 1 x, 1/2 + y, 1/2 z
N3H3B···O3 2.00 2.993(4) 169 x, 3/2 y, 1/2 + z
O3H3C···O1 1.78 2.761(4) 173 x, 1/2 y,1/2 + z
O3H3D···O2 1.75 2.713(4) 166 1 x,1/2 + y, 1/2 z
C17H17···O1 2.48 3.301(4) 131 1 x,1y,1z
Figure 2. Centrosymmetric carboxylic acid dimer in MFA. The methyl
group is disordered in the crystal structure of MFA with sof 0.58(1)
and 0.42(1).
Crystal Growth & Design Article
dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 202320362026
more stable. However, differential scanning calorimetry (DSC)
measurements (discussed later) indicate that the melting point
of the monoclinic form is higher and we show that MFAPPZ-M
is the stable polymorph.
Piperazinium Meclofenamate Salt (2:1, MFAPPZ).
Piperazinium meclofenamate salt (2:1) was obtained when the
components were ground with a few drops of EtOH added.
The crystal structure of MFAPPZ (2:1) was solved in P1̅space
group with two crystallographic meclofenamate anions (conformer
A and B) and two half piperazinium cations in the asymmetric
unit. The two conformers of MFA are arranged in an alternate
fashion ABAB parallel to the b-axis, and piperazinium cations are
sandwiched between the anion layers. Each piperazinium dication
forms four N+H···Oionic hydrogen bonds (N3+H3A···
O1, 2.690(3) Å; N3+H3B···O4, 2.731(3) Å; N4+H4A···
O2, 2.750(3) Å and N4+H4B···O3, 2.695(3) Å with
meclofenamate anions A and B (shown in thick bond and ball
and stick models, Figure 6). This is the only crystal structure of
MFA salts with both conformers A and B in the same lattice.
Piperazinium Meclofenamate Salt Hydrate (1:1:1,
MFAPPZH2O). Piperazinium meclofenamate salt hydrate
(1:1:1) was obtained during grinding of MFA with piperazine
hydrate or MFA and piperazine in the presence of water.
It crystallized in the monoclinic space group P21/cwith conformer B.
Two meclofenamate anions and two water molecules form ring
motifofgraphsetnotationR
44(12). Water molecules
are present as spacers between two meclofenamates. Two
piperazinium cations form N+H···O hydrogen bonds (N2+
H2···O3, 3.029(4) Å) with water (Figure 7a). The NH···N/O
hydrogen bond network (N3+H1B···O1,2.842(4,N3
+
H1A···N2, 2.939(4) Å) is shown in Figure 7b. Water molecules
reside in channels along the c-axis (Figure 7c). Thermogravimetric
analysis (TGA) (Figure S2, Supporting Information) confirmed
Scheme 1. Different Hydrogen Bond Synthons in Crystal Structures
a
a
(1) Acidacid dimer R22(8), (2) acidpyridine R22(7) ring motif, (3) carboxylateaminopyridinium R22(8) ring motif, (4) carboxylate
piperaziniumcarboxylate salt, (5) carboxylatepiperazinium salt, (6) carboxylatewater tetramer R44(12) ring motif.
Figure 3. (a) Acidpyridine heterosynthon in MFAINA (1:1)
cocrystal. (b) Acidpyridine heterosynthon, amide dimer homosyn-
thon, and CCl···O interaction along the b-axis in crystal structure.
Figure 4. (a) Acidpyridine heterosynthon form termolecular unit in
MFABPY (1:0.5) cocrystal structure. (b) Ladder structure along the
c-axis.
Figure 5. Crystal packing in (a) monoclinic form of MFAPPZ-M
(1:1) viewed along the b-axis, and (b) orthorhombic polymorph
MFAPPZ-O (1:1) along the a-axis. The main difference is in the
orientation of Me group in MFA, that is, conformer B (in M) and A
(in O).
Crystal Growth & Design Article
dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 202320362027
the stoichiometry as a monohydrate (calc. 4.68%, obsd. 4.71%).
A mefenamic acid piperazine salt hydrate (2:1:4) was recently
reported,
16
but its crystal structure is completely different
from our results because of hydrogen bonding to four water
molecules.
2-Aminopyridinium Meclofenamate Salt (1:1, MFA2-
APY). The organic cation and anion form R22(8) ring motif
(Scheme 1) between amino-pyridinium and carboxylate via
NH···O bonds (N···O, 2.642(3) Å, 2.841(3) Å). Such dimer
units are connected via amino NH···O(N···O 2.897(3) Å)
hydrogen bond (Figure 8a). The molecular packing extends via
Cl···Cl type I interaction
17
(3.472(2) Å, θ1=θ2= 135.8°) and
CH···Cl interaction (H···Cl, 2.92(3) Å) in a zigzag chain
along the b-axis (Figure 8b). Carboxylate CO bond distances
(1.252(3), 1.263(3) Å) and pyridinium CNC bond angle
(121.7(2)°) are consistent with a salt structure.
4-Aminopyridinium Meclofenamate Salt Hydrate
(1:1:1, MFA4-APYH2O). Proton transfer occurred from
MFA to pyridine base in this salt hydrate. The expected
carboxylatepyridinium R22(7) ring motif is absent presumably
due to stronger H bonds of carboxylate with water molecules.
Two water molecules now make a dimeric R44(12) ring motif
12
capped with two carboxylates (Scheme 1) through OH···O
(O···O, 2.761(4) Å, 2.713(4) Å) hydrogen bonds. The 4-APY
cation interacts with two water molecules through both ion-
ic N+H···O(N···O, 2.837(4) Å) (Figure 9a) and neutral
NH···O(N···O, 2.993 (4) Å) and one meclofenamate ion
through NH···O(N···O, 3.018 (4) Å) hydrogen bonds
(Figure 9b). Water molecules are strongly hydrogen bonded in
channels formed by MFA and 4-APY ions along the c-axis
(Figure 9c). Carboxylate nature (CO1.261,1.266ÅandC
NC121.1°) and water stoichiometry by TGA (Figure S2,
Supporting Information) were confirmed (calc 4.40%, obsd.
4.40%) in the monohydrate salt. Hydrogen bonding and unit cell
parameters similarity of MFAPPZH2O and MFA4-APY
H2O (1:1:1) suggest 2D isostructurality and isomorphism, but
there are differences in the 3D packing (Figure 7c vs 9c).
The conformation of MFA is different in crystal structures
(Table 3). Both conformers A and B are well distributed
in cocrystals/salts. A small contribution from an alternate con-
formation could be detected in a few cases (e.g., MFAINA
cocrystal) by the unassigned difference electron density (Q
peaks of about 0.50.7 electron), but it was difficult to refine
it as Me group occupancy. The small amount of Me group
disorder in these structures is difficult to model accurately. The
intramolecular NH···O hydrogen bond locks the anthranilic
acid fragment in a planar conformation, while the phenyl ring
bearing the Cl and Me groups can rotate around the NC
bond. The formation of cocrystal or salt followed the ΔpKa
rule,
5
that is, ΔpKa> 3 gives salt, ΔpKa< 0 is for cocrystal, and
the region 0 < ΔpKa< 3 is a difficult to predict zone. The pKa
s
(calculated using SPARC calculator in water medium) are listed
in Table 4. The acidacid dimer of MFA crystal structure is
replaced by acidpyridine heterosynthon in MFAINA and
MFABPY. Charge assisted N+H···Ohydrogen bonds
reproducibly give salts of MFA and piperazine/aminopyridine
base. The N base inserts between the carboxylic acid groups to
replace the carboxylic acid OH···O homosynthon with the
pyridinium/piperazinumcarboxylate charge assisted N+H···O
Figure 6. Two crystallographic molecules of meclofenamate (A = thick
bond, B = ball-stick) and two half molecules of piperazine cations form
N+H···Ohydrogen bonds in the 2:1 salt of MFAPPZ.
Figure 7. (a) Tetramer R44(12) ring motif between two meclofenamate anions and two water molecules in MFAPPZH2O (b) A piperazinium
cation is surrounded by two similar cations, one water molecule, and one meclofenamate anion. (c) Water molecules are present in channels along
the c-axis.
Figure 8. (a) Pyridine (NH+)···OOC synthon in MFA2-APY
(conformer B of drug). (b) Cl···Cl and CH···Cl interactions connect
the molecules in a 1D wavelike chain.
Crystal Growth & Design Article
dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 202320362028
heterosynthon. This strong heterosynthon in MFA salts/
cocrystals is a reliable tool for crystal engineering.
Powder X-ray Diffraction. Powder X-ray diffraction
(PXRD)
18
is a reliable characterization technique to establish
the formation of new solid materials. Rapid fingerprintingof
the product phase (cocrystal/salt) compared to characteristic
peaks in the starting materials (drug, coformer) is possible
by eye-balling of the line patterns. PXRD of new materials
prepared in this work (Figure S3, Supporting Information)
confirms the purity and homogeneity of each crystalline phase
by an excellent overlay of the experimental PXRD with the
calculated lines from the crystal structure. Calculated PXRD
line patterns of MFAPPZ-M and MFAPPZ-O salt poly-
morphs (prepared in mixed state) are compared in Figure S4,
Supporting Information.
Thermal Analysis. A change in the melting point of
cocrystal/salt in the DSC thermogram is usually indicative of
a new phase. Any dissociation/decomposition and/or phase
changes upon heating are indicated by endo-/exotherm in the
DSC trace. The monoclinic and orthorhombic polymorphs of
MFAPPZ (1:1) melt at 162.6 and 145.8 °C suggesting that
the high melting monoclinic form is more stable (Figure 10).
The metastable orthorhombic form could not be reproduced in
bulk scale purity. The DSC of polymorphic mixture (O + M)
and pure monoclinic form are shown to compare their thermal
behavior. Melting points are listed in Table 5. Dehydration of
MFAPPZH2O resulted in an anhydrate that matched with
the stable monoclinic form (MFAPPZ-M). The dissociation
of the cocrystal/salt to MFA above 240 °C is indicated by
the broad endotherm after melting (see Figure S5, Supporting
Figure 9. (a) Two meclofenamate anions (conformer A) and two water molecules form R44(12) ring through OH···Ohydrogen bond in MFA
4-APYH2O. (b) 4-Aminopyridinium cation interacts with one meclofenamate anion and two water molecules through NH···Oand N+H···O
hydrogen bonds. (c) Water molecules reside in channels formed along the c-axis.
Table 3. Torsion Angles (°) in MFA Crystal Structures
C2C7N1C8
(deg)
C7N1C8C9
(deg)
C7N1C8C13
(deg)
C7C2C1O1
(deg)
C7C2C1O2
(deg) conformer of MFA
in crystal structure
MFA 173.46(2) 92.72(3) 86.44(3) 1.54(4) 177.71(2) disorder
MFAINA 163.21(2) 99.01(2) 81.84(2) 4.52(3) 174.63(2) A
MFABPY 172.84(1) 86.50(1) 93.76(1) 3.71(2) 175.45(1) A
MFAPPZ-M 165.93(1) 98.87(1) 83.43(2) 4.20(2) 176.64(1) B
MFAPPZ-O 167.98(3) 66.12(4) 115.44(3) 0.16(4) 179.45(2) A
MFAPPZ 157.82(1), 161.44(1) 98.31(1), 84.42(1) 84.52(1), 95.47(1) 6.28(1), 14.42(1) 171.35(1), 165.67(1) A + B
MFAPPZH2O 167.43(2) 108.04(3) 74.59(3) 8.59(3) 173.55(2) B
MFA2-APY 177.91(7) 71.22(11) 111.85(9) 6.49(11) 174.43(7) B
MFA4-APYH2O 171.60(4) 84.65(5) 98.50(5) 5.55(6) 176.02(4) A
Table 4. Coformers Attempted to Make Cocrystal/Salts with
MFA and ΔpKaValues
a
pKa/pKb(water) ΔpKacocrystal/salt
meclofenamic acid 4.56
isonicotinamide 4.17 0.39 1:1 cocrystal
4,4-bipyridine 4.61 0.05 1:0.5 cocrystal
piperazine 9.72 5.16 1:1 salt polymorphs
piperazine 9.72 5.16 1:1:1 salt hydrate
piperazine 9.72 5.16 2:1 salt
2-amino pyridine 6.68 2.12 1:1 salt
4-amino pyridine 8.59 4.03 1:1:1 salt hydrate
a
pKa
s were calculated in water using SPARC pKacalculator, http://
archemcalc.com/sparc/test/login.cfm?CFID=11677&CFTOKEN=
94786654.
Figure 10. DSC endotherm comparisons for MFAPPZ (1:1) salt
polymorphs monoclinic (M) and orthorhombic (O).
Crystal Growth & Design Article
dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 202320362029
Information). The dissociation of salts/cocrystals to MFA is
similar to that for Diclofenac.
21e
FT-IR and FT-Raman Spectroscopy. Spectroscopic
analysis (IR and Raman)
19
showed clear differences in hydrogen
bonding of salt/cocrystal compared to the pure components.
Such peak shifts are useful to know specific functional groups
which are involved in intermolecular interactions. Generally
speaking, for fenamic acids,
19b
NH stretch appears at 3300
3350 cm1. The band at about 3335 cm1arises from the
amino group internally hydrogen bonded to the CO of MFA.
Carboxylic acid CO stretch is normally at 17001720 cm1,
but due to intramolecular hydrogen bonds in MFA the CO
stretch is red-shifted at 1655 cm1. Bands due to CO stretch
and OH bend appear at 1256 cm1and 1400 cm1. The car-
boxylate anion in salts showed two bands, a strong asymmetric
stretch at 16501550 cm1, and a weaker symmetric stretch
near 1450 cm1. FT-IR and FTRaman frequencies are sum-
marized in Tables 6 and 7.
Solid State NMR Spectroscopy. ss-NMR spectroscopy
20
of MFAPPZ (1:1), MFAPPZ monohydrate (1:1:1), and
MFAPPZ (2:1) was informative about local short-range order
in these solid-state structures (Figure 11, Table S3). Two peaks
at δ18.34 and 20.06 ppm represent the methyl group of MFA
which is disordered over two positions. The chemical shifts in
the NMR spectra of MFAPPZ-M (1:1) and MFAPPZ
H2O (1:1:1) salts are similar because the same conformer B of
MFA is present in both structures. There is an extra 13C peak in
MFAPPZ (2:1) for piperazinium cation at δ37.33 ppm,
methyl carbon at 21.23 ppm, and carboxylate peak at δ175.07
ppm, consistent with the crystal structure which showed two
MFA conformers A and B and two nonequivalent piperazinium
cations. The downfield region peaks in the ss-NMR spectrum
of salts (COO) are shifted relative to the free acid (COOH).
Solution Mediated Phase Transformations. Solution
mediated phase transformations
21
are common in pharmaceuti-
cals, that is, the transformation of one phase to another in a
suspension or slurry medium. Such phase changes can also
occur upon wet granulation and thus need to be monitored in
dosage formulation. 50% EtOHwater solvent was used be-
cause the drug and cocrystals/salts are soluble in this medium.
The same solvent system was used for solubility and dissolu-
tion experiments (discussed next). Piperazinium meclofenamate
(1:1) and (1:1:1) salt hydrate converted to piperazinium
meclofenamate (2:1) after 24 h slurry in 50% EtOHwater at
37 °C (Figures 12 and 13). The product 2:1 salt converted to
MFA after another 24 h in the same slurry medium (Figure 14).
Fini et al.
21e
reported 1:1 and 2:1 salts of diclofenac-pirperazine
while our work was in review. Their 1:1 salt transforms to the
less soluble 2:1 salt after one week in distilled water during
slurry experiments, similar to our results. We surmise that the
more soluble coformer piperazine dissociates from the salt in
the aqueous medium and the less soluble drug MFA preci-
pitates after 48 h. The dissociation of piperazine from the salt
occurs in stages: half equivalent in 24 h (1:1 to 2:1 salt),
and then another half equivalent in next 24 h (to give MFA)
as confirmed by PXRD of the solid residue. The stability of
Table 5. Melting Point of Cocrystal/Salt Compared with
MFA and Coformers
mp of MFA/
coformer (°C) mp of cocrystal/salt (°C)
(DSC, Tonset)
MFA 257260
MFAINA 155158 176.8
MFABPY 110114 207.1
MFAPPZ-M 106108 169.3
MFAPPZ-O 106108 144.4
MFAPPZH2O 106108 97.3, 163.3
MFAPPZ 106108 171.9
MFA2-APY 5760 145.6
MFA4-APYH2O 157161 93.6, 111.7
Table 6. FT-IR Stretching Modes (νs,cm
1) of MFA and Its Cocrystals/Salts
a
NH stretch CO stretch NH bend CO stretch (asym) CO stretch (sym)
MFA 3335.1 1655.4 1575.3 1437.4 1256.6
MFAINA 3447.0, 3227.2 1694.8 1559.5 1449.5 1261.5
MFABPY 3245.9 1675.4 1600.4, 1581.4 1451.0 1259.7
MFAPPZ-M 3216.9 1582.4 1452.9 1289.1
MFAPPZH2O 3208.4, 3172.6 1576.4 1452.1 1282.5
MFAPPZ 3497.3, 3267.0 1577.8 1452.5 1287.6
MFA2-APY 3268.5, 3199.8 1577.7 1449.1, 1458.1 1288.1, 1255.0
MFA4-APYH2O 3436.7, 3346.2, 3301.2 1574.7 1455.2 1288.0
a
IR spectrum of MFAPPZ-O salt could not be recorded due to insufficient sample and contamination from the stable monoclinic polymorph in the
bulk phase.
Table 7. FT-Raman Stretching Modes (νs,cm
1) of MFA and Its Cocrystal/Salts
a
CH stretch CO stretch NH bend CO stretch (asymm) CO stretch (sym)
MFA 3079.6, 2925.2 1655.2 1577.5 1439.1 1242.3
MFAINA 3071.4 1675.0 1602.1 1449.2 1243.2
MFABPY 3086.9, 3053.6 1663.2 1582.3 1450.8 1239.8
MFAPPZ-M 3070.8, 2983.6 1580.1 1448.4 1271.6
MFAPPZH2O 3079.2, 3002.6 1581.1 1450.2 1274.4
MFAPPZ 3070.1, 2973.4 1582.5 1445.5 1281.0
MFA2-APY 3073.5, 2924.5 1579.4 1448.4 1273.2
MFA4-APYH2O 3059.7, 2924.4 1573.7 1453.4 1267.1
a
Raman spectrum of MFAPPZ-O salt could not be recorded due to insufficient sample and contamination from the stable monoclinic polymorph
in the bulk phase.
Crystal Growth & Design Article
dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 202320362030
Figure 12. PXRD comparison of MFAPPZ-M (black) after 24 h slurry with the calculated X-ray lines of MFAPPZ-M (red) and MFAPPZ 2:1
salt (blue) in 50% EtOHwater medium. The monoclinic polymorph undergoes solvent mediated phase transformation from MFAPPZ-M to
MFAPPZ salt after 24 h.
Figure 11. Solid state 13C NMR spectra of MFAPPZ-M (1:1, pink) and MFAPPZH2O (1:1:1, green) and MFAPPZ (2:1, yellow) salts along
with the pure components. The peaks in the salt are shifted relative to the pure components because of carboxylic acid to carboxylate in MFA. The
regions of main chemical shift differences are indicated in red line border.
Crystal Growth & Design Article
dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 202320362031
variable stoichiometry salts in slurry medium may be related
to their hydrogen bonding in the crystal structures: the 1:1 salt
contains one N+H···Oand one NH···O hydrogen bond
between meclofenamate and piperazinium, whereas one meclo-
fenamate interacts with a dipiperazinium through two N+
H···Ohydrogen bonds in the 2:1 salt. The stronger ionic
H bonds result in greater stability of the 2:1 salt over the 1:1
salt. MFAPPZ monohydrate (1:1:1) showed similar results.
The cocrystals, on the other hand, were more stable to slurry
conditions: MFAINA transformed to 86% MFA and 14%
cocrystal remained after 24 h, as confirmed by PXRD. A cali-
bration sample of 85% MFA and 15% cocrystal suggested that
the standard deviation of PXRD is ±3%. MFABPY cocrystal
was unusually stable for up to 72 h in 50% EtOHwater
medium (Figure S6, Supporting Information). These stability
trendsmaybeascribedtocongruentandincongruent
Figure 14. PXRD comparison of MFAPPZ 2:1 (black) after 24 h slurry with the calculated X-ray lines of MFAPPZ 2:1 (red) and MFA (blue) in
50% EtOHwater medium. MFAPPZ 2:1 salt transforms to pure MFA after 24 h.
Figure 13. PXRD comparison of MFAPPZH2O (black) after 24 h slurry with the calculated X-ray lines of MFAPPZH2O (red) and MFA
PPZ 2:1 (blue) in 50% EtOHwater medium. The hydrate MFAPPZH2O transforms to MFAPPZ salt after 24 h.
Crystal Growth & Design Article
dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 202320362032
systems.
22
When the solubility of the drug and the coformer are
similar in the same medium (both are insoluble/less soluble),
such systems are congruent and tend to be stable. For incon-
gruent systems, the high solubility coformer leaches into the
solvent medium in which it readily dissolves, and in this way
the salt/cocrystal dissociates faster. The soluble cocrystals will
usually have a highly soluble coformer partner and this in turn
makes the salt/cocrystal of limited stability. This in effect puts a
limit to the stability of more soluble pharmaceutical cocrystals
in the slurry medium. The stability order in aqueous slurry
medium is (less to more stable): MFAPPZ (1:1) < MFA
PPZ monohydrate (1:1:1) < MFAPPZ (2:1) < MFAINA
(1:1) < MFABPY (1:0.5). Solution-mediated phase transfor-
mations
21
can occur during dissolution and solubility measure-
ments (discussed next), and the above stability order will help
to explain any unusual trends.
Solubility and Dissolution Experiments. The rate of
dissolution and solubility of the solid drug in water or aqueous
solvent mixtures is necessary for good oral bioavailability. The
aqueous solubility of the drug must be at least 100 mg/L
for fast dissolution of the tablet.
1c
The aqueous solubility of
MFA is 30 mg/L and other fenamic acids (e.g., mefenamic acid,
tolfenamic acid and diclofenamic acid) are also low solubility
BCS class II drugs. The sodium salt of meclofenamic acid
(MFA-SS) is marketed as capsules (solubility > 250 g/L). The
aqueous solubility for cocrystals and salts of MFA were
measured in 50% EtOHwater medium at 37 °C (Table 8).
Although solubility is a good indicator of drug bioavailability,
the method is applicable only for those solid-state drug forms
which are stable in the test medium/wet slurry. The intrinsic
dissolution rate (IDR) is a kinetic parameter and is a useful
indicator for those solid forms which undergo phase trans-
formation or dissociation during the experiment. Most drugs
exert their therapeutic effect during 468 h of oral admin-
istration, and IDR is related to drug dissolution in such cases.
IDR measurements showed that piperazine salts reached peak
concentration within 3045 min of drug dissolution (Figure 15).
The most stable 2:1 MFAPPZ salt exhibited a gradual in-
crease in dissolution rate up to 24 h. The cumulative amount of
MFA dissolved (mg cm2) vs time (min) was plotted to com-
pare the IDR of piperazine meclofenamates (Figure 15a). The
highest solubility of MFAPPZ-M (1:1) is driven by the high
solubility of piperazine > MFAPPZH2O (1:1:1) because
hydrates are generally less soluble than anhydrate drugs >
(MFAPPZ (2:1) due to lower piperazine content. In contrast
to salts, MFAINA (1:1) and MFABPY (1:0.5) cocrystals
exhibited poor dissolution rates (Figure 15b), but they were
quite stable for over 24 h (Figure S6, Supporting Information).
The highest solubility salt MFAPPZ (1:1) was compared
Table 8. Solubility and Dissolution Rate of MFA and Its
Cocrystals and Salts
a
solid form
absorption
coefficient
(ε,mM
1cm1)
equilibrium solubility
after 24 h slurry in
50% EtOHwater
(mg L1)IDR ((mg cm2)
min1)
a
MFA 6.94 203.11 0.03074
MFAINA 4.10 581.43 (x 2.9) 0.04858 (x 1.6)
MFABPY 5.74 1638.07 (x 7.6) 0.03822 (x 1.2)
MFAPPZ-M 3.38 553223.061 (x 2724) 27.4976 (x 894.5)
MFAPPZ 7.95 10211.954 (x 50) 0.31477 (x 10.2)
MFAPPZH2O 4.82 270966.397 (x 1334) 9.8044 (x 318.9)
a
Number in parentheses indicates how many times higher solubility/
IDR compared to reference drug.
Figure 15. Intrinsic dissolution rate (IDR) measurements of (a)
MFAPPZ-M (1:1), MFAPPZH2O (1:1:1) and MFAPPZ (2:1)
performed up to 3 h, (b) MFAPPZ (2:1), MFAINA (1:1), MFA
BPY (1:0.5) and MFA over 24 h in 50% EtOHwater medium, and
(c) MFAPPZ-M (1:1) and MFA-SS (sodium salt of MFA) over
22 min in distilled water. The amount of MFA dissolved in the test
medium was monitored by UVvis spectroscopy.
Crystal Growth & Design Article
dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 202320362033
with MFA-SS in aqueous medium; IDR of sodium salt is 28.76
and piperazinium salt is 13.72 (mg cm2) min1(Figure 15c).
The marketed sodium salt of MFA has 2.1 times higher IDR
than the equimolar piperazine salt. PXRD of the residue of
MFAPPZ-M dissolution experiment after 11 min matched
with the 2:1 salt (70%) and monohydrate (30%), the 1:1 salt
showed transformation to piperazine salt hydrate and 2:1 salt,
whereas the MFA sodium salt transformed to its hydrate after
8 min in the dissolution medium. These results suggest that 1:1
MFAPPZ-M is more stable than the marketed sodium salt
to hydration in aqueous medium. PXRD of melofenamic acid
sodium salt after 24 h of slurry stirring in water (Figure S7,
Supporting Information) is different from the starting material
(MFA-SS, Sigma-Aldrich). Thermogravemetric analysis and Karl
Fischer titration method indicate a water content of 13.92%
(2.82.9 equivalent) in MFA-SS hydrate (Figure S8, Supporting
Information) formed in the aqueous medium, a value that was
confirmed by DSC. The aqueous solubility of sodium salt of
MFA (>250 g/L)
7b
is 46 times greater than the value for MFA
PPZ-M 1:1 salt in water (5.4 g/L), but its dissolution rate is two
times faster in the first 30 min of measurement.
CONCLUSIONS
MFA is the most potent nonsteroidal anti-inflammatory drug.
A few cocrystals and salts of MFAs were crystallized using crystal
engineering principles. All new crystalline phases were charac-
terized by X-ray diffraction, IR and ss-NMR spectroscopy, and
DSC/TGA. MFA exists in two different conformers A and B
due to m-tolyl group rotation about the NC bond. Whereas
the crystal structure of MFA has disordered methyl groups, the
drug molecule is in an ordered orientation in its cocrystals and
salts, perhaps rigidified due to stronger hydrogen bonding. The
dissolution rate and solubility of MFAPPZ-M 1:1 salt is the
highest among the novel solid-state forms studied and slightly
lower than the marketed sodium salt of MFA. Their stability to
hydration is comparable in aqueous medium. Thus, meclofena-
mate piperazinium has an equivalent dissolution and stability
profile to the marketed sodium salt. Four different piperazinium
salts were crystallized: monoclinic and orthorhombic poly-
morphs of MFAPPZ (1:1), its monohydrate, and MFAPPZ
(2:1). This is the first example of polymorphic and variable
stoichiometry piperazinium salts with X-ray crystal structures
solved to good accuracy.
EXPERIMENTAL SECTION
MFA and INA were purchased from Sigma-Aldrich (Hyderabad,
Andhra Pradesh, India) and used directly for experiments. All other
chemicals were of analytical or chromatographic grade. Melting points
were measured on a Fisher-Johns melting point apparatus. Water
filtered through a double deionized purification system (AquaDM,
Bhanu, Hyderabad, India) was used in all experiments. Single crystals
were obtained via slow evaporation of stoichiometric amounts of
starting materials in an appropriate solvent. Cocrystals and salts were
characterized by infrared spectroscopy (IR), powder X-ray diffraction
(PXRD), differential scanning calorimetry (DSC), thermogravimetric
analysis (TGA), and single crystal X-ray diffraction (SC-XRD).
Meclofenamic Acid, MFA. Normally cracked crystals of MFA
appeared after crystallization from organic solvents. Sublimation of
MFA at 190200 °C produced good quality block shaped crystal after
23 h. Melting point 257260 °C (literature value).
7a,23
Meclofenamic AcidIsonicotinamide, MFAINA (1:1) Co-
crystal. MFA (100 mg, 0.34 mmol) and INA (41.5 mg, 0.34 mmol)
were ground in a mortar pestle for 15 min using acetonitrile as solvent-
assisted grinding. After a new solid phase was confirmed by IR and
PXRD, the bulk material was dissolved in acetonitrile. Good quality
single crystals appeared at ambient conditions after 23 days, mp
175177 °C.
Meclofenamic Acid4,4-Bipyridine, MFABPY (1:0.5) Co-
crystal. MFA (100 mg, 0.34 mmol) and BPY (26.5 mg, 0.17 mmol)
were ground in a mortar pestle for 15 min using acetonitrile as solvent-
assisted grinding. After a new solid phase was confirmed by IR and
PXRD, the bulk material was dissolved in acetonitrile. Thick plate
crystals were harvested at ambient conditions after 23 days, mp 204
207 °C.
Piperazinium Meclofenamate, MFAPPZ (1:1) Salt Poly-
morphs. MFA (100 mg, 0.34 mmol) and piperazine (29.3 mg,
0.34 mmol) were ground in a mortar pestle for 15 min using acetonitrile
as solvent-assisted grinding. After a new solid phase was confirmed by IR
and PXRD, the bulk material was dissolved in acetonitrile. Block
(monoclinic, form M) and thick long needle (orthorhombic, form O)
crystals were harvested concomitantly at ambient conditions after 23
days. Monoclinic form (plate) was obtained exclusively from CH3NO2
solvent; mp of monoclinic and orthorhombic forms are 162166 °C
and 144147 °C, respectively.
Piperazinium Meclofenamate, MFAPPZ (2:1) Salt. MFA
(100 mg, 0.34 mmol) and piperazine (29.3 mg, 0.34 mmol) were
ground in a mortar pestle for 15 min using acetonitrile as solvent-
assisted grinding. After a new solid phase was confirmed IR and
PXRD, bulk material was dissolved in EtOH. Suitable thick plate
(triclinic) crystals were harvested at room temperature after 34 days,
mp 166169 °C.
Piperazinium Meclofenamate Monohydrate, MFAPPZ
H2O (1:1:1) Salt. MFA (100 mg, 0.34 mmol) and piperazine hydrate
(35.4 mg, 0.34 mmol) were ground in a mortar pestle for 15 min After
a new solid phase was confirmed by IR and PXRD, the bulk material
was dissolved in nitromethane, mp of salt 154158 °C and dehydra-
tion temperature 9295 °C.
2-Aminopyridinium Meclofenamate, MFA2-APY (1:1) Salt.
MFA (100 mg, 0.34 mmol) and 2-aminopyridine (32 mg, 0.34 mmol)
were ground in a mortar pestle for 15 min using acetonitrile as solvent-
assisted grinding. After a new solid phase was confirmed by IR and
PXRD, the bulk material was dissolved in acetonitrile. Plate crystals
were harvested after 23 days at ambient condition, mp 143146 °C.
4-Aminopyridinium Meclofenamate Monohydrate, MFA
4-APYH2O (1:1:1) Salt. MFA (100 mg, 0.34 mmol) and 4-amino-
pyridine (32 mg, 0.34 mmol) were ground in a mortar pestle for
15 min using acetonitrile as solvent-assisted grinding. After a new solid
phase was confirmed by IR and PXRD, bulk material was dissolved in
acetonitrile. Suitable thick plate crystals were harvested after 23 days
at ambient condition, mp 104108 °C and dehydration temperature
8084 °C.
Single Crystal X-ray Diffraction. A single crystal obtained from
the crystallization solvent(s) was mounted on the goniometer of
Oxford CCD X-ray diffractometer (Yarnton, Oxford, UK) equipped
with MoKαradiation (λ= 0.71073 Å) source. Data reduction was
performed using CrysAlisPro 171.33.55 software.
24
Crystal structures
were solved and refined using Olex21.0
25
with anisotropic displace-
ment parameters for non-H atoms. Hydrogen atoms were experi-
mentally located through the Fourier difference electron density maps
in all crystal structures. All OH, NH, and CH atoms were geo-
metrically fixed using HFIX command in SHELX-TL program of
Bruker-AXS.
26
A check of the final cif file with PLATON
27
did not
show any missed symmetry. X-Seed
28
was used to prepare the figures
and packing diagrams. Crystallographic parameters of both structures
are summarized in Table 1. Hydrogen bond distances in Table 2 are
neutron-normalized to fix the DH distance to its accurate neutron
value in the X-ray crystal structures (OH 0.983 Å, NH 0.82 Å, C
H 1.083 Å). Crystallographic .cif files (CCDC Nos. 859220859229)
are available at www.ccdc.cam.ac.uk/data_request/cif or as part of the
Supporting Information.
Powder X-ray Diffraction. Bulk samples were analyzed by PXRD
on a Bruker AXS D8 diffractometer (Bruker-AXS, Karlsruhe,
Germany). Experimental conditions: CuKαradiation (λ=1.54056
Å); 40 kV; 30 mA; scanning interval 550°2θat a scan rate of
1°min1; time per step 0.5 s. The experimental PXRD patterns and
Crystal Growth & Design Article
dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 202320362034
calculated X-ray lines from the single crystal structure were compared
to confirm the purity of the bulk phase using Powder Cell.
29
Thermal Analysis. DSC and TGA were performed on a Mettler
Toledo DSC 822e module and a Mettler Toledo TGA/SDTA 851e
module, respectively. Samples were placed in open alumina pans for
TGA and in crimped but vented aluminum sample pans for DSC.
A typical sample size is 46 mg for DSC and 912 mg for TGA.
The temperature range was 30250 °C at 2 K min1for DSC and
10 K min1for TGA. Samples were purged with a stream of dry N2
flowing at 150 mL min1for DSC and 50 mL min1for TGA.
Vibrational Spectroscopy. A Thermo-Nicolet 6700 FT-IR
spectrometer (Waltham, MA, USA) with a NXR FT-Raman Module
(Nd:YAG laser source, 1064 nm wavelength) was used to record IR
and Raman spectra. IR spectra were recorded on samples dispersed in
KBr pellets. Raman spectra were recorded on samples contained
in standard NMR diameter tubes or on compressed samples contained
in a gold-coated sample holder.
Solid-State NMR Spectroscopy. Solid-state 13C NMR (ss-NMR)
spectroscopy provides structural information about differences in
hydrogen bonding, molecular conformations, and molecular mobility
in the solid state. The solid-state 13C NMR spectra were obtained on
a Bruker Ultrashield 400 spectrometer (Bruker BioSpin, Karlsruhe,
Germany) utilizing a 13C resonant frequency of 100 MHz (magnetic
field strength of 9.39 T). Approximately 100 mg of crystalline sample
was lightly packed into a zirconium rotor with a Kel-F cap. The cross-
polarization, magic angle spinning (CP-MAS) pulse sequence was used
for spectral acquisition. Each sample was spun at a frequency of 5.0 ±
0.01 kHz and the magic angle setting was calibrated by the KBr
method. Each data set was subjected to a 5.0 Hz line broadening factor
and subsequently Fourier transformed and phase corrected to produce
a frequency domain spectrum. The chemical shifts were referenced to
TMS using glycine (δglycine = 43.3 ppm) as an external secondary
standard.
Dissolution and Solubility Measurements. Intrinsic dissolution
rate (IDR) and solubility measurements were carried out on a USP-
certified Electrolab TDT-08 L Dissolution Tester (Electrolab, Mumbai,
MH, India). A calibration curve was obtained for all the new solid
phases including MFA by plotting absorbance vs concentration UVvis
spectra curves on a Thermo Scientific Evolution EV300 UVvis spec-
trometer (Waltham, MA, USA) for known concentration solutions in
50% EtOHwater medium. The mixed solvent system (EtOHwater)
was selected for its higher solubility of MFA in this medium. The slope
of the plot from the standard curve gave the molar extinction coeffi-
cient (ε) by applying the BeerLamberts law. Equilibrium solubility
was determined in 50% EtOHwater medium using the shake-flask
method.
27
To obtain the equilibrium solubility, 100 mg of each solid
material was stirred for 24 h in 5 mL of 50% EtOHwater at 37 °C,
and the absorbance was measured at 318 nm. The concentration of the
saturated solution was calculated at 24 h, which is referred to as the
equilibrium solubility of the stable solid form.
100 mg of the solid (drug, cocrystal, salt) was taken in the intrinsic
attachment and compressed to a 0.5 cm2pellet using a hydraulic press
at a pressure of 2.5 ton/inch2for 2 min. The pellet was compressed to
provide a flat surface on one side and the other side was sealed. Then
the pellet was dipped into 900 mL of 50% EtOHwater medium at
37 °C with the paddle rotating at 150 rpm. At a regular interval of 5
10 min, 5 mL of the dissolution medium was withdrawn and replaced
by an equal volume of fresh medium to maintain a constant volume.
Samples were filtered through 0.2 μm nylon filter and assayed for drug
content spectrophotometrically at 318 nm on a Thermo-Nicolet
EV300 UVvis spectrometer. There was no interference to MFA
UVvis maxima at 318 nm by coformers INA and bipyridine which
absorb strongly at 250270 nm. Piperazine is UVvis inactive. The
amount of drug dissolved in each time interval was calculated using the
calibration curve. The linear region of the dissolution profile was used
to determine the intrinsic dissolution rate (IDR) of the compound
(= slope of the curve, that is, the amount of drug dissolved divided by
the surface area of the disk (0.5 cm2) per minute). The dissolution
rates for MFA, its cocrystals, and salts were computed from their IDR
values. Similarly IDR experiments of MFAPPZ-M and MFA-SS salts
were carried out in distilled water and absorbance was measured at
318 nm in a UVvis spectrophotometer.
ASSOCIATED CONTENT
*
SSupporting Information
Crystallographic information files; Refcodes of cocrystals and
salts of piperazine (Table S1) and piperazinium dications
with carboxylate anions (Table S2); 13C solid-state NMR of
piperazinium meclofenamate salts compared to values for the
pure coformers (Table S3); ORTEP diagrams (Figure S1);
TGA results (Figure S2); PXRD results (Figures S3, S6, S7);
comparison of calculated X-ray diffraction lines of monoclinic
and orthorhombic forms of piperazinium meclofenamate (1:1)
salts (Figure S4); DSC endotherm (Figure S5); DSC and TGA
thermograms (Figure S8). This material is available free of
charge via the Internet at http://pubs.acs.org.
AUTHOR INFORMATION
Corresponding Author
*E-mail: ashwini.nangia@gmail.com.
Notes
The authors declare no competing financial interest.
ACKNOWLEDGMENTS
We thank the DST (SR/S1/OC-67/2006), JC Bose fellowship
(SR/S2/JCB-06/2009), and CSIR (01(2410)/10/EMR-II) for
research funding, and DST (IRPHA) and UGC (PURSE grant)
for providing instrumentation and infrastructure facilities. P.S.
and G.B. thank the UGC for fellowship. We thank Dr. Naba
Kamal Nath for his assistance to resolve disorder issues in
crystal structure refinement of MFA.
REFERENCES
(1) (a) Lipinski, C. Am. Pharm. Rev. 2002,5, 82. (b) Vippagunta,
S. R.; Brittain, H. G.; Grant, D. J. W. Adv. Drug Delivery Rev. 2001,48,
3. (c) Byrn, S. R.; Pfeiffer, R. R.; Stowell, J. G. Solid-State Chemistry of
Drugs, 2nd ed.; SSCI, Inc.: West Lafayette, IN, 1999.
(2) (a) Desiraju, G. R. Crystal Engineering: The Design of Organic
Solids; Elsevier: Amsterdam, 1989. (b) Desiraju, G. R. Angew. Chem.,
Int. Ed. 2007,46, 8342. (c) Blagden, N.; Matas, M. D.; Gavan, P. T.;
York., P. Adv. Drug Delivery Rev. 2007,59, 617.
(3) (a) Karki, S.; Friščić, T.; Fábián, L.; Laity, P. R.; Day, G. M.;
Jones, W. Adv. Mater. 2009,21, 3905. (b) Childs, S. L.; Chyall, L. J.;
Dunlap, J. T.; Smolenskaya, V. N.; Stahly, B. C.; Stahly, G. P.
J. Am. Chem. Soc. 2004,126, 13335. (c) Remenar, J. F.; Morissette,
S. L.; Peterson, M. L.; Moulton, B.; MacPhee, J. M.; Guzmán, H. R.;
Almarsson, O
̈J. Am. Chem. Soc. 2003,125, 8456. (d) Cheney, M. L.;
Shan, N.; Healey, E. R.; Hanna, M.; Wojtas, L.; Zaworotko, M. J.; Sava,
V.; Song, S.; Sanchez-Ramos, J. R. Cryst. Growth Des. 2010,10, 395.
(e) Sanphui, P.; Goud, N. R.; Khandavilli, U. B. R.; Nangia, A. Cryst.
Growth Des. 2011,11, 4135.
(4) (a) Portell, A.; Barbas, R.; Font-Bardia, M.; Dalmases, P.;
Prohens, R.; Puigjaner, C. CrystEngComm 2009,11, 791. (b) Thakuria,
R.; Nangia, A. CrystEngComm 2011,13, 1759. (c) Banerjee, R.; Bhatt,
P. M.; Ravindra, N. V.; Desiraju, G. R. Cryst. Growth Des. 2005,5,
2299.
(5) (a) Aakeröy, C. B.; Fasulo, M. E.; Desper, J. Mol. Pharmaceutics
2007,4, 317. (b) Sarma, B.; Nath, N. K.; Bhogala, B. R.; Nangia, A.
Cryst. Growth Des. 2009,9, 1546. (c) Childs, S. L.; Stahly, G. P.; Park,
A. Mol. Pharmaceutics 2007,4, 323.
(6) (a) Flower, R. J. Pharm. Res. 1974,26, 33. (b) Moser, P.;
Salimann., A.; wissenberg., I. J. Med. Chem. 1990,33, 2358. (c) Bowen,
E. J.; Eland, J. H. D. Proc. Chem. Soc. 1963, 202. (d) Narsinghania, T.;
Chaturvedi, S. C. Biorg. Med. Chem. Lett. 2006,16, 461. (e) Peretz, A.;
Degani, N.; Nachman, R.; Uziyel, Y.; Gibor, G.; Shabat, D.; Attali, B.
Crystal Growth & Design Article
dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 202320362035
Mol. Pharmacol. 2005,67, 1053. (f) Belsole, S. C.; Chester, N. J. U.S.
Patent 1986, 4602040. (g) Philip, J.; Szulczewski, D. H. J. Pharm. Sci.
1973,62, 1479. (h) http://www.elephantcare.org/Drugs/meclofen.htm.
(7) (a) http://www.drugbank.ca/drugs/DB00939. (b) http://www.
orgyn.com/resources/genrx/D001710.asp.
(8) Fábián, L.; Hamill, N.; Eccles, K. S.; Moynihan, H. A.; Maguire,
A. R.; McCausland, L.; Lawrence, S. E. Cryst. Growth Des. 2011,11,
3522.
(9) Murthy, H. M. K.; Vijayan, M. Acta Crystallogr. 1981,B37, 1102.
(10) Dhanraj, V.; Vijayan, M. Biochim. Biophys. Acta 1987,924, 135.
(11) (a) Takasuka, M.; Nakai, H.; Shiro, M. J. Chem. Soc., Perkin
Trans. II 1982, 1061. (b) McConnell, J. F.; Company, F. Z. Cryst.
Struct.Commun. 1976,5, 861. (c) Lόpez-Mejías, V.; Kapyf, J. W.;
Matzger, A. J. J. Am. Chem. Soc. 2009,131, 4554.
(12) (a) Etter, M. C.; Macdonald, J. C. Acta Crystallogr. 1990,B46,
256. (b) Bernstein, J.; Davis, R. E.; Shimoni, L.; Chang, N.-L. Angew.
Chem., Int. Ed. Engl. 1995,34, 1555.
(13) (a) Stahl, P. H., Wermuth, C. G., Eds.; Handbook of Pharma-
ceutical Salts: Properties, Selection and Use; Verlag Helvetica Chimica
Acta: Zürich, 2002. (b) Generally Regarded as Safe: http://www.cfsan.
fda.gov/rdb/opa-gras.html and http://www.cfsan.fda.gov/dms/
grasguid. html.
(14) Cambridge Structural Database, Ver. 5.32, 2010, November
2011 update, Cambridge Crystallographic Data Center, UK, www.
ccdc.cam.ac.uk.
(15) Cerius2Materials Studio, http://accelrys.com/products/cerius2,
Lattice energy calculation of crystal structures.
(16) Fonari, M. S.; Ganin, E. V.; Vologzhanina, A. V.; Antipin, M. Y.;
Kravtsov, V. C. Cryst. Growth Des. 2010,10, 3647.
(17) (a) Hathwar, V. R.; Roopan, S. M.; Subashini, R.; Khan, F. N.;
Guru Row, T. N. J. Chem. Sci. 2010,122, 677. (b) Nath, N. K.; Saha,
B. K.; Nangia, A. New J. Chem. 2008,32, 1693. (c) Rajput, L.;
Chernyshev, V. V.; Biradha, K. Chem. Commun. 2010,46, 6530.
(18) (a) Karki, S.; Friščić, T.; Fábián, L.; Jones, W. CrystEngComm
2010,12, 4038. (b) Remenar, J. F.; Peterson, M. L.; Stephens, P. W.;
Zhang, Z.; Zimenkov, Y.; Hickey, M. B. Mol. Pharmaceutics 2007,4,
386. (c) André, V.; Fernandes, A.; Santos, P. P.; Duarte, M. T. Cryst.
Growth Des. 2011,11, 2325.
(19) (a) Silverstein, R. M. Spectrometric Identification of Organic
Compounds, 6th ed.; John Wiley & Sons, Inc.: New York, 2002;
pp 71143. (b) Gilpin, R. K.; Zhou, W. Vib. Spectrosc. 2005,37, 53.
(c) Romero, S.; Escalera, B.; Bustamante, P. Int. J. Pharm. 1999,178,
193. (d) Smith, E. Dent, G. Modern Raman Spectroscopy: A Practical
Approach; John-Wiley: New York, 2005.
(20) (a) Vogt, F. G.; Clawson, J. S.; Strohmeier, M.; Edwards, A. J.;
Pham, T. N.; Watson, S. A. Cryst. Growth Des. 2009,9, 921. (b) Li,
Z. J.; Abramov, Y.; Bordner, J.; Leonard, J.; Medek, A.; Trask, A. V.
J. Am. Chem. Soc. 2006,128, 8199.
(21) (a) Greco, K.; Mcnamara, D. P.; Bogner, R. J. Pharm. Sci. 2011,
100, 2755. (b) Murphy, D.; Rodríguez-Cintrón, F.; Langevin, B.; Kelly,
R. C.; Rodríguez-Hornedo, N. Int. J. Pharm. 2002,246, 121. (c) Ferrari,
E. S.; Davey, R. J. Cryst. Growth Des. 2004,4, 1061. (d) Cherekuvada,
S.; Babu, N. J.; Nangia, A. J. Pharm. Sci. 2011,100, 3233. (e) Fini, A.;
Cavallari, C.; Bassini, G.; Ospitalli, F.; Morigi, R. J. Pharm. Sci. 2012,
DOI: 10.1002/jps.23052.
(22) (a) Friščić, T.; Childs, S. L.; Rizvi, S. A. A.; Jones, W.
CrystEngComm 2009,11, 418. (b) Alhalaweh, A.; Velega, S. P. Cryst.
Growth Des. 2010,10, 3302. (c) Rodríguez-Hornedo, N.; Nehm, S. J.;
Seefeldt, K. F.; Pagán-Torres, Y.; Falkiewicz, C. J. Mol. Pharmaceutics
2006,3, 362. (d) Goud, N. R.; Gangavaram, S.; Suresh, K.; Pal, S.;
Manjunatha, N. G.; Nambiar, S. J. Pharm. Sci. 2012,101, 664.
(23) Fawzi, M.; Mahjour, M. U.S. Patent, 5,373,022, 1994.
(24) CrysAlis CCD and CrysAlis RED, versions 1.171.33.55; Oxford
Diffraction: Oxford, 2008.
(25) Diffraction Ltd, Yarnton, Oxfordshire, UK.
(26) Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A. K.;
Puschmann, H. OLEX2: A complete structure solution, refinement
and analysis program. J. Appl. Crystallogr. 2009,42, 339.
(27) SMART, version 5.625 and SHELX-TL, version 6.12; Bruker
AXS Inc.: Madison, Wisconsin, USA, 2000.
(28) Spek, A. L. PLATON, A Multipurpose Crystallographic Tool;
Utrecht University: Utrecht, Netherlands, 2002. Single-crystal struc-
ture validation with the Program PLATON. J. Appl. Crystallogr. 2003,
36,7.
(29) Barbour, L. J. X-Seed, Graphical interface to SHELX-97 and POV-
Ray, Program for Better Quality of Crystallographic Figures; University of
Missouri-Columbia: Missouri, USA, 1999
(30) Kraus, N.; Nolze, G. Powder Cell, version 2.3, A Program for
Structure Visualization, Powder Pattern Calculation and Profile Fitting;
Federal Institute for Materials Research and Testing: Berlin, Germany,
2000.
(31) Glomme, A.; Marz, J.; Dressman, J. B. J. Pharm. Sci. 2005,94,1.
Crystal Growth & Design Article
dx.doi.org/10.1021/cg300002p |Cryst. Growth Des. 2012, 12, 202320362036
... In this context, several efforts have been made to modify NSAID scaffolds to minimize possible side effects such as gastrointestinal [24] and cardiovascular [25] risks associated with their prolonged use, and to develop more effective drugs with novel mechanisms of action [26]. Likewise, the existence of some polymorphic forms of NSAIDs and NSAID derivatives has been identified [27][28][29]. For instance, both flufenamic [30] and tolfenamic [31] acids presented nine polymorphs each, respectively, and aceclofenac was reported to have at least six polymorphs [32]. ...
Article
Full-text available
Stability, thermal characterization, and identification of possible polymorphism are relevant in the development of novel therapeutic drugs. In this context, thirty new nonsteroidal anti-inflammatory drug (NSAID) derivatives containing selenium (Se) as selenoesters or diacyl diselenides with demonstrated anticancer activity were thermally characterized in order to establish thermal stability criteria and detect possible polymorphic forms. Compounds were analyzed by a combination of thermogravimetry, differential scanning calorimetry, and X-ray diffraction techniques, and five different calorimetric behaviors were identified. Two compounds based on naproxen ( I.3d and I.3e ) and an indomethacin-containing derivative ( II.2 ) presented two crystalline forms. The stability under acid, alkaline and oxidative conditions of selected polymorphs was also assessed using high-performance liquid chromatography. In addition, the cytotoxic activity of Se-NSAID crystalline polymorphs was studied in several cancer cell lines in vitro. Remarkably, no significant differences were found among the polymorphic forms tested, thus proving that these compounds are thermally qualified for further drug development. Graphical abstract
... However, it should be noted that the slight shift in the resulting C=O stretching is mainly due to the hydrogen bonding. Furthermore, the vibrational bands around 1270 cm − 1 and 1430 cm − 1 for compounds 1 and 2 are ascribed to the symmetric and antisymmetric stretching frequencies of the C -O, respectively [79]. In addition, the IR band at 1560 cm − 1 is assigned to the -NH bending vibration of piperazine for compound 2. ...
... However, it should be noted that the slight shift in the resulting C=O stretching is mainly due to the hydrogen bonding. Furthermore, the vibrational bands around 1270 cm -1 and 1430 cm -1 for compounds 1 and 2 are ascribed to the symmetric and antisymmetric stretching frequencies of the C-O, respectively [79]. In addition, the IR band at 1560 cm -1 is assigned to the -NH bending vibration of piperazine for compound 2. Fig. 8. FT-IR spectra of compound 1, 2 and starting materials. ...
... Salt formation, which can increase the solubility and dissolution of APIs, still has the greatest potential [14][15][16]. In terms of stability, however, salts have the drawback that they are more likely to form hydrates compared to cocrystals and many exhibit strong hygroscopicity, making them prone to deliquescence upon moisture sorption [17,18]. ...
Article
Full-text available
Vonoprazan (VPZ) is the first-in-class potassium-competitive acid blocker (P-CAB), and has many advantages over proton pump inhibitors (PPIs). It is administered as a fumarate salt for the treatment of acid-related diseases, including reflux esophagitis, gastric ulcer, and duodenal ulcer, and for eradication of Helicobacter pylori. To discover novel cocrystals of VPZ, we adopted an artificial neural network (ANN)-based machine learning model as a virtual screening tool that can guide selection of the most promising coformers for VPZ cocrystals. Experimental screening by liquid-assisted grinding (LAG) confirmed that 8 of 19 coformers selected by the ANN model were likely to create new solid forms with VPZ. Structurally similar benzenediols and benzenetriols, i.e., catechol (CAT), resorcinol (RES), hydroquinone (HYQ), and pyrogallol (GAL), were used as coformers to obtain phase pure cocrystals with VPZ by reaction crystallization. We successfully prepared and characterized three novel cocrystals: VPZ–RES, VPZ–CAT, and VPZ–GAL. VPZ–RES had the highest solubility among the novel cocrystals studied here, and was even more soluble than the commercially available fumarate salt of VPZ in solution at pH 6.8. In addition, novel VPZ cocrystals had superior stability in aqueous media than VPZ fumarates, demonstrating their potential for improved pharmaceutical performance.
Article
Full-text available
New derivatives of non-steroidal anti-inflammatory drugs were synthesized via conjugation with L-amino acid isopropyl esters. The characteristics of the physicochemical properties of the obtained pharmaceutically active ionic liquids were determined. It has been shown how the incorporation of various L-amino acid esters as an ion pair affects the properties of the parent drug. Moreover, the antimicrobial activity of the obtained compounds was evaluated. The proposed structural modifications of commonly used drugs indicate great potential for use in topical and transdermal preparations.
Article
The manuscript describes a novel small building block, 1,4-piperazinediol (PipzDiol), which has an extended H-bond donor structure compared to piperazine.
Article
A series of novel rhein-piperazine-dithiocarbamate hybrids 3 were efficiently synthesized from rhein through a catalyst-free and one-pot, three-step sequence involving chlorination and N-acylation followed by dithiocarbamate formation. Hybrids 3 were evaluated for their in vitro cytotoxic potency by MTT assay against several human cancer and non-cancer cells. Five of the hybrids were more cytotoxic to human lung cancer cell line A549 than the parent rhein and the reference, cytarabine (CAR). Structure-activity relationship (SAR) analysis indicated that cytotoxicity was significantly enhanced when ester groups were incorporated into the hybrids (3h–j). In particular, hybrid 3h (IC50 = 10.93 μg/mL), containing a long-chain alkyl ester, was the most potent compound toward A549 tumor cells, being 7- and 5-fold more toxic than rhein (IC50 = 77.11 μg/mL) and CAR (IC50 = 49.27 μg/mL), respectively. Additionally, hybrid 3h was less toxic to the corresponding normal human lung fibroblast cell line, WI-38, with a higher selectivity index (SI, WI-38/A549 ≈ 5) than doxorubicin (DOX, SI ≈ 0), CAR (SI ≈ 2) and rhein (SI ≈ 1). Furthermore, hybrid 3h displayed more toxicity against four types of lung cancer cells (A549, Calu-1, PC-9, and H460; IC50 = 10.81–23.78 μg/mL) than against six other types of cancer cells (Huh-7, 786-O, HCT116, Hela, SK-BR-3, and SK-OV-3; IC50 = 23.85–51.98 μg/mL). Further mechanistic studies showed that hybrid 3h induced apoptosis in a concentration-dependent manner in human lung adenocarcinoma cell line PC-9. In vivo safety studies showed that hybrid 3h had no acute toxicity to the major organs of mice and did not lead to blood biochemical index changes. Our results exhibit prominent anti-cancer cell inhibition ability and no obvious systemic toxicity to normal organs, indicating that hybrid 3h has promising potential for further applications in anti-lung cancer drug development.
Article
Different crystal forms of molecular solids have different physical, mechanical and chemical properties and the control of the solid-state form or the preparation of new forms (polymorphs, solvates, salts, cocrystals) is an important and challenging task in many fields ranging from pharmaceuticals, pigments, agrochemicals, and explosives to molecular electronics. The shelf-life, bioavailability and processability of an active pharmaceutical ingredient depends on the crystal form and screening for new solid-state forms including polymorphs, salts and cocrystals is now a routine part of the drug development process. A range of techniques are used to obtain different solid-state forms with high phase purity. Besides the traditional solution crystallization method these include solid-state grinding, extrusion, spray-drying and supercritical fluid methods. Recently, sublimation has emerged as a method that can give excellent crystal form selectivity and often allows the crystallization of solid-state forms that are inaccessible by other methods. In addition, sublimation is a ‘green chemistry’ technique as it produces no waste and does not use solvent. In this highlight article we summarize and discuss the literature on the crystallization of polymorphs, cocrystals and salts from the gas phase showing the tremendous potential of sublimation in solid-state form research.
Article
Full-text available
Olanzapinium monomaleate and dimaleate salts (OLNH+ MA− 1:1 and OLN2H+ 2MA− 1:2) of crystalline and amorphous states were prepared. The crystalline salts exhibit significantly higher solubility (225–550 times) compared to the almost insoluble free base drugOlanzapine (43 mg L−1). The faster dissolution rate of the 1:2 salt compared to 1:1 is explained by their crystal structure analysis.
Article
An infrared method has been developed and used to study the thermal conversion of mefenamic acid from its polymorphic I form to the polymorphic II form. Rates of conversion for the crystal to crystal transition have been measured at temperatures of 150, 155, and 160 °C and subsequently used to calculate the activation energy for the process. The value of 71.6 kcal/mol obtained using the current infrared method is significantly smaller than a previously reported value of 86.4 kcal/mol that was determined by differential scanning calorimetry. In order to explain this difference, additional HPLC experiments were carried out to evaluate the potential loss of analyte due to sublimation when samples of mefenamic acid are maintained at elevated temperatures in the DSC pans prior to measuring the exothermic event indicative of the polymorph I content of a sample. Based on the current study, the rate of anaylte loss due to sublimation is the likely mechanism explaining the larger reported value for the activation energy of the polymorphic conversion of mefenamic acid obtained by DSC.
Article
The polymorphic forms of 2-(2-methyl-3-chloroanilino) nicotinic acid, C13H11ClO2N2, were classified as (I)–(IV) by their i.r. and u.v. spectra, thermograms, and X-ray powder diffraction patterns. Form (I) is monoclinic, space group P21/c, a= 7.625(1), b= 14.201(1), c= 11.672(1)Å, β= 101.65(1)°, and Z= 4. Form (II) is orthorhombic, space group Pca21a= 23.597(6), b= 4.042(1), c= 12.127(3)Å, and Z= 4. Form (III) is triclinic, space group P, a= 13.810(1), b= 3.858(1), c= 10.984(2)Å, α= 94.98(1), β= 94.42(1), γ= 95.57(1)°, and Z= 2. Form (IV) is triclinic, space group P, a= 7.670(1), b= 7.254(1), c= 10.882(1)Å, α= 100.66(1), β= 102.02(1), γ= 86.97(1)°, and Z= 2. Assignments of the i.r. vibration bands to the intra- and inter-molecular hydrogen bonds have been performed taking into consideration the deuterium isotopic effects, the spectral changes observed at low temperature and the spectra of related compounds. Three types of intermolecular hydrogen bonds (A)–(C) were respectively allocated to forms (I), (II), and (III) and (IV). The intramolecular hydrogen bonds found in all forms were identified as NH O. The crystal structures were determined by X-ray analysis. The results confirmed the hydrogen bonds indicated by the i.r. spectral study. Relationships were established between the AH stretching frequencies and the A B distances for the intra- and inter-molecular AH B hydrogen bonds detected in all the forms. Shifts of the longest wavelength band in the u.v. spectra were correlated with the dihedral angles between the pyridine and benzene rings.
Article
New solid forms of the important antimalarialartemisinin have been constructed by cocrystallisation; the difficulty of identifying successful cocrystals formers for artemisinin has been overcome through a mechanochemical screening procedure and the results are compared to predictions based on molecular descriptors obtained from the Cambridge Structural Database (CSD).
Article
Meclofenamic acid, C I4HIICI2NO2, probably the most potent among analgesic fenamates, crystallizes in the triclinic space group P1, with a = 8.569 (5), b = 8.954(8), c -- 9.371 (4) A, ct = 103.0 (2), fl -- 103.5 (2), y = 92.4 (2) ° , Z = 2, D m = 1.43 (4), D c = 1.41 Mg m -3. The structure was solved by direct methods and refined to R = 0.135 for 1062 observed reflections. The anthranilic acid moiety in the molecule is nearly planar and is nearly perpendicular to the 2,6-dichloro-3-methylphenyl group. The molecules, which exist as hydrogen-bonded dimers, have an internal hydrogen bond involving the imino and the carboxyl groups. The methyl group is disordered and occupies two positions with unequal occupancies. The disorder can be satisfactorily explained in terms of the rotational isomerism of the 2,6-dichloro-3-methylphenyl group about the bond which connects it to the anthranilic acid moiety and the observed occupancies on the basis of packing considerations.
Article
Exposure of dilute solutions of N-(2,6-dichloro-m-tolyl)anthranilic acid to visible or UV light results in fairly rapid decomposition with concurrent formation of approximately equimolar amounts of 8-chloro-7-methylcarbazole-1-carboxylic and 8-chloro-5-methylcarbazole-1-carboxylic acids.
Article
A new salt of ziprasidone with high aqueous solubility has been discovered as a result of a salt screening and it has been found to exist in three anhydrous polymorphic forms. One of these forms shows the highest aqueous solubility ever reported for ziprasidone salts in combination with a high kinetic stability, becoming a good candidate for further pharmaceutical development.
Article
Whereas much of organic chemistry has classically dealt with the preparation and study of the properties of individual molecules, an increasingly significant portion of the activity in chemical research involves understanding and utilizing the nature of the interactions between molecules. Two representative areas of this evolution are supramolecular chemistry and molecular recognition. The interactions between molecules are governed by intermolecular forces whose energetic and geometric properties are much less well understood than those of classical chemical bonds between atoms. Among the strongest of these interactions, however, are hydrogen bonds, whose directional properties are better understood on the local level (that is, for a single hydrogen bond) than many other types of non-bonded interactions. Nevertheless, the means by which to characterize, understand, and predict the consequences of many hydrogen bonds among molecules, and the resulting formation of molecular aggregates (on the microscopic scale) or crystals (on the macroscopic scale) has remained largely enigmatic. One of the most promising systematic approaches to resolving this enigma was initially developed by the late M. C. Etter, who applied graph theory to recognize, and then utilize, patterns of hydrogen bonding for the understanding and design of molecular crystals. In working with Etter's original ideas the power and potential utility of this approach on one hand, and on the other, the need to develop and extend the initial Etter formalism was generally recognized. It with that latter purpose that we originally undertook the present review.
Article
Solid-state NMR (SSNMR) is capable of providing detailed structural information about organic and pharmaceutical cocrystals and complexes. SSNMR nondestructively analyzes small amounts of powdered material and generally yields data with higher information content than vibrational spectroscopy and powder X-ray diffraction methods. These advantages can be utilized in the analysis of pharmaceutical cocrystals, which are often initially produced using solvent drop grinding techniques that do not lend themselves to single crystal growth for X-ray diffraction studies. In this work, several molecular complexes and cocrystals are examined to understand the capabilities of the SSNMR techniques, particularly their ability to prove or disprove molecular association and observe structural features such as hydrogen bonding. Dipolar correlation experiments between spin pairs such as 1H−1H, 1H−13C, and 19F−13C are applied to study hydrogen bonding, intermolecular contacts, and spin diffusion to link individual molecules together in a crystal structure and quickly prove molecular association. Analysis of the principal components of chemical shift tensors is also utilized where relevant, as these are more sensitive to structural effects than the isotropic chemical shift alone. In addition, 1H T1 relaxation measurements are also demonstrated as a means to prove phase separation of components. On the basis of these results, a general experimental approach to cocrystal analysis by SSNMR is suggested.