Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Bioorganic & Medicinal Chemistry Letters 27 (2017) 5100–5108

Contents lists available at ScienceDirect

Bioorganic & Medicinal Chemistry Letters


journal homepage: www.elsevier.com/locate/bmcl

Digest

Improvement in aqueous solubility achieved via small molecular


changes
Michael A. Walker
Dart Neuroscience, 12278 Scripps Summit Dr., San Diego, CA 92131, USA

a r t i c l e i n f o a b s t r a c t

Article history: Overcoming poor solubility is a significant issue in drug discovery. The most common solution is to
Received 4 May 2017 replace carbon atoms with polar heteroatoms such as N and O or by attaching a solubilizing appendage.
Revised 14 September 2017 This approach can lead to other issues such as poor activity and PK or the increased risk for toxicity.
Accepted 17 September 2017
However, there are more subtle structural changes which can be employed that lead to an increase in sol-
Available online 19 October 2017
ubility. These include, excising hydrophobic groups which do not efficiently contribute to binding, mod-
ifying stereo- and regiochemistry, increasing or decreasing the degree of unsaturation or adding small
Keywords:
hydrophobic groups such as fluorine or methyl.
Aqueous solubility
Drug properties
Ó 2017 Elsevier Ltd. All rights reserved.
General Solubility Equation (GSE)
Hydrophobicity
Physicochemical properties

Contents

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5100
Standard approach for improving solubility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5101
Improving solubility through small changes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5101
Removal of hydrophobic functionality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5102
Modification of molecular geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5102
Reduction or increase in the degree of unsaturation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5104
Carbon-carbon based degree of unsaturation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5105
Carbon-heteroatom based degree of unsaturation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5105
Addition of small hydrophobic groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5106
Addition of fluorine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5106
Addition of methyl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5107
Summary and conclusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5108
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5108

Introduction tein. Hydrophobicity can play a positive role as a non-specific driv-


ing force for the partitioning of the drug into the binding site by
Discovering a new drug is very challenging since the molecular raising its free energy in water.1–3 Additional binding affinity
properties which drive optimal biology and pharmacology are comes from specific interactions of the compound with the target.
often in opposition to those which lead to favorable physicochem- However, from a physicochemical property point of view,
ical properties such as solubility. For example, drug binding can be hydrophobicity is undesirable since it leads to decreased solubility
viewed as a distribution equilibrium between the surrounding as well as other issues. Thus, finding the appropriate balance can
aqueous solution and the hydrophobic cavity of the targeted pro- be difficult.
Thermodynamic solubility of a solid is defined as its concentra-
E-mail address: mwalker@dartneuroscience.com
tion in solution when it is at equilibrium with a stable crystalline

https://doi.org/10.1016/j.bmcl.2017.09.041
0960-894X/Ó 2017 Elsevier Ltd. All rights reserved.
M.A. Walker / Bioorganic & Medicinal Chemistry Letters 27 (2017) 5100–5108 5101

form. Therefore, thermodynamic solubility is not only affected by Standard approach for improving solubility
hydrophobicity but also by the crystal lattice stability of the com-
pound in the solid state. Kinetic solubility is defined as the concen- Since hydrophobicity is the most often encountered factor driv-
tration at which a compound separates from solution. It is ing poor solubility, increasing polarity is usually the path most
generally assumed that kinetic solubility is driven mainly by medicinal chemists follow for overcoming it. Thus, as a simplified
hydrophobicity. In the experience of the author, solubility is some- generalization, the standard approach, as defined here, involves
times defined in the literature according to the experimental one of two types of structural modifications. In the first method
method used to measure it and might not be an accurate reflection (method a) hydrophobic carbon atoms are replaced by polar atoms,
of the intrinsic thermodynamic or kinetic solubility. There is a large most commonly nitrogen or oxygen. This can be accomplished by
variety of methods employed which differ in the form of the ana- replacing phenyl rings with heterocycles such as pyridine, or ali-
lyte (crystalline solid, amorphous-solid, concentrated stock solu- phatic CH2’s with oxygen or nitrogen resulting in an ether or
tion, or dried semi-solid from stock-solution) the aqueous phase, amine. Numerous other substitutions are possible. In most cases
temperature, and equilibration time. When reporting solubility, it these modifications alter the core-template of the molecule and
is highly recommended that it be accompanied by a description as such can have an impact on drug binding. In order to avoid this
of the physical state of the solute, the make-up of the aqueous issue, an alternative approach (method b) is to attach a polar or
phase (pH and buffer concentration), temperature and equilibra- ionizable solubilizing appendage to the molecule. In a majority of
tion time. cases this is a basic nitrogen containing group but it can also be
The General Solubility Equation (GSE, Eq. (1))4,5 equates the sol- a carboxylic acid or neutral-polar moiety. The appendage is
ubility of a solid organic compound to its log P (hydrophobicity) attached to a site on the molecule which does not bind to the target
and melting point (solid state stability). While the equation was and which, theoretically should not interfere with affinity. This
developed as a predictive tool, it is useful for evaluating the rela- method can be applied as a late stage fix following the optimiza-
tive contribution of hydrophobicity and solid state stability to sol- tion of activity.
ubility. Poorly soluble compounds, where the log P is larger than While these tactics are effective, the changes to the physical
the melting point term, are said to display solvation limited solu- properties are significant enough that one risks unfavorably mod-
bility. Alternatively, when the melting point term dominates, the ifying in vivo behavior. A small sample of the physicochemical
compounds are characterized as exhibiting solid state limited sol- properties altered by employing the standard approach are shown
ubility. A recent analysis of poorly soluble drugs found that most in Table 1. The favorable decrease in log P is accompanied by an
suffer from solvation limited solubility.6 This is to be expected unfavorable increase in molecular weight, hydrogen bond donor
given the positive correlation between affinity and hydrophobicity or acceptor count, acidity or basicity, and polar surface area. There
mentioned above. Nonetheless, there are cases where high crys- are upper bounds to these properties, above which, drug absorp-
talline stability is the underlying factor. This can be the case for tion and blood brain barrier permeability are significantly ham-
compounds which bind their targets via strong hydrogen bonds. pered.7,8 In addition, carboxylic acids can be glucuronidated to
The hydrogen bonding functionality can lead to strong intermolec- form reactive intermediates while basic amines are recognition
ular interactions in the crystalline state. elements for P-glycoprotein efflux,9 inhibition of hERG and
phospholipidosis.10
Log SðMÞ ¼ 0:5  log P  0:01ðMPð CÞ  25Þ ð1Þ
The GSE also offers insight into the expected change in solubil- Improving solubility through small changes
ity associated with changes in log P or melting point. Assuming a
constant melting point, a decrease in log P of 1 log unit will result Subtle structural modifications which reduce log P or melting
in a 10-fold increase in solubility. Likewise, holding log P constant point can also lead to improved solubility and are available as an
and reducing the melting point by 100 °C will result in the same alternative to the methods described in the previous section. As
change in solubility. However, it is usually not possible to indepen- shown in Table 2, these modifications include removing hydropho-
dently change either log P or melting point without affecting the bic groups, modifying geometry by altering the position of func-
other. Moreover, they often change in opposite directions. This is tional groups, changing stereochemistry, reducing or increasing
usually because functionality which improves one property will the degree of unsaturation, or adding substituents such as methyl
often have the opposite effect on the other. For example, groups or fluorine. These changes increase solubility by reducing molecu-
such as H-bond donors and acceptors, which lead to lower log P, lar size, H-bond acceptor/donor count or by altering the shape of
raise the possibility of intermolecular H-bonding in the solid state. the molecule. Basicity/acidity, polar surface area and other param-
Other modifications which lead to an increase in compound polar- eters are not negatively impacted.
ity can have a similar effect due to more favorable intermolecular As illustrated in the following examples, alterations to these
interactions in the solid state. In contrast, structural modifications properties can have a significant effect on solubility. Since a num-
which weaken crystalline stability can sometimes increase ber of publications covering this topic have appeared in the litera-
hydrophobicity. ture in the last few years11–13 the current review is focused on
more recent examples and should be considered a snapshot of
Table 1 the most recent literature. Also, the primary goal here is to demon-
Molecular properties affected by the standard approach for improving solubility. strate solubility SAR, therefore the compounds were ‘cherry picked’
a b b
Property Method a Method b
Log P decrease decrease Table 2
mol wt NA increase Small molecular changes which can improve solubility.
#HBD/#HBA increase increase
Basicity/acidity increase increase Structural change Effect on molecular weight
PSA increase increase Remove unnecessary hydrophobic groups reduce
a Find more soluble stereo/region-isomer maintain
Abbreviations: mol wt; molecular weight, #HBD; hydrogen bond donor count,
Change degree of unsaturation reduce/increase by 2
#HBA; hydrogen bond acceptor count, PSA; polar surface area
b Add F or Me increase by 18 or 15
Defined in the text.
5102 M.A. Walker / Bioorganic & Medicinal Chemistry Letters 27 (2017) 5100–5108

from larger data-sets and may not be representative of the optimal Table 4
progression of the chemotype discussed in the reference. In cases Log D, pKa and kinetic solubility (pH 7.4) of the 2,3-anti and 2,3-syn isomers 3a-e and
4a-e.
where previously reviewed topics are presented, the goal is to aug-
ment, through additional examples the original findings. The OH OH
examples are ordered according to the net effect the modification O O
has on the molecular weight of the parent (Table 2) as a measure N +/- N +/-
of the relative size of the change. Where possible the conditions N 3
N 3
used in the solubility assay are reported. Activity data is included 2 2
O O
in order to monitor the effect the structural change has on target N N
engagement. Finally, for the sake of brevity, the principal ideas of H H N
O N O
this review are illustrated using a minimal number of examples
and compounds.

Removal of hydrophobic functionality 3a-e (2,3-anti) 4a-e (2,3-syn) OPh


OPh

Improving solubility via molecular size reduction is very attrac-


tive since it can potentially help avoid other issues later in the Property 3a–e 4a–e
development process.14 Although hydrophobicity drives potency, a
IC50 nM 1–5 16–450
compounds sometimes possess lipophilic groups or domains which Log D 4.3–4.5 3.5–3.9
make only a moderate contribution to the overall affinity of the pKa.(N15) 6.08–6.18 7.07–7.16
b
molecule. The presence of such groups can be identified by evalu- sol7.4 mg/mL 0.5–1.2 51–58
ating the binding affinity of the molecule (pKi or pIC50) relative to a
Inhibition of T. cruzi growth in mouse fibroblasts.
its hydrophobicity (log P). The Lipophilic Ligand Efficiency (LLE) b
DMSO stock solution with pH 7.4 PBS buffer/1% DMSO, 18 h at room
index (Eq. (2)) is one method for doing this.15 Compounds which temperature.
bind efficiently will have a pKi (or pIC50) which is significantly
higher than log P. A net difference of 5 is generally considered
the minimal benchmark for efficient binding. Thus, a higher LLE, kinase and low solubility (sol. < 0.5 mg/mL). Both of these issues
indicates that affinity is driven more by specific interactions with were solved by truncating the left-hand portion of the molecule
the target (pKi), and less by hydrophobic binding to the target leading to compound 2 which is less hydrophobic (clog P = 2.03
(log P). It is possible that a highly potent molecule could exhibit versus 0.74 for 1 and 2 respectively) and more soluble (sol. =
a low LLE if the log P is high due to being excessively large. One 10.6 mg/mL). In addition, the corresponding increase in activity
remedy for this is to look for portions of the molecule which can (IC50 = 0.012 mM) combined with the lower clog P resulted in much
be paired down without affecting binding affinity. more efficient binding (LLE = 4.32 versus 7.18 for 1 and 2
LLE ¼ pK i ðor pIC 50 Þ  log P ð2Þ respectively).

A recent example from the literature illustrates this approach


nicely. Compound 1 (Table 3) was identified in a cell-based screen Modification of molecular geometry
as a Hedgehog pathway inhibitor.16 Although it was potent (IC50 =
0.45 mM) it suffered from both poor selectivity against p38 MAP Stereochemistry and regiochemistry can affect solubility by
altering the shape and geometry of a molecule in the solid state
and in solution, leading to changes in melting point and/or log P.
Table 3 For example, according to Carnelly’s rule, isomers with higher sym-
Solubility, clog P, IC50 and LLE of 1 and 2. metry exhibit higher melting points.17 Therefore, solubility can
sometimes be improved by breaking the symmetry of the
O
molecule. This approach has been recently reviewed.18 Molecular
R N O geometry can also strengthen or weaken inter- and intra-molecu-
H
N O lar H-bonds in the solid state which influence crystalline stability.
H Stereochemistry also affects solution state behavior (i.e. Log P)
N
as illustrated by a recent example (Table 4).19 The antiparasitic
compounds 3 and 4 contain 3-stereocenters resulting in 8 possible
O N ∗ diastereomers (3a–e, 4a–e). Examination of the corresponding log
N ∗ D, pKa, and solubility at pH 7.4 revealed an interesting trend
N NH
H related to the relative stereochemistry at positions 2 and 3. The
2 four 2,3-anti diastereomers 3a–e where found to be more
F hydrophobic (log D = 4.3–4.5) and less soluble at pH 7.4 (sol7.4 =
0.5–1.2 mg/mL) than the corresponding 2,3-syn isomers, 4a-e (log
1
D = 3.5–3.9, sol7.4 = 51–58 mg/mL). In addition, the basicity of the
a
IC50/bEC50 mM 0.04/2.3 0.01/0.01 benzylic amine was also dependent on the relative 2,3-geometry.
p38a IC50 mM 0.03 2.7
Conformational analysis indicated that the benzylic amine is favor-
mw 448 308
c
sol. mg/mL <0.5 10.6 ably positioned to form an intramolecular H-bond with the nicoti-
IC50 mM 0.45 0.012 noyl amide NH. This type of intramolecular interaction is known to
clog P 2.03 0.74 increase hydrophobicity.20 Intramolecular H-bonding can increase
LLE 4.32 7.18
hydrophobicity in two ways. First, it stabilizes the unprotonated
a
Inhibition of Luciferase reporter gene in NIH-3T3 cells. form of the C2-benzylic-methylamine thereby reducing the degree
b
Inhibition of Hedgehog dependent differentiation of C3H10T1/2 cells. of ionization in solution and second it blocks the amine from H-
c
Method not reported. bonding to water. In the case of the syn isomers, steric strain
M.A. Walker / Bioorganic & Medicinal Chemistry Letters 27 (2017) 5100–5108 5103

Table 5 Table 7
Solubility, log P and clog Kb of compounds 5–7. Solubility of compounds 12–15.

R
N Het
N N CO2 H
CN
N
O O
N CN

a
compd pEC50 (S1P1) sol. mg/mL
N N N O O N
Property
Het ∗
O N N ∗
∗ ∗ ∗
O N
12 9.7 46
5 6 7

a
sol.7.4 mg/mL 63 2.5 5.3 ∗ N
Log P 2.6 3.3 3.1
N N
13 8.5 187
clog Kb 2.9/2.6 1.8/0.8 1.6/1.6

a
pH 7.4, 24 h at 25 °C ∗ O

S N
14 8.6 160

between the C3-methyl and C2-benzylic–methylamine disfavors ∗ N
the intramolecular H-bonded conformation reducing its propensity
N N
15 8.9 256
to exist in solution. As a result the amine is available to interact ∗
with water leading to higher solubility compared to the anti ∗ S
isomers. a
Method not provided.
In a separate publication it is revealed that the more soluble anti
diastereomers are less active than the syn analogues.21 However,
this example illustrates two points. First, intramolecular H-bond- probably due to the higher dipole moment associated with the
ing often leads to an increase in hydrophobicity. Second, the H- 1,3,4- compared to the 1,2,4-substitution pattern.23, In addition,
bond can be disrupted without having to remove the donor or the authors suggest that the lower log P is also the result of the
acceptor groups. It is feasible that if one had a molecule where stronger H-bonding potential of the nitrogen atoms of the 1,3,4-
the donor and/or acceptor group directly engaged the target that oxadiazole. The calculated H-bond basicities (clog Kb)24 of the
both activity and solubility could be improved by hindering nitrogen atoms of the parent oxadiazoles are shown in the table.
intramolecular H-bonding through conformational control. It is assumed that higher H-bond acceptor basicity leads to a more
The regioisomers of multi-heteroatom containing heterocycles favorable H-bonding interaction with water.
often exhibit large differences in physicochemical properties such A broader analysis of oxadiazole compounds using matched
as log P, melting point and solubility despite the fact that they molecular pair analysis (MMPA) was conducted by researchers at
are structurally very similar. In a recent paper22 the properties of AstraZeneca.25 They found that in accordance with the results in
a series of oxadiazole regioisomers were examined (Table 5). For Table 5, 1,3,4-oxadiazoles tend to be less hydrophobic and more
this set of compounds the 1,3,4-oxidiazole 5 was found to be more soluble than 1,2,4-oxadiazoles. The difference in log D between
soluble than its 1,2,4-isomeric counterparts 6 and 7. The solubility the two regioisomers is on the order of 1 log unit. However, despite
difference is likely the result of the lower log P of 5 (log P = 2.6) rel- the lower log D, some exceptions to the general trend of higher sol-
ative to 6 and 7 (log P = 3.3 and 3.1, respectively). This suggests ubility for the 1,3,4-regioisomers were noted. A separate example
that 5 is better solvated by water compared to 6 and 7. This is from the literature was found which demonstrates this behavior
in a series of soluble epoxide hydrolase inhibitors and suggests
one possible reason for the outliers.26 As seen in Table 6, while
Table 6
Solubility, clog P and melting point of compounds 8–11. compounds 8 and 9 are predicted to have lower clog P than 10
and 11, they are slightly less soluble at pH 7.4. The reason for this
Het is that the lower clog P for the 1,3,4-isomers is offset by corre-
spondingly higher melting points. It can hypothesized that the lar-
O
ger dipole moment and stronger H-bond basicity of 8 and 9 lead to
N stronger intermolecular interactions in the solid state. In this way
H
the improved solution behavior imparted by the 1,3,4-oxadiazole
R is accompanied by an unfavorable increase in crystalline stability
leading to lower solubility. This illustrates that changes in log P
a
and melting point are often negatively correlated with one
compd R clog P mp °C sol.7.4 mg/mL
Het ∗ another. The activity of the compounds against soluble epoxide
∗ hydrolase was measured using a fluorescence based assay. Deriva-
tives 8–11 inhibited the enzyme by 46–64% at 1 nM. Thus, the
N N
8 H 3.54 228 5.89
structure of the oxadiazole core appeared to have had little impact
9 ∗ F 3.37 235 5.00
O
on activity.
∗ Interestingly, there is evidence from the literature that the
O N
10 H 4.41 152 6.12
regioisomeric effect seen for the oxadiazoles might apply to other
11 ∗ F 4.62 173 5.23
N 5-membered ring heterocycles. Shown in Table 7, during the
∗ course of optimizing a series of S1P1 agonists the oxadiazole
a
pH 7.4, 25 ± 1.0 °C. regioisomers, 12 and 13 and thiadiazoles 14 and 15 were synthe-
5104 M.A. Walker / Bioorganic & Medicinal Chemistry Letters 27 (2017) 5100–5108

Table 8 Table 10
Solubility difference between pyridazine (16) and pyrimidine (17) containing Solubility, clog P and melting point of hydroxyl-quinoxaline (21),- quinazoline (22)
inhibitors. and -cinnoline (23) heterocycles.

Cl OH OH OH
N N N
HN SO2 N
N N het N
O N
S 22 23
21
F3 C SO2 HN
sol. mg/mL 6.6 1.3 0.6
mp °C 102–103 149–150 185–186
N

O P31,32 progressively change as the nitrogens draw closer together


in the ring. Therefore, the higher solubility of 16 is consistent with
a
the higher dipole moment, pKHBX, pKa, and lower log P of the pyri-
Compd sol.7.4 mg/mL
∗ ∗ dazine compared to the corresponding pyrimidine of 17.
Het
As with the 5-membered ring heterocycles, the solubility of 6-
16 N N 35 membered ring heterocycles is determined by the net balance of
∗ ∗ solution (log P) and solid state (melting point) properties. The com-
pounds shown in Table 10 illustrate this point.34–37 Solubility is
17 N 3
driven more by crystalline stability than hydrophobicity for this
∗ ∗
N series. Compound 21 has a much lower melting point (mp = 102–
103 °C) than 22 and 23 (mp = 149–150 and 185–186 °C, respec-
a
Dried from DMSO sample, pH 7.4, 24 h, 25 °C.33 tively) which correlates well with their observed solubilities.
The observed trend in melting point for 21, 22 and 23 is oppo-
site to what one would expect from the monocyclic heterocycles in
sized. Similar to the oxadiazoles the 1,3,4-thiadiazole regioisomer Table 9. This can be rationalized based on the putative nitrogen H-
15 (sol. = 256 mg/mL) was more soluble than the corresponding bonding basicities of the isomers and the presence of an H-bond
1,2,4-thiadiazole analogue 14 (sol. = 160 mg/mL).27 While the donor group (OH). According to Etter’s rule38 which states that
1,2,4-oxadiazole is nearly 10-fold more potent than the 1,3,4-iso- ‘‘all good proton donors and acceptors are used in hydrogen bond-
mer, the thiadiazole regioisomers are similar in activity. A more ing”, one can assume that compounds 21–23 will be stabilized by
thorough investigation of the literature is needed in order to see the formation of such bonds in the solid state. If true, the inter-
how general this phenomenon is for other 5-membered ring molecular H-bonding in 23 would be expected to be stronger than
heterocycles. 21 and 22 due to the higher H-bond basicity. This would lead to
Six membered nitrogen-containing heterocycles also display higher crystalline stability and melting point. The correlation of
interesting solubility SAR related to their regiochemistry. For solubility with H-bond basicity would also account for the differ-
example in the course of optimizing inhibitors of Bcl-2/Bcl-xL both ence in solubility between 21 and 22.
the pyridazine and pyrimidine analogues 16 and 17 were synthe-
sized (Table 8).28 Although 16 was found to be approximately
Reduction or increase in the degree of unsaturation
10-fold less active than 17 (data not shown), it exhibited a 10-fold
improvement in solubility at pH 7.4 (sol. = 35 and 3 mg/mL for 16
The textbook definition for the total number of p-bonds plus
and 17 respectively). This is relatively large compared to their
rings in a molecule is ‘‘degree of unsaturation”.39 In this section
structural similarity and the fact that pyrimidine and pyridazine
both sp2-hybridized atoms and ring-count are considered. A dou-
make up only a small portion of the structures.
ble bond or ring are each equivalent to one degree of unsaturation
Examination of the physical properties of the parent heterocy-
and reflect the loss of 2 hydrogen atoms. The addition or subtrac-
cles offers some interesting insights. The three possible arrange-
tion of 2 hydrogen atoms is perhaps one of the smallest changes to
ments of 2-nitrogen atom containing, 6-membered ring
the constitutional make-up of a molecule one can make. This
heterocycles, are shown in Table 9. In this series the dipole
increases or decreases the total number of rings and/or p-bonds
moment,29 H-bond basicity (pKHBX),30 pKa, melting point and log
contained in the molecule. This seemingly small change in struc-
ture can result in large improvements in solubility.
A number of retrospective analyses have appeared in the last
Table 9
Dipole moment, pKHBX, pKa, melting point and log P of compounds pyrazine (18), few years evaluating the effect of unsaturation (sp2-hybridization
pyrimidine (19) and pyridazine (20). and aromaticity) on the solubility of drug like molecules. The gen-
eral consensus is that solubility is reduced as the relative-number
Compd N N N
N of unsaturated atoms increases.40 For example, Lovering et al.
N found that solubility was positively correlated with the fraction
N
of sp3-hybridized carbon atoms (Fsp3).41 In addition, Ritchie and
19 20
18 Mcdonald reported an increase in hydrophobicity and decrease in
a
solubility as the number or aromatic rings increased.42 Although
DM 0.0D 2.4D 3.9D
pKHBX 0.92 1.07 1.65
these studies and others show a clear correlation between solubil-
b
pKa 0.7 1.3 2.3 ity and the level of unsaturation one should be cautious in not
mp °C 52 20 -8 over-interpreting the results to suggest that this reflects the funda-
log P -0.26 -0.44 -0.72 mental properties of sp3 and sp2 carbon atoms. The effect of unsat-
a
Dipole moment (DM) measured in dioxane. uration on solubility is context dependent, especially in the case of
b
pKa of the conjugate acid hydrocarbon-aromatic versus heteroaromatic systems. The follow-
M.A. Walker / Bioorganic & Medicinal Chemistry Letters 27 (2017) 5100–5108 5105

ing sections describe situations under which unsaturation


decreases solubility and others where it increases it.

Carbon-carbon based degree of unsaturation


The correlation of physical properties of hydrocarbons or hydro- H H
carbon-moieties accompanying alterations in the degree of unsat-
uration can be appreciated by examining the series of C8- H H H H
HO HO
hydrocarbons shown in Table 11. Removal of 2 hydrogen atoms
from n-octane (25) increases the degree of unsaturation by 1 and 29 30
yields an alkene or a ring (26 and 27). Both of these modifications sol. = 0.026 μg/mL (30 oC) sol. = 0.009 μg/mL (35 oC)
lead to a reduction in lipophilicity (log P) which is sufficient to
yield an improvement in solubility. The lower lipophilicity of 1-
octene (log P = 4.57) compared to octane (log P = 4.93) is likely
the result of increased polarity. On the other hand one would not In summary, for non-aromatic hydrocarbon based molecules or
expect cyclooctane to be more polar than n-octane but it displays moieties, increasing the degree of unsaturation through sp2-carbon
a lower log P of 4.07 nonetheless. The effect of cyclization is to atoms or cyclization can lead to a reduction in lipophilicity. On the
reduce the size of the molecule, and thus the solvation cavity rel- other hand lowering the degree of unsaturation via deannulation
ative to the acyclic form, which translates to a lower entropic pen- can lead to improved solubility via a lowering the melting point.
alty. Therefore, cyclooctane is more soluble than n-octane because With some caveats these principles can also be extended to aro-
it is smaller in size. matic hydrocarbons when used as a substitute for Fsp3. Carbo-
The degree of unsaturation plays a role in melting point as well. cyclic aromatic compounds tend to be more soluble than their
Although the C8-alkanes in Table 11 are liquids at room tempera- saturated counterparts due to a more favorable enthalpy of solva-
ture the effects of sp2/sp3-hybridization and cyclization on melting tion.49 For example, benzene has a clog P = 2.14 while cyclohexa-
point can be extrapolated to compounds which are solids. In the triene, cyclohexene and cyclohexane have clog P = 2.45, 2.87 and
case of isolated carbon-based sp2 centers saturation usually leads 3.35 respectively. On the other hand unlike isolated sp2 functional-
to an increase in melting point. In the case of aromatic systems sat- ity, aromatic compounds tend to have higher melting points due to
uration will lead to a reduction in melting point. In addition to sat- better crystal packing. The increase in log P associated with satu-
uration, melting point is usually positively correlated with the rating a carbocyclic-aryl might be offset by a reduction in melting
cyclization. The melting point of n-octane (mp = 57 °C) versus point.
cyclooctane (mp = 13 °C) illustrates this point. Moreover, melting This is interesting given the conclusions of the empirical studies
point continues to increase as the number of rings increases such mentioned above which suggest that increasing Fsp3 yields
that bicyclo[3.2.1]octane (28) has a melting point of 135 °C.43 improved solubility. However, since the effect of unsaturation on
Therefore, in certain cases solubility might be improved by dean- solubility is context dependent, this would suggest that the results
nulation in order to lower the melting point even though this of these studies are biased by the make-up of the compounds in
might increase log P. the data sets. This seems plausible since, heterocycles make up a
large portion of the compounds encountered in medicinal chem-
istry. As seen in the next section, the observations in these studies
most likely reflect the properties of the heteroatoms contained
within these molecules. However, since phenyl is one of the most
commonly encountered rings found in drug-like molecules50 the
28 modifications described in the current section should prove useful.

Although the C8-compounds in Table 11 are simple mole-


Carbon-heteroatom based degree of unsaturation
cules they reflect the fundamental properties of hydrocarbon
As mentioned above, there is an important distinction between
systems and the same effects can be seen in more drug-like
all-carbon unsaturated functionalities versus those containing het-
molecules. For example, cholesterol (29) and cholestanol (30)
eroatoms. The change in polarity of the heteroatom accompanying
are naturally occurring compounds differing from one another
the change in the degree of unsaturation (ie. sp2 to sp3) needs to be
by a single site of saturation. Similar to the octane example
considered. For example, as shown in Table 13, the partially satu-
in Table 11 the saturated derivative cholestanol is significantly
rated heterocycle tetrahydroisoquinoline 33 is less hydrophobic
less soluble in water (sol. = 0.009 mg/mL at 35 °C)44 than choles-
than the fully aromatic isoquinoline, 34 (log P = 1.71 and 2.09
terol (sol. = 0.026 mg/mL at 30 °C).45 Likewise, in Table 12 the
respectively). This is the opposite trend observed for the all-carbon
relative solubilities of the steroid pair testosterone 31 (monohy-
templates where saturation results in increased hydrophobicity.
drate, sol. = 23 mg/mL)46,47 and stanolone 32 (sol. = 5 mg/mL)48
follow the same trend.
Table 12
Solubility, clog P and melting point of testosterone (31) and stanolone (32).
Table 11
Degree of unsaturation, solubility, clog P and melting point (mp °C) of n-octane (25), OH OH
1-octene (26) and cyclooctane (27).
H H

25 26 H H H H
O O
H
27 31
32
deg. unsat. 0 1 1
sol. mg/mL 1 3 7 sol. mg/mL 23 5
log P 4.93 4.57 4.07 log P 3.32 3.55
mp °C 57 102 13 mp °C 155 181
5106 M.A. Walker / Bioorganic & Medicinal Chemistry Letters 27 (2017) 5100–5108

The increase in hydrophilicity for tetrahydroisoquinoline can be activity in the kinesin spindle assay was also improved by
correlated to its higher H-bond acceptor basicity (pKBHX) = 2.04) deannulation.
compared to isoquinoline (pKBHX) = 1.94).27 In addition the more
basic sp3-nitrogen of tetrahydroisoquinoline (pKa = 9.3 at 37 °C)51 Addition of small hydrophobic groups
will be more highly protonated than the sp2-hybridized nitrogen
of isoquinoline (pKa = 5.4 at 25 °C) at neutral pH further adding The final area of structural modification considered is the addi-
to the increased solubility.52 For heteroatom containing templates tion of small hydrophobic substituents such as methyl or fluorine.
the effect of unsaturation on melting point is variable. In some The effect of replacing hydrogen by either of these groups has been
cases, like 33, the fully aromatic analogue is higher melting com- evaluated using matched molecular pair analysis (MMPA). The
pared to its saturated counterpart, while for others, for example results from three separate literature reports are summarized in
quinolone (mp = 15 °C) and tetrahydroquinoline (mp = 14 °C), Table 15.55–57 The mean change in solubility (Dlog S) associated
the opposite is true. with replacing hydrogen with F or CH3 is given while the percent-
A recent example demonstrates the large effect that partial sat- age of matched pairs where an increase was observed is provided
uration of a heterocyclic ring can have on aqueous solubility. In a in parenthesis. As will be discussed below, the context of the sub-
report describing the optimization of a series of androgen receptor stitution is important. The MMPA studies in Table 15 were carried
down regulators, hERG-inhibition and poor solubility arose as crit- out using different collections of compounds, and the observations
ical issues.53 Lead compound 35 was found to be highly hydropho- reflect the structural make-up of these sets. Nonetheless, the stud-
bic (log D7.4 = 3.2) and poorly soluble (sol. = 3 mg/mL, crystalline ies yield interesting observations that provide hints to some under-
solid, 0.1 M PBS, 24 h). Partial saturation of the triazolopyridazine lying trends.
ring yields 36 resulting in a significant reduction in hydrophobicity
(log D7.4 = 2.5) and improved solubility (sol. = 299 mg/mL, crys- Addition of fluorine
talline solid, 0.1 M PBS, 24 h). Although compound 36 (pIC50 = As shown in the Table 15, while the mean effect of replacing H
5.75) was less active than 35 (pIC50 = 6.4) it showed much lower by F results in a reduction in solubility (Dlog S = 0.22, 0.45 and
hERG-inhibition as well as higher plasma free fraction and lower 0.10 respectively) there appears to be a significant percentage of
intrinsic clearance in human hepatocytes. The overall profile was instances where the reverse is observed. Leach and Zhang reported
such that compound 36 was advanced as a candidate (AZD3514) an increase in solubility for 22% and 34% of the compounds respec-
for clinical trials. tively, while Gleeson observed it for 9%. Taken together this sug-

O O
N N
N Me N Me

N O N O

N N
N N N N
N N
CF3 35 CF3 36
logD7.4 = 3.2 logD7.4 = 2.5
sol7.4 = 3 μg/mL sol 7.4 = 299 μg/mL

The kinesin spindle inhibitors 37 and 38 (Table 14) demonstrate gests that there might be a special chemical property of fluorine
the effect of ring opening on solubility.54 The carbazole based which imparts improved solubility when deployed in certain struc-
derivative 37 contains 4-rings and is very insoluble. Its solubility tural contexts.
in water (pH 7.4 PBS) was less than 1 mg/mL leading the research- Much earlier than this Hansch made some pioneering observa-
ers to use a solution of 50% EtOH in pH 7.4 PBS in their measure- tions concerning fluorine’s effect on hydrophobicity (log P or log D)
ments. Under these conditions the solubility of 37 was 0.5 mg/
mL. It also exhibits a very high melting point (mp > 300 °C) sug-
Table 14
gesting that the poor solubility is the result of strong crystal pack-
Solubility, HPLC retention time and melting point of kinesin spindle inhibitors 37 and
ing. Deannulation yields compound 38 which has a lower melting 38.
point (mp = 190 °C) and is 3-fold more soluble. Interestingly, ring-
opening appears to increase hydrophobicity as indicated by the O O
NH NH
longer retention time via reverse phase HPLC (logP not reported).
This might be expected given the C8-examples in Table 11. Finally,

N CF3 HN
H
Table 13 CF3
Solubility, clog P and melting point of tetrahydroisoquinoline (33) and isoquinoline
37
(34). 38
a
IC50 mM 0.18 0.045
b
sol. mg/mL 0.5 1.8
NH N c
HPLC-RT min. 21.7 24.2
mp °C >300 190
33 34
a
Inhibition of microtubule-activated KSP ATPase assay.
sol. mg/mL NA 5.66 b
50% EtOH/7.4 PBS, 48 h, 25 °C.
pKBHX 2.04 1.94 c
HPLC analysis was carried out on a Cosmosil 5C18-ARII column (4.6  250 mm),
log P 1.71 2.09
and the material eluted by a linear MeCN gradient (70:30–30:70 over 40 min) in
mp °C 30 25
0.1% TFA; flow rate of 1 mL/min.
M.A. Walker / Bioorganic & Medicinal Chemistry Letters 27 (2017) 5100–5108 5107

Table 15 F R F O F R
Matched molecular pair analysis (MMPA) for replacement of H by F, and CH3. O R
OH
R Dlog S (% comps with increased sol.) F F O F
a
Leach et al. Zhang et al. a
Gleeson et al. O
OH
F 0.22 (34)[Ar] 0.45 (22) 0.10 (9)[Ar]
CH3 0.21 (33)[Ar] 0.50 (26) 0.11 (11)
a Fig. 1. Examples of fluorinated functionality leading to reduced log D.
[Ar]; substitution of an aromatic H by R.

shedding light onto structural contexts where its dual nature is A much more profound example comes from a series of com-
manifested. It was found that fluorine behaves differently when pounds designed as potential agents for the treatment of amy-
attached to an aryl ring versus an alkyl group. A summary of these otrophic lateral sclerosis (ALS).64 The lead compound 41 was
results are shown in Table 16.58–61 Using X = H as a reference found to be potent, exhibiting good oral bioavailability and brain
(entry 1), fluorine yields a lower log P when attached to an alkyl penetration. Fluorination was examined as one of two approaches
chain (log P = 2.95 and 3.68 for X = F and H respectively) versus to further optimize the solubility and stability in microsomes and
phenyl (log P = 2.27 and 2.13 for X = F and H respectively). The plasma. Introduction of F onto both of the phenyl rings (42) yielded
order of magnitude in the reduction in log P is 0.73 log units no change in solubility (from DMSO stock solution, PBS buffer, 45
which could potentially translate into a 10-fold increase in solu- min.–16 h). In contrast bis-fluorination of the ethyl-spacer group
bility. Interestingly, this behavior is shared, to a certain degree, by provided a 10-fold increase in solubility for 43 and 44. In this
chlorine and bromine (entries 3 and 4) where log P is either case, despite the improved solubility, the clog P predicted an
reduced (Cl) or left unchanged (Br) when attached to an alkyl-site. increase in hydrophobicity. This could represent the inability of
Therefore, the unusual behavior noted for fluorine might be better the calculation protocol to accurately predict the log P of this
classified as a halogen atom property. chemotype. Fluorination of the alkyl portion of the molecule also
While this would explain Zhang’s results (Table 15) it does not resulted in increased stability in plasma and human microsomes.
account for Leach and Gleeson studies where aryl substitution was
examined. Interestingly, an earlier report by Böhm62 examined the O O O O
effect of fluorine substitution on log D of compounds from the N N F N N F
Roche compound collection. While the mean effect was an increase
O O
in log D there were certain structural configurations associated 41 42
with a decrease. As shown in Fig. 1, in addition to alkyl-based clog P = 3.42 clog P = 3.70
groups, a number of aryl based sub-structures were found. It is sol. = 10 μg/mL sol. = 12 μg/mL
not certain whether these structural motifs were present in the
Leach and Gleeson data sets but it reinforces the idea that the O O O O
F F F F F F F F
behavior of fluorine is context dependent. N N F N N F
The solubility enhancing effects of fluorine have been demon-
O O
strated in a number of drug discovery programs. As an example,
43 44
fluorination of the BRaf inhibitor 39 (sol. = 20 mg/mL, from solid, clog P = 4.61 clog P = 4.89
pH 6.5 PBS, room temperature, 24 h) provides a modest improve- sol. > 200 μg/mL sol. = 111 μg/mL
ment in solubility for 40 (sol. = 60 mg/mL, from solid, pH 6.5 PBS,
room temperature, 24 h).63 Both compounds exhibit favorable
inhibition of B-Raf (IC50 = 3.8 and 0.6 nM for 39 and 40, respec-
tively) Addition of methyl
The MMPA analysis in Table 15 also indicates that methylation
F F can result in an increase in solubility. One mechanism by which
O O methylation can increase solubility is by blocking the formation
O S O S F
N O N O of intra- and intermolecular hydrogen bonds. Intramolecular H-
H H
NH Cl NH Cl bonding hampers interactions with water while intermolecular
N N H-bonds lead to higher crystalline stability. In the final example,
N N 39 N N 40 the Mycobacterium tuberculosis inhibitor 45 (Table 17) features a
H H
clog P = 2.19 clog P = 1.61 primary amide which is positioned in close proximity to two H-
sol. = 20 μg/mL sol. = 60 μg/mL bond acceptors, the adjacent ether-oxygen and the distal pyridine
nitrogen.65 As demonstrated in Table 4, it can be hypothesized that
intramolecular H-bonding leads to higher lipophilicity (log P not

Table 16
Hansch’s analysis of the effect of halogenation at alkyl or aryl positions.

entry X
X X

Log P Dlog P Log P Dlog P


1 H 3.68 – 2.13 –
2 F 2.95 -0.73 2.27 0.14
3 Cl 3.55 -0.13 2.84 0.71
4 Br 3.72 0.04 2.99 0.86
5108 M.A. Walker / Bioorganic & Medicinal Chemistry Letters 27 (2017) 5100–5108

Table 17 4. Yalkowsky SH, Valvani SC. J Pharm Sci. 1980;69:912.


Solubility of 45–47. 5. Jain N, Yalkowsky SH. J Pharm Sci. 2001;90:234.
6. Bergström CAS, Wassvik CM, Johansson K, Hubatsch I. J Med Chem.
O 2007;50:5858.
O R
7. Wager TT, Hou X, Verhoest PR, Villalobos A. ACS Chem Neurosci. 2016;7:767.
N O
N R 8. García-Sosa AT, Maran U, Hetényi C. Curr Med Chem. 2012;19:1646.
O 9. Hitchcock SA. J Med Chem. 2012;55:4877.
10. Sun H, Xia M, Shahane SA, Jadhav A, Austin CP, Huang R. Bioorg Med Chem Lett.
N 2013;23:4587.
11. Walker MA. Topics in medicinal chemistry. In: Meanwell NA, ed. Tactics in
contemporary drug design, vol. 9. Berlin, Heidelberg: Springer; 2015:69–106.
a b 12. Walker MA. Exp Opin Drug Disc. 2014;9:1421.
compd R R MIC mM sol. mg/mL
13. Ahmad NM. Bioorg Med Chem Lett. 2016;26:2975.
45 H H 0.6 2 14. Hann MM. Med Chem Commun. 2011;2:349.
46 H Me 1.2 95 15. Leeson PD, Springthorpe B. Nat Rev Drug Discovery. 2007;6:881.
47 Me Me >75 327 16. Yang B, Hird AW, Russell DJ, et al. Bioorg Med Chem Lett. 2012;22:4907.
17. Yalkowsky SH. J Pharm Sci. 2014;103:2629.
a 18. Ishikawa M, Hashimoto Y. J Med Chem. 2011;54:1539.
M. tuberculosis H27Rv.
b
Dried-from solvent isolate, PBS buffer, 24 h, 25 °C 19. Over B, McCarren P, Artursson P, et al. J Med Chem. 2014;57:2746.
20. Kuhn B, Mohr P, Stahl M. J Med Chem. 2010;53:2601.
21. Dandapani S, Germain AR, Jewett I, et al. ACS Med Chem Lett. 2013;5:149.
22. Goldberg K, Groombridge S, Hudson J, et al. MedChemComm. 2012;3:600.
reported). Invoking Etter’s rule as above (examples 21–23), one 23. Tsukerman SV, Orlov VD, Rozum YS, Lavrushin VF. Dipole moments of
would also expect that the presence of both an H-bond donor diphenyl-1,3,4-oxidiazole and diphenyl-1,2,4-oxidiazole = 3.5 and 1.5 D,
respectively. Chem Heterocycl Compd. 1969;5:461.
and acceptor in a molecule will result in the formation of inter- 24. Toulmin A, Wood JM, Kenny PW. J Med Chem. 2008;51:3720.
molecular H-bonding in the solid state. The introduction of a single 25. Boström J, Hogner A, Llinas A, Wellner E, Plowright AT. J Med Chem.
methyl group (46) results in a significant increase in solubility (sol. 2012;55:1817.
26. Zavareh ER, Hedayati M, Rad LH, et al. Lett Drug Des Disc. 2014;11:721.
= 95 mg/mL,). Assuming that the remaining hydrogen of the sec- 27. Ren F, Deng G, Wang H, et al. J Med Chem. 2012;55:4286.
ondary amide in 46 is capable of acting as an intramolecular H- 28. Varnes JG, Gero T, Huang S, et al. Bioorg Med Chem Lett. 2014;24:3026.
bond donor, solubility should increase by adding an additional 29. Schneider WC. J Am Chem Soc. 1948;70:627.
30. Laurence C, Brameld KA, Graton J, Le Questel JY, Renault EJ. Med Chem Lett.
methyl. Presumably, tertiary amide 47 blocks both intramolecular 2009;52:4073.
an intermolecular H-bonding which would explain its improved 31. Willighagen EL, Denissen HMGW, Wehrens R, Buydens RLMC. J Chem Inf Model.
solubility (sol. = 327 mg/mL) over 45 and 46. The presence of an 2006;46:487.
32. Valko K, Du CM, Bevan CD, Reynolds DP, Abraham MH. J Pharm Sci.
amide-NH appears to be important for activity since 47 is com-
2000;89:1085.
pletely inactive. However, compound 46 achieves a more favorable 33. Alelyunas YW, Liu R, Pelos-Kilby L, Shen C, Eur J. Pharm Sci. 2009;37:172.
balance between activity and solubility compared to 45. 34. Albert A, Hampton A. J Chem Soc. 1954;505.
35. Sorkin E, Roth W. Helv Chim Acta. 1951;34:427.
36. Albert A, Hampton A. J Chem Soc. 1952;4985.
Summary and conclusion 37. Alford EJ, Irving H, Marsh HS, Schofield K. J Chem Soc. 1952;3009.
38. Etter MC. Acc Chem Res. 1990;23:120.
39. Vollhardt P, Schore N. Organic chemistry, structure and function. 7th ed. New
In summary, the methods described in this review provide an York: W.H. Freeman and Company; 2014.
alternative to the standard approach which medicinal chemists 40. Ritchie TJ, Macdonald SJF. The number of papers on this topic is too large to
resort to in order to improve solubility. In theory a drug only needs include in this space. J Med Chem. 2014;57:7206. for a compiled list and
analysis of a number these studies.
to be soluble enough to allow dissolution of the targeted dose.
41. Lovering F, Bikker J, Humblet C. J Med Chem. 2009;52:6752.
Since the dose is likely a multiple of the in vitro activity of the com- 42. Ritchie TJ, Macdonald SJ. Drug Discovery Today. 2009;21–22:1011.
pound high potency will lower the bar for the required solubility. 43. Bergman E. J Org Chem. 1963;28:2210.
44. Matsuoka K, Kajimoto E, Horiuchi M, Honda C, Endo K. Chem Phys Lipids.
However, the problem is that activity usually comes at the price
2010;163:397.
of increased hydrophobicity which leads to lower solubility. The 45. Saad HY, Higuchi WI. J Pharm Sci. 1965;54:1205.
solution to this dilemma is to take advantage of H-bonding interac- 46. Frokjoer S, Anderson VS. Arch Pharm Sci. 1974;2:50.
tions and to employ lipophilic functionality parsimoniously so that 47. Tomida H, Yotsuyanagi T, Ikeda K. Chem Pharm Bull. 1978;26:2832.
48. Bowen DB, James KC, Roberts MJ. J Pharm Pharmacol. 1970;22:518.
LLE is maximized. This will not necessarily result in high solubility. 49. Makhatadze GI, Privalov PL. Biophys Chem. 1994;50:285.
For example, a compound with a Ki = 1 nM can have a log P of 4 and 50. Taylor RD, MacCoss M, Lawson ADG. J Med Chem. 2014;57:5845.
still be considered to bind efficiently (i.e. LLE = 5). Alternatively, 51. Bojarski AJ, Mokrosz Maria J, Paluchowska Maria H. Pharmazie. 1995;50:569.
52. Chrystiuk E, Williams A. J Am Chem Soc. 1987;109:3040.
optimizing LLE by building in H-bonding functionality to increase 53. Bradbury RH, Acton DG, Broadbent NL, et al. Bioorg Med Chem Lett.
activity and lower log P often leads to increases in melting point. 2013;23:1945.
Rather, LLE ensures that the gap between the solubility of the lead 54. Takeuchi T, Oishi S, Kaneda M, et al. ACS Med Chem Lett. 2014;5:566.
55. Leach AG, Jones HD, Cosgrove DA, et al. J Med Chem. 2006;49:6672.
compound and that needed to deliver the targeted dose is narrow. 56. Gleeson P, Bravi G, Modi S, Lowe D. Bioorg Med Chem. 2009;17:5906.
Under such circumstances the best approach for optimizing solu- 57. Zhang L, Zhu H, Mathiowetz A, Gao H. Bioorg Med Chem. 2011;19:5763.
bility further would be to employ minor structural changes that 58. Iwasa J, Fujita T, Hansch C. J Med Chem. 1965;8:150.
59. Toxnet Toxicology Data Network, https://toxnet.nlm.nih.gov, accessed
do not disturb potency and PK. This review identifies several minor 14.04.17.
structural changes which lead to improved solubility and can be 60. Suzuki TJ. Comput. Aided Mol Des. 1991;5:149.
applied to diverse templates. 61. Ruelle P, Kesselring UW. J Pharm Sci. 1998;87:1015.
62. Böhm H-J, Banner D, Bendels S, Kansy M, Kuhn B, Müller K, Obst-Sander U, Stahl
M. ChemBioChem. 2004;5:637.
References 63. Wenglowsky S, Moreno D, Rudolph J, Ran Y, Ahrendt KA, Arrigo A, Colson B,
Gloor SL, Hastings G. Bioorg Med Chem Lett. 2012;22:912.
1. Hansch C. Acc Chem Res. 1969;2:232. 64. Xia G, Benmohamed R, Morimoto R, Kirsch DR, Silverman RB. Bioorg Med Chem
2. Hansch C, Dunn WJ. J Pharm Sci. 1972;61:1. Lett. 2014;24:5098.
3. Leo AJ, Hansch C. Perspect Drug Disc. Des.. 1999;17:1. 65. Kedari CK, Choudhury NR, Sharma S, et al. ACS Med Chem Lett. 2014;5:491.

You might also like