ArticlePDF Available

Abstract and Figures

New pseudopolymorphs of ivermectin (IVM), a potential anti-COVID-19 drug, were prepared. The crystal structure for three pseudopolymorphic crystalline forms of IVM has been determined using single-crystal X-ray crystallographic analysis. The molecular conformation of IVM in crystals has been compared with the conformation of isolated molecules modeled by DFT calculations. In a solvent with relatively small molecules (ethanol), IVM forms monoclinic crystal structure (space group I2), which contains two types of voids. When crystallized from solvents with larger molecules, like -valerolactone (GVL) and methyl tert-butyl ether (MTBE), IVM forms orthorhombic crystal structure (space group P212121). Calculations of the lattice energy indicate that interactions between IVM and solvents play a minor role; the main contribution to energy is made by the interactions between the molecules of IVM itself, which form a framework in the crystal structure. Interactions between IVM and molecules of solvents were evaluated using Hirshfeld surface analysis. Thermal analysis of the new pseudopolymorphs of IVM was performed by differential scanning calorimetry and thermogravimetric analysis.
Content may be subject to copyright.
crystals
Article
Crystal Structures of New Ivermectin Pseudopolymorphs
Kirill Shubin 1, Agris B¯
erzin
,š2and Sergey Belyakov 1,*


Citation: Shubin, K.; B¯
erzin
,š, A.;
Belyakov, S. Crystal Structures of
New Ivermectin Pseudopolymorphs.
Crystals 2021,11, 172. https://
doi.org/10.3390/cryst11020172
Academic Editor: Alexander Pöthig
Received: 14 January 2021
Accepted: 30 January 2021
Published: 9 February 2021
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in
published maps and institutional affil-
iations.
Copyright: © 2021 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
1Latvian Institute of Organic Synthesis, 21 Aizkraukles St., LV-1006 Riga, Latvia; kir101@osi.lv
2Faculty of Chemistry, University of Latvia, 1 Jelgavas St., LV-1004 Riga, Latvia; agris.berzins@lu.lv
*Correspondence: serg@osi.lv; Tel.: +371-67014897
Abstract:
New pseudopolymorphs of ivermectin (IVM), a potential anti-COVID-19 drug, were
prepared. The crystal structure for three pseudopolymorphic crystalline forms of IVM has been
determined using single-crystal X-ray crystallographic analysis. The molecular conformation of
IVM in crystals has been compared with the conformation of isolated molecules modeled by DFT
calculations. In a solvent with relatively small molecules (ethanol), IVM forms monoclinic crystal
structure (space group I2), which contains two types of voids. When crystallized from solvents
with larger molecules, like
γ
-valerolactone (GVL) and methyl tert-butyl ether (MTBE), IVM forms
orthorhombic crystal structure (space group P2
1
2
1
2
1
). Calculations of the lattice energy indicate that
interactions between IVM and solvents play a minor role; the main contribution to energy is made by
the interactions between the molecules of IVM itself, which form a framework in the crystal structure.
Interactions between IVM and molecules of solvents were evaluated using Hirshfeld surface analysis.
Thermal analysis of the new pseudopolymorphs of IVM was performed by differential scanning
calorimetry and thermogravimetric analysis.
Keywords: ivermectin; pseudopolymorph; crystal structure analysis; Hirshfeld surface analysis
1. Introduction
Ivermectin (IVM) is a macrocyclic lactone developed in the 1980s as an antipara-
sitic multitarget drug with nematocidal, acaricidal and insecticidal activities [
1
,
2
]. It is a
semisynthetic substance, which is used as a mixture of two components: major B
1a
(R = Et)
and minor B1b (R = Me), as shown in Figure 1.
Figure 1. Ivermectin as a mixture of two components B1a (R = Et) and B1b (R = Me).
Crystals 2021,11, 172. https://doi.org/10.3390/cryst11020172 https://www.mdpi.com/journal/crystals
Crystals 2021,11, 172 2 of 15
Currently hundreds of millions of people are using IVM for treatment of various
parasitic diseases, including onchocerciasis, lymphatic filariasis, scabies, etc. [
3
]. While at
nanomolar concentrations it is effective mostly against nematodes, at higher concentrations,
multiple new targets were identified [
4
,
5
]. Activity against various types of cancer has been
reported [
6
9
]. IVM is an approved drug for treatment of rosacea due to its antiparasitic
properties complemented by anti-inflammatory activity [
10
12
]. In addition, the activity
of IVM against various viruses [
13
], including COVID-19 [
14
,
15
], is of a special interest.
Several clinical studies are planned or have been started for this indication [16,17].
Nature and properties of a solid form of drugs is important for their production and all
relevant applications [
18
]. Analysis and understanding of the internal molecular arrange-
ments in crystalline materials bear the key to preparation of materials with controllable
and predictable solubility, hygroscopicity and mechanical properties [19,20].
Interestingly, only few crystalline structures of IVM have been reported so far. First,
two single-crystal diffraction data sets on a close analogue avermectin are deposited
in the Cambridge Crystallographic Data Centre with CCDC refcodes BASVAS [
21
] and
YOCYAT [
22
]. IVM was discussed in a context of interaction with the transmembrane
domain of certain receptors using models with low resolution [
23
,
24
]. Additionally, several
crystalline polymorphs were characterized by powder X-ray diffraction [2528]. The only
single crystal data of IVM B
1a
published to date was reported by Seppala et al.: CSD,
Version 5.40, November 2019; CCDC refcode BIFYOF [
29
]. The IVM crystal structure
represents monoclinic modification (space group I2) of IVM as acetone-chloroform solvate.
This form is not satisfactory for drug application due to the presence of a toxic chlorinated
solvent (chloroform).
In this study, we performed crystallization of IVM from several solvents to investigate
whether other crystal structures of this compound can be obtained. It was found that
new pseudopolymorphs, isomorphic to the already reported monoclinic structure, contain
various solvent molecules in the structure cavities. Moreover, in the presence of larger
solvent molecules, orthorhombic structure can be also obtained having bigger cavities able
to accommodate larger solvent molecules. Both types of crystal structures were analyzed
and compared by characterizing intermolecular interactions and molecular conformation,
and the ability of IVM to incorporate different solvents is discussed.
2. Materials and Methods
2.1. Synthesis of Ivermectin Pseudopolymorphs
IVM was obtained from Key Organics,
γ
-valerolactone (GVL) from Carbosynth, UK.
New pseudopolymorphs of IVM were prepared by crystallization of the starting material
from an appropriate solvent. Preparation of IVM as ethanol solvate (
I
) was carried out
by dissolution of IVM (1 g) in EtOH (5 mL) at reflux. Solution was cooled to room
temperature and left for 48 h to effect the crystallization. Crystals of IVM as GVL solvate
(
II
) were prepared by brief heating of IVM (1 g) in GVL (3 mL) up to 120
C, and the
obtained clear solution was left at room temperature for 72 h to effect the crystallization.
Pseudopolymorph of IVM as methyl tert-butyl ether (MTBE) solvate (
III
) was prepared
by dissolution of IVM (1 g) in MTBE (20 mL) at reflux and addition of hexanes (20 mL).
Crystals were obtained at room temperature in 24 h.
2.2. Single Crystal X-ray Diffraction
For compounds
I
(ethanol solvate),
II
(GVL solvate) and
III
(MTBE solvate) diffrac-
tion data were collected at low temperature on a Rigaku, XtaLAB Synergy, Dualflex,
HyPix (Rigaku Corporation, Tokyo, Japan) diffractometer using copper monochromated
Cu-K
α
radiation (
λ
= 1.54184 Å). The crystal structures were solved by direct methods
with the ShelXT (Version 2014/5, Georg-August Universität Göttingen, Germany) [
30
]
structure solution program using intrinsic phasing and refined with the SHELXL (ver-
sion 2016/6, Georg-August Universität Göttingen, Germany) refinement package [
31
].
All calculations were performed with the help of Olex2 software (version Olex2.refine,
Crystals 2021,11, 172 3 of 15
Durham University, UK) [
32
]. The lattice parameters for solvate
I
were determined also at
room temperature; the density of the compound was measured by the flotation method
in ethanol-chloroform system. For calculation of the density, the actual composition of
crystal
I
is 2(IVM)
×
2C
2
H
5
OH
×
1.5H
2
O (with Z= 2), where IVM = 0.8B
1a ×
0.2B
1b
,
where B1a = C48H74O14 and B1b = C47H72O14. Molecular crystals of bulky molecules with
many degrees of freedom, with disordered solvents and not containing heavy atoms (the
heaviest atom in IVM is oxygen) cannot be close to ideal, therefore, R-factors for such
crystal structures are quite high. Table 1lists the main crystal data for these compounds.
Table 1. Crystal data and structure refinement parameters for solvates I,II and III 1.
Parameter I at Low
Temperature
I at Room
Temperature II III
Empirical formula (IVM) ×
C2H5OH ×0.75H2O(IVM) ×
C2H5OH ×0.75H2O2(IVM) ×
0.5C5H8O2
2(IVM) ×
0.5C5H12O
Formula weight 931.85 931.85 1794.59 1787.605
Temperature (K) 173 293 160 160
Crystal size (mm3)0.21 ×0.11 ×0.08 0.17 ×0.09 ×0.06 0.22 ×0.16 ×0.11 0.21 ×0.17 ×0.12
Crystal system Monoclinic Monoclinic Orthorhombic Orthorhombic
Space group I2I2P212121P212121
a(Å) 14.8197(7) 14.8612(9) 16.7127(2) 16.7309(1)
b(Å) 9.1753(5) 9.1973(6) 24.5777(2) 24.5805(2)
c(Å) 39.094(2) 39.201(4) 24.5908(2) 24.5797(2)
β() 94.490(5) 95.04(5) 90.0 90.0
Unit cell volume (Å3)5299.5(5) 5337.4(7) 10100.9(2) 10108.5(1)
Molecular multiplicity 4 4 4 4
Calculated density
(g/cm3)1.168 1.161 1.180 1.175
Measured density
(g/cm3)1.16
Absorption coefficient
(mm1)0.703 0.698 0.702 0.696
F(000) 2023.5 2023.5 3887.2 3875.2
2θmax () 156.0 150.0 155.0 155.0
Reflections collected 29158 5217 71647 75795
Number of
independent reflections
9710 - 20489 20330
Reflections with I>2
σ
(I)
9485 - 18694 19029
Number of refined
parameters 601 - 1143 1143
R-factors (for I>2σ(I)
and for all data) 0.0971, 0.0982 - 0.0981, 0.1051 0.0972, 0.1010
1IVM = 0.8(C48H74 O14)×0.2(C47 H72O14).
Overlay of IVM geometry was done in BIOVIA Discovery Studio 4.5 Visualizer
v4.5.0.15071 (Dassault Systèmes, France), by matching the position of atoms C3, C14
and O26.
2.3. Modeling and Quantitative Analysis of Crystal Structures
To better characterize the differences and similarities between the crystal structures
of IVM, their Hirshfeld surfaces were generated using CrystalExplorer17 (University of
Western Australia, Perth, Australia) [
33
]. They were analyzed by performing the generation
and analysis of Hirshfeld surface 2D fingerprint plots and summarizing the information
about intermolecular interactions [
34
,
35
]. Additionally, for ethanol solvate
I
, pairwise
intermolecular interaction energies for molecules, for which atoms are within 3.8 Å of
the central molecule, were estimated in CrystalExplorer17 at the B3LYP-D2/6-31G(d,p)
level [36] with electronic structure calculations performed in Gaussian09.
Crystals 2021,11, 172 4 of 15
The gas phase geometry optimizations were carried out using Schrödinger software
at the B3LYP/6-311G(d,p) level of theory [37].
2.4. Thermal Analysis
The differential scanning calorimetric/thermogravimetric (DSC/TGA) analysis was
performed on a TGA/DSC2 (Mettler Toledo, Greifensee, Switzerland) apparatus. Open
100
µ
L aluminum pans were used. Heating of the samples from 25 to 600
C was performed
at a heating rate 10
C
·
min
–1
. Samples of 5–10 mg mass were used, and the nitrogen flow
rate was 100 ±10 mL·min–1.
Thermal analysis (TGA and DSC) shows typical data for solid solutions (see files in
the Supplementary Materials). The peaks in the DSC patterns, which correspond to the
melting process, are quite wide: at 165
C for
I
and at 172
C for
II
. Their half-widths are:
17
C for solvate
I
and 8
C for solvate
II
. Thermal analysis data also show that solvate
II
loses the solvent (GVL) at 83
C. In
I
, this process is not observed. This is due to the fact
that solvent in II is not stabilized by strong hydrogen bonds.
3. Results
3.1. Molecular Structure in the Crystal Cell
For investigation of the molecular structure of IVM by means of X-ray diffraction
method, single crystals of
I
have been grown from ethanol solution. It is well known that
the semisynthetic substance of IVM represents a mixture of two compounds—B
1a
and B
1b
in molar ratio 80:20. Thus, a solid solution of these two components as a single crystal
structure was obtained during crystallization. The formation of solid solutions, where
several chemically distinct components occupy the same position in the crystal lattice, is
not such a rare occurrence in organic and inorganic chemistry [
38
]. However, there are not
so many crystal structures of this type in crystallographic databases. This is largely due to
the technical difficulties observed during their crystallographic studies. Despite the fact
that there are two components of IVM in crystals, this paper will further focus on the main
component of IVM, namely B
1a
. Figure 2illustrates an Oak Ridge Thermal-Ellipsoid Plot
(ORTEP) diagram of solvate
I
showing the atom-labeling scheme and thermal displacement
ellipsoids for non-H atoms. For ethyl group (carbon atoms C33 and C34 and hydrogen
atoms H33a, H33b, H34a, H34b and H34c), the value of occupancy g-factor = 0.8 and, for
methyl group (carbon atom C33 and hydrogen atoms H33a, H33b and H33c), the value of
g-factor = 0.2.
The major figure of the molecular structure is the 16-membered macrocycle, which
consists of atoms C3, C10–C20, O21, C22, C4 and C9. In the crystals, the least-squares
plane of this cycle corresponds to the crystallographic plane of (0 3 2). Atoms C15 and C12
have maximal deviations (0.406(5) and –0.382(5) Å, respectively) from this plane. It should
be noted that positive and negative atomic deviations from the plane alternate. Thus, in
crystal
I
, the macrocycle has a “crown” conformation. In the fused bicyclic system, both
cycles are characterized by an envelope conformation: atom C8 deviates on 0.590(5) Å from
the plane of other atoms in the tetrahydrofuran cycle, and atom C9 deviates on 0.627(4) Å
from the plane of other five atoms in the cyclohexene cycle. All the other cycles in the
molecule have a usual chair conformation.
For the comparison of molecular structures in crystal
I
, in vacuo geometry optimiza-
tion of the molecule using density functional theory (DFT) calculations was performed. A
perspective view of the molecule in the free state and in crystals is shown in Figure 3.
Crystals 2021,11, 172 5 of 15
Figure 2.
ORTEP diagram of IVM molecule in crystal
I
showing atomic labels and 50% probability
displacement ellipsoids. Hydrogen atoms are shown as small spheres of arbitrary radii.
Figure 3.
Overlay of IVM molecules present in crystal
I
as determined in the crystal structure (colored
by elements) and after in vacuo geometry optimization (blue).
Crystals 2021,11, 172 6 of 15
As seen from the figure, the molecular conformation in the free state is close to the
one in the crystals. Main differences are associated with the presence of intramolecular
hydrogen bonds of the OH
· · ·
O type, which are formed in vacuo between the hydroxy
groups present in the molecule, whilst, in the crystals, these groups are involved in in-
termolecular hydrogen bonds. The geometrical parameters of these bonds are as follows:
angle O39H39· · · O1 is equal to 113.5, H39· · · O1 length is 2.146 Å; O37H37· · · O23 is
144.9
, H37
· · ·
O23 is 1.865 Å; O57
H57
· · ·
O58 is 111.5
and H57
· · ·
O58 is 2.278 Å. Table
S2 (Supporting Information) lists the values of selected torsion angles; for these angles, a
difference is observed in the free state and in the crystal structure.
As already mentioned, the hydroxy groups form intermolecular hydrogen bonds
in the crystal structure. The hydroxy group O39
H39 forms a moderate hydrogen
bond O39
H39
· · ·
O58 (
1 + x, 1 + y,z) with length 3.048(7) Å (H39
· · ·
O58 = 2.14 Å,
O39
H39
· · ·
O58 = 178
). The hydroxy group O57
H57 forms bifurcated hydrogen
bonds O57
H57
· · ·
O37 (1+x,y,z) (O57
· · ·
O37 = 2.856(6) Å, H57
· · ·
O37 = 2.35 Å,
O57
H57
· · ·
O37 = 121
) and O57
H57
· · ·
O23 (1+x,y,z) (O57
· · ·
O23 = 3.024(6) Å,
H57
· · ·
O23 = 2.25 Å, O57
H57
· · ·
O23 = 158
). Hydrogen bond C61
H61c
· · ·
O39 (
x,
1+y, 1
z), which can be considered a moderate hydrogen bond of the CH
· · ·
O type,
should be also noted. The parameters of this bond are as follows: C61
· · ·
O39 = 3.263(7) Å,
H61c
· · ·
O39 = 2.80 Å and C61
H61c
· · ·
O39 = 110
. By means of these intermolecular
hydrogen bonds, three-dimensional framework containing voids is formed from IVM
molecules. Figure 4shows a projection of the unit cell of crystal
I
along the monoclinic
axis. For geometric modeling of voids in crystals, solvents were removed formally and
volumes of voids were then calculated. As it can be seen, there are two types of the
voids: one of them lies in special positions (on symmetry axes of order 2), its volume is
82 Å3; the second void with volume of 221 Å3corresponds to general positions.
Figure 4. A projection of the unit cell of crystal Ialong the monoclinic axis showing the voids.
It is known that many molecules of macrocyclic compounds are characterized by
the fact that the function of distribution of the electrostatic potential has considerable
extrema [
39
]; this allows the molecules to form supramolecular adducts with ions and polar
molecules. However, for the IVM molecule, the electrostatic potential obtained from the
DFT calculation of the distribution of electron density does not contain significant extrema.
This is also the case for other macrocyclic molecules [
40
]. For this reason, molecules
of IVM can form inclusion compounds with polar molecules due to the formation of
hydrogen bonds.
Crystals 2021,11, 172 7 of 15
Disordered ethanol molecules fill the larger voids in the crystal structure and form
hydrogen bonds of the OH
· · ·
O type with oxygen atom O57. This oxygen atom and carbon,
which is attached to O57 atom, were located from a differential Fourier synthesis and
refined with g= 1.0. However, the methyl group of ethanol is disordered, and two carbon
atoms of this group were located with a differential synthesis and refined with g= 0.5. The
length of this hydrogen bond is 2.832(9) Å. The smaller voids are filled with disordered
water molecules, which form hydrogen bonds with the lengths of 2.67(1)–3.12(1) Å. It
should be noted that this crystal structure is isomorphous to the structure of avermectin
B1a, in which the voids are filled with methanol molecules [21].
The crystal structure of solvate
I
is isomorphous to the structure of IVM-acetone-
chloroform solvate (refcode BIFYOF in the Cambridge Crystallographic Database). The
void that is occupied by chloroform molecules in BIFYOF in crystal
I
is filled with disor-
dered water molecules. That is why the cell volume is by 96.3 Å
3
lower than that of the
BIFYOF structure.
We were interested in testing whether IVM can be crystallized with solvent molecules
exceeding the size of the voids (see Figure 4). It turned out that, upon crystallization of IVM
from GVL, IVM forms molecular crystals
II
of orthorhombic system (space group P2
1
2
1
2
1
)
with two independent molecules of IVM in the asymmetric unit. For both molecules, the
occupancy g-factor of IVM B
1a
is 0.8. IVM molecules form a three-dimensional framework
by means of a system of intermolecular hydrogen bonds. This framework also contains
voids; Figure 5gives a projection of the unit cell along the crystallographic parameter
ashowing their layout. The volume of voids is 552 Å
3
, and they are filled with GVL
molecules. Figure 6shows a content of the asymmetric unit of the crystal. The value of
occupation g-factor for the solvent is 0.5; this means that not all voids are filled with GVL.
It should be noted that, in the crystal structure, there molecules of (R)-enantiomer of GVL
are present despite the fact that a racemic solvent was used for the crystallization. This
suggests that IVM is suitable for the separation of racemic solvents.
The conformation changes of the IVM molecule in solvate
II
are small, but the system
of hydrogen bonds in the crystal structure differs from solvate
I
. The strongest intermolec-
ular hydrogen bonds that form the framework are as follows: O57I–H
· · ·
O43II (
1/2 + x,
3/2
y, 1
z) with length 2.904(5)Å (H
· · ·
O = 2.01 Å, O–H
· · ·
O = 159
); O37II
H
· · ·
O57I
(x,y,z) with length 2.694(5)Å (H
· · ·
O = 1.82 Å, O–H
· · ·
O = 166
) and O57II
H
· · ·
O43I (1
x, 1/2 + y, 1/2
z) with length 2.869(5)Å (H
· · ·
O = 2.11 Å, O–H
· · ·
O = 154
), where I and
II are the designations of the independent molecules. Among the weak hydrogen bonds,
the contact O37I–H
· · ·
O1s (x,y,z) (O
· · ·
O = 2.494(9) Å, H
· · ·
O = 2.99Å, O–H
· · ·
O = 122
)
that binds the IVM molecule with the solvent (GVL) should be distinguished.
In continuation of the study, crystallization of IVM from MTBE solution was also car-
ried out. The size of the molecule of the solvent (MTBE) is relatively large and approaches
the size of GVL. It turns out that IVM crystallizes in orthorhombic system and forms crystal
structure
III
that is isomorphous to crystal structure
II
with g= 0.8 for the B
1a
component
of IVM. The voids of crystal structure III are filled with disordered MTBE molecules.
Conformation of IVM in both isomorphous monoclinic structures (
I
and BIFYOF) and
conformation of each symmetrically unique molecule in both orthorhombic structures (
II
and
III
) are identical, as shown in Figures S1–S3 (Supplementary Materials). Meanwhile,
conformation of IVM present in monoclinic structures and in orthorhombic structures is dif-
ferent (see Figure 7). The conformation of IVM in monoclinic structures is the most differing,
while the conformation for both symmetrically independent molecules in orthorhombic
structures are also notably different.
Crystals 2021,11, 172 8 of 15
Figure 5.
A projection of the unit cell of crystal
II
on the crystallographic plane (100) showing
the voids.
Figure 6.
ORTEP diagram of two independent molecules of IVM in the asymmetric unit of crystal
II
showing thermal ellipsoids with a 50% probability level. For the sake of clarity, hydrogen atoms and
solvent have been omitted.
Crystals 2021,11, 172 9 of 15
Figure 7.
Overlay of IVM molecules present in crystal
I
(colored by elements, representing
conformation in monoclinic structures; overlay of conformation in
I
and BIFYOF is given in
Figure S1, Supplementary Materials) and crystal
II
(blue and red, representing conformation in
orthorhombic structures, overlays of conformations in
II
and
III
are given in Figures S2 and S3,
(Supplementary Materials).
3.2. Qualitative Analysis of Intermolecular Interactions: Hirshfeld Surface and 2D
Fingerprint Plots
Crystal structures of IVM solid forms were also analyzed using Hirshfeld surfaces.
This, however, was complicated by the fact that part of the structures contained disordered
solvent molecules and the fact that all three of them were actually solid solutions. Therefore,
ethanol molecules in one of its potential position was used for IVM ethanol solvate
I
in
this analysis. Hydrogen atoms were added to the solvent molecule in IVM MTBE solvate
III
in Mercury 2020.2.0. Two hydrogen atoms were removed from the acetone molecule in
BIFYOF to obtain molecule with reliable atom arrangement. In all structures of
I
,
II
and
III
,
only the geometry corresponding to B1a (R = Et) was used.
The Hirshfeld surfaces of IVM molecule in the analyzed structures are given in
Figure 8(both sides of the molecule are shown). It can be seen that, as expected, the
closest normalized distances between the atoms involved in intermolecular interactions
are present for the atoms that form conventional and also weak hydrogen bonds (most
obviously, hydroxy groups containing oxygen atoms O37, O39 and O57, carbonyl group
oxygen atom O23 and part of the ether-type oxygen atoms). As could be expected, Hirshfeld
surfaces of both monoclinic structures were quite similar and exhibited more pronounced
difference if compared to the surfaces of IVM in orthorhombic structures. Meanwhile,
Hirshfeld surfaces of both symmetrically independent molecules of the orthorhombic
structures also demonstrated notable differences showing that each of the molecule forms
different intermolecular interactions. These differences can be partially associated with the
different conformation of the molecules in monoclinic and orthorhombic structures and
each symmetrically independent molecule in orthorhombic structures.
Crystals 2021,11, 172 10 of 15
Figure 8.
Hirshfeld surfaces of IVM in the analyzed structures. All surfaces designated by 1 (on the
left) correspond to front view, whereas surfaces designated by 2 (on the right) correspond to the back
view. A and B designates each of symmetrically independent molecules.
Two-dimensional fingerprint plots of these Hirshfeld surfaces are given in Figure 9.
Again, it can be seen that plots obtained from monoclinic structures are highly similar and
there are differences to the plots obtained from the orthorhombic structures, most notably,
in the points representing H
· · ·
O contacts, showing that there are shorter contacts present in
the monoclinic structures. Meanwhile, the differences between plots for both symmetrically
independent molecules of orthorhombic structures are notably less pronounced with the
most notable difference again being the arrangement of points representing H
· · ·
O contacts
and illustrating that both symmetrically independent molecules form hydrogen bonds of
different geometric parameters.
Summary of the quantitative analysis of the contact types present in the Hirshfeld
surfaces is given in Figure 10. Using this representation, it can be seen that there were no
major differences in the relative contribution of different contact types in Hirshfeld surfaces
of IVM structures. In solvate BIFYOF containing acetone and chloroform as solvents, part
of the H
· · ·
H contacts were replaced with H
· · ·
Cl, but the sum of these two contact types
was the same as for purely H
· · ·
H contacts for the remaining surfaces. The most notable
difference among all surfaces seemed to be the larger number of H
· · ·
O contacts present
on the surface of A molecule II.
Crystals 2021,11, 172 11 of 15
Figure 9.
Two-dimensional fingerprint plots of Hirshfeld surfaces of IVM structures showing the
regions associated with the types of interatomic contacts H
· · ·
O, O
· · ·
H and H
· · ·
Cl (present only in
the structure BIFYOF).
Crystals 2021,11, 172 12 of 15
Figure 10.
Percentage contributions of individual contacts to the Hirshfeld surface area for the
analyzed IVM crystal structures.
3.3. Intermolecular Interaction Energy of IVM Ethanol Solvate I
In order to get an additional quantitative picture of the intermolecular interactions
of the crystal structure of solvate
I
, calculation of interaction energies was performed in
CrystalExplorer17 by assessing the electrostatic (E
ele
), polarization (E
pol
), dispersion (E
dis
)
and exchange-repulsion (E
rep
) terms that together form the total interaction energy (E
tot
)
(see Table 2) [
36
,
41
]. However, as water molecules exhibit partial occupancy factors and are
not a critical part of the hydrogen bonding network, they were excluded from the structure
prior to these calculations.
As can be seen from Table 2, the greatest contribution to stabilization of this structure
is made by the interactions between neighboring IVM molecules and, in most cases,
these interactions are dominated by dispersion energy component. Additionally, even
for the hydrogen-bonded molecule pairs the dispersion energy component is dominant
or comparable to the electrostatic components (the highest importance of electrostatic
components is observed in a pair having E
ele
= –43.3 kJ mol
–1
, E
pol
= –13.6 kJ mol
–1
and
E
disp
= –56.2 kJ mol
–1
). This can be easily understood considering the size of the molecule
and the relatively small amount of hydrogen bonds present in this structure (compared to
the weak and dispersion interactions). Notably lower contribution in the stabilization of this
IVM crystal structure is provided by interaction between IVM and ethanol molecules, and
even here electrostatic interactions and dispersion interactions provide similar contribution,
and only in the pair connected by a hydrogen bond the electrostatic components are the
dominant ones.
It follows from the calculations that the energy of the crystal lattice consists mainly
of the energies of interactions between IVM molecules, which form the framework of the
structure. This means that substance
I
is an exemplary compound of the host-guest type. In
this crystal structure, the voids can be filled with molecules of other solvents if the size of
these solvent molecules corresponds to these voids. This is observed in the crystal structure
of BIFYOF, where the structure and symmetry of the IVM framework are preserved.
Crystals 2021,11, 172 13 of 15
Table 2.
The pairwise total interaction energy (E
tot
) and its components (electrostatic (E
ele
), polarization (E
pol
), dispersion
(E
dis
) and exchange-repulsion (E
rep
) energy terms) for the closest molecule pairs (having atoms within 3.8 Å radius from the
central molecule) for IVM-ethanol solvate Icalculated in CrystalExplorer17.
Symmetry R/Å Eele
kJ/mol
Epol
kJ/mol
Edis
kJ/mol
Erep
kJ/mol
Etot
kJ/mol Contact 1
x, y, z 9.18 4.4 1.7 46.8 23.4 32.3 IVM-IVM
x + 1/2, y + 1/2, z + 1/2 10.42 12.5 1.9 87.1 50.1 59.5 IVM-IVM
x + 1/2, y + 1/2, z + 1/2 17.43 3.3 0.4 18.5 10.4 13.4 IVM-IVM
x, y, z 14.82 43.3 13.6 56.2 57 69.6 IVM-IVM
HBS
x, y, z 11.35 8.8 4.6 77.2 62.4 41.3 IVM-IVM
x, y, z 14.60 51.4 8.7 4.1 11.5 IVM-IVM
x, y, z 22.32 0 0.1 6.3 3.2 3.6 IVM-IVM
- 2.65 14.5 340.6 27.9 35.7 IVM-EtOH
x, y, z 17.43 14.3 47.9 29.7 27.7 IVM-IVM
- 7.28 1.7 0.6 13.3 7.7 9.1 IVM-EtOH
- 12.39 0.5 0.1 1.9 0.1 1.2 IVM-EtOH
x, y, z 14.10 0.6 0.8 22 3.8 18 IVM-IVM
- 13.5 35.4 7.8 12.2 37.8 30.5 IVM-EtOH
HB
- 14.9 1.1 0.1 2.9 1.6 2.8 IVM-EtOH
1
IVM-IVM denotes that this is an interaction between two IVM molecules, while IVM-EtOH is an interaction between IVM and ethanol.
HB and HBS indicate that there are one or multiple hydrogen bonds connecting the respective molecules.
4. Conclusions
In summary, three new pseudopolymorphs of IVM were prepared:
I
as the ethanol
solvate,
II
as the GVL solvate and
III
as the MTBE solvate. In all cases, crystallization of
the commercially available IVM provided a solid solution of two components B
1a
and B
1b
with the retaining of the natural 80:20 ratio.
The major feature of the molecular structure is the crown-conformation of the main
macrocyclic ring. Propensity of IVM to crystallize together with solvent molecules is
associated with the molecule being bulky and, as other similar compounds, IVM cannot
pack efficiently without leaving voids in the crystal structure, which are filled with solvent
molecules. In a solvent with relatively small molecules (ethanol), IVM forms monoclinic
crystal structure (space group I2), which contains two types of voids with volumes of
82 and 221 A
3
. The largest void contains one disordered solvent molecule, while the
small one contains disordered water molecules. When crystallized from solvents with
larger molecules (GVL and MTBE), IVM forms orthorhombic crystal structure (space group
P2
1
2
1
2
1
). In case of solvates
II
and
III
, only one type of void is formed with a much bigger
volume: 552 A3.
Conformation of IVM found in crystals is generally retained through all obtained
forms
I
,
II
and
III
and also in solvate BIFYOF, data for which has been previously deposited
in the Cambridge Crystallographic Data Centre. Conformation determined for a molecule,
modeled by DFT calculations, is near to the conformation found in crystals. The Hirshfeld
surface analysis indicated the dominant role of dispersive contacts for H
· · ·
H (80% on
average), O· · · H (10% on average) and H· · · O (10% on average).
The energy of crystal lattice was calculated for model crystal system
I
, which con-
tains IVM molecule and one ethanol molecule. The interaction between IVM molecules
themselves provides the biggest contribution to the crystal energy. By means of these
interactions between IVM molecules, the molecular framework in the crystal structure
is formed.
Thermal analysis shows wide peaks in DSC patterns, typical for solid solutions see
Figures S4 and S5 (Supplementary Materials). The peak of the melting process is observed
at 165
C for pseudopolymorph
I
and at 172
C for
II
. Additionally, the thermal analysis
Crystals 2021,11, 172 14 of 15
data show that, in solvate
II
, the substance loses the solvent (GVL) at 83
C. In crystal
I
,
this process is not observed.
Supplementary Materials:
The following are available online at https://www.mdpi.com/2073-435
2/11/2/172/s1, Table S1: Atomic Cartesian coordinates for IVM B
1a
from DFT, Table S2: Selected
torsion angles for
I
, Figure S1: Overlay of IVM molecules present in
I
and BIFYOF, Figure S2: Overlay
of the first kind of symmetrically unique IVM molecules present in
II
and
III
, Figure S3: Overlay of
the second kind of symmetrically unique IVM molecules present in
II
and
III
, Figure S4: Processed
DSC and TGA data for I, Figure S5: Processed DSC and TGA data for II.
Author Contributions:
K.S. conceived and designed the experiments, conceptualized the work
and prepared the manuscript for publication; A.B. provided crystal structure analysis, reviewed
and edited the manuscript; S.B. provided acquisition of funding and supervision of the research,
conducted the X-ray analysis, reviewed and edited the manuscript. All authors discussed the contents
of the manuscript. All authors have read and agreed to the published version of the manuscript.
Funding:
This research was funded by project “Development of innovative face cosmetics with
controlled release of active ingredients by use of Metal Organic Frameworks or Cocrystals” (ERAF
project number 1.1.1.1/18/A/176).
Conflicts of Interest: The authors declare no conflict of interest.
References
1.
Campbell, W.; Fisher, M.; Stapley, E.; Albers-Schonberg, G.; Jacob, T. Ivermectin: A potent new antiparasitic agent. Science
1983
,
221, 823–828. [CrossRef] [PubMed]
2.
Crump, A. Ivermectin: Enigmatic multifaceted ‘wonder’ drug continues to surprise and exceed expectations. J. Antibiot.
2017
,70,
495–505. [CrossRef]
3. Ashour, D.S. Ivermectin: From theory to clinical application. Int. J. Antimicrob. Agents 2019,54, 134–142. [CrossRef]
4. Laing, R.; Gillan, V.; Devaney, E. Ivermectin—Old Drug, New Tricks? Trends Parasitol. 2017,33, 463–472. [CrossRef]
5.
Perez-Garcia, L.A.; Mejias-Carpio, I.E.; Delgado-Noguera, L.A.; Manzanarez-Motezuma, J.P.; Escalona-Rodriguez, M.A.; Sordillo,
E.M.; Mogollon-Rodriguez, E.A.; Hernandez-Pereira, C.E.; Marquez-Colmenarez, M.C.; Paniz-Mondolfi, A.E. Ivermectin:
Repurposing a multipurpose drug for Venezuela’s humanitarian crisis. Int. J. Antimicrob. Agents
2020
,56, 106037. [CrossRef]
[PubMed]
6.
Juarez, M.; Schcolnik-Cabrera, A.; Dueñas-Gonzalez, A. The multitargeted drug ivermectin: From an antiparasitic agent to a
repositioned cancer drug. Am. J. Cancer Res. 2018,8, 317–331.
7.
Laudisi, F.; Marônek, M.; Di Grazia, A.; Monteleone, G.; Stolfi, C. Repositioning of Anthelmintic Drugs for the Treatment of
Cancers of the Digestive System. IJMS 2020,21, 4957. [CrossRef]
8.
Mudassar, F.; Shen, H.; O’Neill, G.; Hau, E. Targeting tumor hypoxia and mitochondrial metabolism with anti-parasitic drugs to
improve radiation response in high-grade gliomas. J. Exp. Clin. Cancer Res. 2020,39, 208. [CrossRef] [PubMed]
9.
Tang, M.; Hu, X.; Wang, Y.; Yao, X.; Zhang, W.; Yu, C.; Cheng, F.; Li, J.; Fang, Q. Ivermectin, a potential anticancer drug derived
from an antiparasitic drug. Pharmacol. Res. 2020, 105207. [CrossRef]
10.
Sahni, D.R.; Feldman, S.R.; Taylor, S.L. Ivermectin 1% (CD5024) for the treatment of rosacea. Expert Opin. Pharmacother.
2018
,19,
511–516. [CrossRef]
11.
McGregor, S.P.; Alinia, H.; Snyder, A.; Tuchayi, S.M.; Fleischer, A.; Feldman, S.R. A Review of the Current Modalities for the
Treatment of Papulopustular Rosacea. Dermatol. Clin. 2018,36, 135–150. [CrossRef]
12.
Feaster, B.; Cline, A.; Feldman, S.R.; Taylor, S. Clinical effectiveness of novel rosacea therapies. Curr. Opin. Pharmacol.
2019
,46,
14–18. [CrossRef]
13.
Heidary, F.; Gharebaghi, R. Ivermectin: A systematic review from antiviral effects to COVID-19 complementary regimen.
J. Antibiot. 2020,73, 593–602. [CrossRef] [PubMed]
14.
Jans, D.A.; Wagstaff, K.M. Ivermectin as a Broad-Spectrum Host-Directed Antiviral: The Real Deal? Cells
2020
,9, 2100. [CrossRef]
15.
Sen Gupta, P.S.; Rana, M.K. Ivermectin, Famotidine, and Doxycycline: A Suggested Combinatorial Therapeutic for the Treatment
of COVID-19. Acs Pharmacol. Transl. Sci. 2020,3, 1037–1038. [CrossRef]
16.
Rajter, J.C.; Sherman, M.S.; Fatteh, N.; Vogel, F.; Sacks, J.; Rajter, J.-J. ICON (Ivermectin in COvid Nineteen) study: Use of
Ivermectin is Associated with Lower Mortality in Hospitalized Patients with COVID19; Public and Global Health. 2020. Available
online: https://www.medrxiv.org/content/10.1101/2020.06.06.20124461v2 (accessed on 2 February 2021).
17.
Gupta, D.; Sahoo, A.K.; Singh, A. Ivermectin: Potential candidate for the treatment of Covid 19. Braz. J. Infect. Dis.
2020
,24,
369–371. [CrossRef]
18.
Healy, A.M.; Worku, Z.A.; Kumar, D.; Madi, A.M. Pharmaceutical solvates, hydrates and amorphous forms: A special emphasis
on cocrystals. Adv. Drug Deliv. Rev. 2017,117, 25–46. [CrossRef] [PubMed]
Crystals 2021,11, 172 15 of 15
19.
Pindelska, E.; Sokal, A.; Kolodziejski, W. Pharmaceutical cocrystals, salts and polymorphs: Advanced characterization techniques.
Adv. Drug Deliv. Rev. 2017,117, 111–146. [CrossRef]
20.
Calvo, N.L.; Maggio, R.M.; Kaufman, T.S. Chemometrics-assisted solid-state characterization of pharmaceutically relevant
materials. Polymorphic substances. J. Pharm. Biomed. Anal. 2018,147, 518–537. [CrossRef] [PubMed]
21.
Springer, J.P.; Arison, B.H.; Hirshfield, J.M.; Hoogsteen, K. The absolute stereochemistry and conformation of avermectin B2a
aglycone and avermectin B1a. J. Am. Chem. Soc. 1981,103, 4221–4224. [CrossRef]
22.
Keates, A.C. CCDC 1904192: Experimental Crystal Structure Determination 2019. Available online: https://www.ccdc.cam.ac.
uk/structures/Search?Ccdcid=1904192&DatabaseToSearch=Published (accessed on 3 February 2021).
23.
Jelínkova, I.; Vávra, V.; Jindrichova, M.; Obsil, T.; Zemkova, H.W.; Zemkova, H.; Stojilkovic, S.S. Identification of P2X4 receptor
transmembrane residues contributing to channel gating and interaction with ivermectin. Pflug. Arch. Eur. J. Physiol.
2008
,456,
939–950. [CrossRef]
24.
Huang, X.; Chen, H.; Shaffer, P.L. Crystal Structures of Human GlyR
α
3 Bound to Ivermectin. Structure
2017
,25, 945–950.e2.
[CrossRef] [PubMed]
25.
Grobler, M.L.J. Genome-Wide Analysis of Wolbachia-Host Interactions. Master’s Thesis, North-West University, Potchefstroom,
South Africa, 2000.
26.
Rolim, L.A.; dos Santos, F.C.M.; Chaves, L.L.; Gonçalves, M.L.C.M.; Freitas-Neto, J.L.; da Silva do Nascimento, A.L.; Soares-
Sobrinho, J.L.; de Albuquerque, M.M.; do Carmo Alves de Lima, M.; Rolim-Neto, P.J. Preformulation study of ivermectin raw
material. J. Anal. Calorim. 2015,120, 807–816. [CrossRef]
27.
Starkloff, W.J.; Bucalá, V.; Palma, S.D.; Gonzalez Vidal, N.L. Design and
in vitro
characterization of ivermectin nanocrystals liquid
formulation based on a top–down approach. Pharm. Dev. Technol. 2017,22, 809–817. [CrossRef] [PubMed]
28.
Lu, M.; Xiong, D.; Sun, W.; Yu, T.; Hu, Z.; Ding, J.; Cai, Y.; Yang, S.; Pan, B. Sustained release ivermectin-loaded solid lipid
dispersion for subcutaneous delivery: In vitro and in vivo evaluation. Drug Deliv. 2017,24, 622–631. [CrossRef]
29.
Seppala, E.; Kolehmainen, E.; Osmialowski, B.; Gawinecki, R. CCDC 187337: Experimental Crystal Structure Determination 2005.
Available online: https://www.ccdc.cam.ac.uk/structures/Search?Ccdcid=187337&DatabaseToSearch=Published (accessed on 3
February 2021).
30.
Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. A Found. Adv.
2015
,71, 3–8.
[CrossRef]
31. Sheldrick, G.M. A short history of SHELX.Acta Cryst. A Found. Cryst. 2008,64, 112–122. [CrossRef]
32.
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. Cryst. 2009,42, 339–341. [CrossRef]
33.
Turner, M.J.; McKinnon, J.J.; Wolff, S.K.; Grimwood, D.J.; Spackman, P.R.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer17; The
University of Western Australia: Perth, WA, Australia, 2017.
34.
McKinnon, J.J.; Jayatilaka, D.; Spackman, M.A. Towards quantitative analysis of intermolecular interactions with Hirshfeld
surfaces. Chem. Commun. 2007, 3814. [CrossRef]
35. Spackman, M.A.; Jayatilaka, D. Hirshfeld surface analysis. CrystEngComm 2009,11, 19–32. [CrossRef]
36.
Mackenzie, C.F.; Spackman, P.R.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer model energies and energy frameworks:
Extension to metal coordination compounds, organic salts, solvates and open-shell systems. IUCrJ
2017
,4, 575–587. [CrossRef]
[PubMed]
37.
Bochevarov, A.D.; Harder, E.; Hughes, T.F.; Greenwood, J.R.; Braden, D.A.; Philipp, D.M.; Rinaldo, D.; Halls, M.D.; Zhang, J.;
Friesner, R.A. Jaguar: A high-performance quantum chemistry software program with strengths in life and materials sciences. Int.
J. Quantum Chem. 2013,113, 2110–2142. [CrossRef]
38.
Sarš
¯
uns, K.; B
¯
erzi
n
,
š, A.; Rekis, T. Solid Solutions in the Xanthone–Thioxanthone Binary System: How Well Are Similar Molecules
Discriminated in the Solid State? Cryst. Growth Des. 2020,20, 7997–8004. [CrossRef]
39.
Marczenko, K.M.; Mercier, H.P.A.; Schrobilgen, G.J. A Stable Crown Ether Complex with a Noble-Gas Compound. Angew. Chem.
Int. Ed. 2018,57, 12448–12452. [CrossRef] [PubMed]
40.
Popov, I.; Chen, T.-H.; Belyakov, S.; Daugulis, O.; Wheeler, S.E.; Miljani´c, O.Š. Macrocycle Embrace: Encapsulation of Fluoroarenes
by m-Phenylene Ethynylene Host. Chem. A Eur. J. 2015,21, 2750–2754. [CrossRef] [PubMed]
41.
Turner, M.J.; Grabowsky, S.; Jayatilaka, D.; Spackman, M.A. Accurate and Efficient Model Energies for Exploring Intermolecular
Interactions in Molecular Crystals. J. Phys. Chem. Lett. 2014,5, 4249–4255. [CrossRef] [PubMed]
... The results of the polymorphism screening therefore suggest that, under the conditions in our laboratory, new polymorphs of ivermectin could not be prepared. This inability to obtain new polymorphs can possibly be explained by the results presented by Shubin et al. [56], in a recent paper where they prepared solvates of ivermectin with ethanol, γ-valerolactone (GVL) and methyl tert-butyl ether (MTBE). Their single crystal X-ray diffraction (SXRD) data showed that the structural conformation of ivermectin was largely retained in all three solvates. ...
... Their single crystal X-ray diffraction (SXRD) data showed that the structural conformation of ivermectin was largely retained in all three solvates. This general conformation of ivermectin was also consistent with that of an ivermectin acetone-CHCl 3 solvate, BIFYOF, which had previously been deposited on the Cambridge Crystallographic Data Centre (CCDC), and with the structure generated from density function theory (DFT) calculations [56]. Shubin et al. further found that that the different solvents mainly determined the space groups, and crystal structures, of their solvates, while the crystal frameworks were stabilized by interactions between neighboring ivermectin molecules. ...
... It should be noted here that Span® 60 is not crystalline, as is evident by its diffuse halo in Fig. 2. This suggests that the peak positions were mostly influenced by the crystal packing of ivermectin, an observation that was consistent with the previously mentioned findings by Shubin et al. regarding the dominant role of neighboring ivermectin molecules in stabilizing the crystal structure [56]. It is possible that the intermolecular interactions between ivermectin and Span® 60, in both the co-crystal candidates and physical mixtures, were not only very similar, but also very limited. ...
Article
Full-text available
Despite being discovered over five decades ago, little is still known about ivermectin. Ivermectin has several physico-chemical properties that can result in it having poor bioavailability. In this study, polymorphic and co-crystal screening was used to see if such solid-state modifications can improve the oil solubility of ivermectin. Span® 60, a lipophilic non-ionic surfactant, was chosen as co-former. The rationale behind attempting to improve oil solubility was to use ivermectin in future topical and transdermal preparations to treat a range of skin conditions like scabies and head lice. Physical mixtures were also prepared in the same molar ratios as the co-crystal candidates, to serve as controls. Solid-state characterization was performed using X-ray powder diffraction (XRPD), Fourier-transform infrared spectroscopy (FTIR), differential scanning calorimetry (DSC) and thermogravimetric analysis (TGA). The FTIR spectra of the co-crystal candidates showed the presence of Span® 60’s alkyl chain peaks, which were absent in the spectra of the physical mixtures. Due to the absence of single-crystal X-ray data, co-crystal formation could not be confirmed, and therefore these co-crystal candidates were referred to as co-processed crystalline solids. Following characterization, the solid-state forms, physical mixtures and ivermectin raw material were dissolved in natural penetration enhancers, i.e., avocado oil (AVO) and evening primrose oil (EPO). The co-processed solids showed increased oil solubility by up to 169% compared to ivermectin raw material. The results suggest that co-processing of ivermectin with Span® 60 can be used to increase its oil solubility and can be useful in the development of oil-based drug formulations. Graphical Abstract
... All IVR derivatives share similar structural featuresa rigid sixteenmembered lactone ring, a spiroketal comprising two six-membered rings, and a cyclohexene ring fused to a five-membered cyclic ether with two or three hydroxyl groups (Fig. 4) [42,43]. Although the introduction of an additional hydroxyl group at position C4 does not significantly influence the lactone ring and spiroketal conformation, the rearrangement does modify the cyclohexene ring. ...
Article
Full-text available
The Debye scattering equation (DSE) [Debye (1915). Ann. Phys.351, 809–823] is widely used for analyzing total scattering data of nanocrystalline materials in reciprocal space. In its modified form (MDSE) [Cervellino et al. (2010). J. Appl. Cryst.43, 1543–1547], it includes contributions from uncorrelated thermal agitation terms and, for defective crystalline nanoparticles (NPs), average site-occupancy factors (s.o.f.’s). The s.o.f.’s were introduced heuristically and no theoretical demonstration was provided. This paper presents in detail such a demonstration, corrects a glitch present in the original MDSE, and discusses the s.o.f.’s physical significance. Three new MDSE expressions are given that refer to distinct defective NP ensembles characterized by: (i) vacant sites with uncorrelated constant site-occupancy probability; (ii) vacant sites with a fixed number of randomly distributed atoms; (iii) self-excluding (disordered) positional sites. For all these cases, beneficial aspects and shortcomings of introducing s.o.f.’s as free refinable parameters are demonstrated. The theoretical analysis is supported by numerical simulations performed by comparing the corrected MDSE profiles and the ones based on atomistic modeling of a large number of NPs, satisfying the structural conditions described in (i)–(iii).
Article
Full-text available
Predictions of the structures of stoichiometric, fractional, or nonstoichiometric hydrates of organic molecular crystals are immensely challenging due to the extensive search space of different water contents, host molecular placements throughout the crystal, and internal molecular conformations. However, the dry frameworks of these hydrates, especially for nonstoichiometric or isostructural dehydrates, can often be predicted from a standard anhydrous crystal structure prediction (CSP) protocol. Inspired by developments in the field of drug binding, we introduce an efficient data-driven and topologically aware approach for predicting organic molecular crystal hydrate structures through a mapping of water positions within the crystal structure. The method does not require a priori specification of water content and can, therefore, predict stoichiometric, fractional, and nonstoichiometric hydrate structures. This approach, which we term a mapping approach for crystal hydrates (MACH), establishes a set of rules for systematic determination of favorable positions for water insertion within predicted or experimental crystal structures based on considerations of the chemical features of local environments and void regions. The proposed approach is tested on hydrates of three pharmaceutically relevant compounds that exhibit diverse crystal packing motifs and void environments characteristic of hydrate structures. Overall, we show that our mapping approach introduces an advance in the efficient performance of hydrate CSP through generation of stable hydrate stoichiometries at low cost and should be considered an integral component for CSP workflows.
Article
Full-text available
The binary system of xanthone–thioxanthone has been explored, showing that two solid solutions (formed based on xanthone and thioxanthone parent structures, respectively) exist for this system. One of the solid solutions shows miscibility of both molecules in a large composition range (>0–80 mol % of xanthone). The structure of thioxanthone has been redetermined to reveal a special case of nonmerohedral twinning in the crystals. Such a twinning feature has apparently been the reason for incorrect crystal structure determination previously. A structure of thioxanthone:xanthone (75:25 mol %) solid solution is also presented. Several similar molecules to the title compounds have been found in the Cambridge Structural Database and shown to crystallize in structures isostructural to that of thioxanthone. The different packing of pure xanthone is thus an exception among the explored compounds.
Article
Full-text available
High-grade gliomas (HGGs), including glioblastoma and diffuse intrinsic pontine glioma, are amongst the most fatal brain tumors. These tumors are associated with a dismal prognosis with a median survival of less than 15 months. Radiotherapy has been the mainstay of treatment of HGGs for decades; however, pronounced radioresistance is the major obstacle towards the successful radiotherapy treatment. Herein, tumor hypoxia is identified as a significant contributor to the radioresistance of HGGs as oxygenation is critical for the effectiveness of radiotherapy. Hypoxia plays a fundamental role in the aggressive and resistant phenotype of all solid tumors, including HGGs, by upregulating hypoxia-inducible factors (HIFs) which stimulate vital enzymes responsible for cancer survival under hypoxic stress. Since current attempts to target tumor hypoxia focus on reducing oxygen demand of tumor cells by decreasing oxygen consumption rate (OCR), an attractive strategy to achieve this is by inhibiting mitochondrial oxidative phosphorylation, as it could decrease OCR, and increase oxygenation, and could therefore improve the radiation response in HGGs. This approach would also help in eradicating the radioresistant glioma stem cells (GSCs) as these predominantly rely on mitochondrial metabolism for survival. Here, we highlight the potential for repurposing anti-parasitic drugs to abolish tumor hypoxia and induce apoptosis of GSCs. Current literature provides compelling evidence that these drugs (atovaquone, ivermectin, proguanil, mefloquine, and quinacrine) could be effective against cancers by mechanisms including inhibition of mitochondrial metabolism and tumor hypoxia and inducing DNA damage. Therefore, combining these drugs with radiotherapy could potentially enhance the radiosensitivity of HGGs. The reported efficacy of these agents against glioblastomas and their ability to penetrate the blood-brain barrier provides further support towards promising results and clinical translation of these agents for HGGs treatment.
Article
Full-text available
At times, combination therapy has proven to be very effective. While no cure is available to date, herein we put forward with rationale and supporting evidence that if adminis-trated simultaneously, a combination of FDA-approved drugs comprising ivermectin, famotidine, and doxycycline may provide robust chemoprophylaxis effective against COVID-19.
Article
Full-text available
The small molecule macrocyclic lactone ivermectin, approved by the US Food and Drug Administration for parasitic infections, has received renewed attention in the last eight years due to its apparent exciting potential as an antiviral. It was identified in a high-throughput chemical screen as inhibiting recognition of the nuclear localizing Human Immunodeficiency Virus-1 (HIV-1) integrase protein by the host heterodimeric importin (IMP) α/β1 complex, and has since been shown to bind directly to IMPα to induce conformational changes that prevent its normal function in mediating nuclear import of key viral and host proteins. Excitingly, cell culture experiments show robust antiviral action towards HIV-1, dengue virus (DENV), Zika virus, West Nile virus, Venezuelan equine encephalitis virus, Chikungunya virus, Pseudorabies virus, adenovirus, and SARS-CoV-2 (COVID-19). Phase III human clinical trials have been completed for DENV, with >50 trials currently in progress worldwide for SARS-CoV-2. This mini-review discusses the case for ivermectin as a host-directed broad-spectrum antiviral agent for a range of viruses, including SARS-CoV-2.
Article
Full-text available
Tumors of the digestive system, when combined together, account for more new cases and deaths per year than tumors arising in any other system of the body and their incidence continues to increase. Despite major efforts aimed at discovering and validating novel and effective drugs against these malignancies, the process of developing such drugs remains lengthy and costly, with high attrition rates. Drug repositioning (also known as drug repurposing), that is, the process of finding new uses for approved drugs, has been gaining popularity in oncological drug development as it provides the opportunity to expedite promising anti-cancer agents into clinical trials. Among the drugs considered for repurposing in oncology, compounds belonging to some classes of anthelmintics—a group of agents acting against infections caused by parasitic worms (helminths) that colonize the mammalian intestine—have shown pronounced anti-tumor activities and attracted particular attention due to their ability to target key oncogenic signal transduction pathways. In this review, we summarize and discuss the available experimental and clinical evidence about the use of anthelmintic drugs for the treatment of cancers of the digestive system.
Article
Ivermectin is a macrolide antiparasitic drug with a 16-membered ring that is widely used for the treatment of many parasitic diseases such as river blindness, elephantiasis and scabies. Satoshi ōmura and William C. Campbell won the 2015 Nobel Prize in Physiology or Medicine for the discovery of the excellent efficacy of ivermectin against parasitic diseases. Recently, ivermectin has been reported to inhibit the proliferation of several tumor cells by regulating multiple signaling pathways. This suggests that ivermectin may be an anticancer drug with great potential. Here, we reviewed the related mechanisms by which ivermectin inhibited the development of different cancers and promoted programmed cell death and discussed the prospects for the clinical application of ivermectin as an anticancer drug for neoplasm therapy.
Article
Ivermectin proposes many potentials effects to treat a range of diseases, with its antimicrobial, antiviral, and anti-cancer properties as a wonder drug. It is highly effective against many microorganisms including some viruses. In this comprehensive systematic review, antiviral effects of ivermectin are summarized including in vitro and in vivo studies over the past 50 years. Several studies reported antiviral effects of ivermectin on RNA viruses such as Zika, dengue, yellow fever, West Nile, Hendra, Newcastle, Venezuelan equine encephalitis, chikungunya, Semliki Forest, Sindbis, Avian influenza A, Porcine Reproductive and Respiratory Syndrome, Human immunodeficiency virus type 1, and severe acute respiratory syndrome coronavirus 2. Furthermore, there are some studies showing antiviral effects of ivermectin against DNA viruses such as Equine herpes type 1, BK polyomavirus, pseudorabies, porcine circovirus 2, and bovine herpesvirus 1. Ivermectin plays a role in several biological mechanisms, therefore it could serve as a potential candidate in the treatment of a wide range of viruses including COVID-19 as well as other types of positive-sense single-stranded RNA viruses. In vivo studies of animal models revealed a broad range of antiviral effects of ivermectin, however, clinical trials are necessary to appraise the potential efficacy of ivermectin in clinical setting.
Article
For decades, Ivermectin (IVM) has been recognized as a robust antiparasitic drug with excellent tolerance and safety profiles. Historically it has been used as the drug of choice for onchocerciasis and lymphatic filariasis global elimination programs. IVM is also a standard treatment against intestinal helminths and ectoparasites given its action as an oral insecticide. The current humanitarian crisis in Venezuela is a regional public health threat that requires immediate action. Venezuela's public health system has now crumbled due to a 70% shortage of medicines in public hospitals, low vaccination campaigns, and the massive exodus of medical personnel. Herein we discuss the repurposing of IVM to attenuate the burden imposed by the most prevalent neglected tropical diseases (NTDs) in Venezuela including soil-transmitted helminths, ectoparasites and, possibly, vector-borne diseases such as malaria. Additionally, novel experimental evidence has shown that IVM is active and efficacious against Chagas disease, Leishmaniases, arboviruses, and SARS-CoV-2 in vitro. In crisis-hit Venezuela, all of the aforementioned infectious diseases are public health emergencies that have been long ignored and that also require immediate attention. IVM's versatile nature could serve as a powerful tool to tackle the multiple overlapping endemic and emergent diseases that affect Venezuela today. The repurposing of this multipurpose drug would be without a doubt a timely therapeutical approach to help mitigate the tremendous burden of NTDs nationwide.
Article
Rosacea is a common inflammatory skin disease that is difficult to manage because of the unknown etiology and due to its variable manifestations. These facts and the few new available treatment options make it difficult to select a really effective treatment. This review aims to assess the efficacy and safety of novel treatment options for rosacea. The topical alpha adrenergic agonist oxymetazoline reduces rosacea-related erythema. Topical ivermectin improves lesion count, inflammation, and maintenance of remission of rosacea compared to topical metronidazole. Procedural therapies including pulsed dye laser, radiofrequency, and dual frequency ultrasound are promising as both monotherapies or in combination. Although there are several effective treatment modalities for rosacea management, treatments options should be tailored for the specific clinical scenario.