You are on page 1of 12

1066 Biomacromolecules 2004, 5, 1066-1077

Molecular Modeling and Simulation of Mycobacterium


tuberculosis Cell Wall Permeability
Xuan Hong and A. J. Hopfinger*
Laboratory of Molecular Modeling and Design (MC 781), College of Pharmacy, The University of Illinois at
Chicago, 833 South Wood Street, Chicago, Illinois 60612-7231
Received December 8, 2003; Revised Manuscript Received February 9, 2004

The low permeability of the mycobacterial cell wall is thought to contribute to the intrinsic drug resistance
of mycobacteria. In this study, the permeability of the Mycobacterium tuberculosis cell wall is studied by
computer simulation. Thirteen known drugs with diverse chemical structures were modeled as solutes
undergoing transport across a model for the M. tuberculosis cell wall. The properties of the solute-membrane
complexes were investigated by means of molecular dynamics simulation, especially the diffusion coefficients
of the solute molecules inside the cell wall. The molecular shape of the solute was found to be an important
factor for permeation through the M. tuberculosis cell wall. Predominant lateral diffusion within, as opposed
to transverse diffusion across, the membrane/cell wall system was observed for some solutes. The extent of
lateral diffusion relative to transverse diffusion of a solute within a biological cell membrane may be an
important finding with respect to absorption distribution, metabolism, elimination, and toxicity properties
of drug candidates. Molecular similarity measures among the solutes were computed, and the results suggest
that compounds having high molecular similarity will display similar transport behavior in a common
membrane/cell wall environment. In addition, the diffusion coefficients of the solute molecules across the
M. tuberculosis cell wall model were compared to those across the monolayers of dipalmitoylphosphati-
dylethanolamine and dimyristoylphosphatidylcholine, are two common phospholipids in bacterial and animal
membranes. The differences among these three groups of diffusion coefficients were observed and analyzed.

Introduction Mycobacterial mycolic acids have several distinctive


Tuberculosis (TB) kills approximately 2 million people features compared to most fatty acids: (1) they are long-
each year.1 The global incidence of TB is still increasing at chain molecules having a long branch, mero chain, of 40-
approximately 0.4% per year and much faster in some areas.2 60 carbons and a short branch, R branch, of typically 24
The devastating impact of this disease is attributed to three carbons; and (2) in addition to their extraordinary lengths,
factors: the breakdown in health services, the co-infection mycolic acids contain very few double bonds or cyclopropane
with HIV/AIDS, and the emergence of multidrug-resistant groups. The short branch, R branch, is totally saturated, and
TB.1 the long branch, mero chain, only has two positions that are
Mycobacteria infections are, in general, difficult to treat observed to be occupied by functional groups. Of these two
because mycobacteria are resistant to the majority of common positions, the proximal position (nearer the β-hydroxy acid)
antibiotics and chemotherapeutic agents.3,4 An important contains exclusively cis- or trans-olefin or cyclopropane,
factor causing this resistance is the barrier imposed by the while the distal position may be the same as the proximal
mycobacterial cell wall.5-10 The mycobacterial cell wall is position or may contain one of a variety of oxygen moieties
extraordinarily thick and tight, having two main compo- such as R-methyl ketone, R-methyl methyl ether, methyl-
nents: (1) characteristic long chain fatty acids, mycolic acids, branched ester, or R-methyl epoxide.
and (2) unique polysaccharides, arabinogalactan (AG).11-13
These two constituents are covalently linked together by ester There are three kinds of mycolates in the M. tuberculosis
bonds. The mycolyl-AG complex is then attached to cell wall, and they are R-mycolates, methoxymycolates, and
peptidoglycan, a porous layer between the cell wall and the ketomycolates (Figure 1). Both the R- and methoxymycolates
plasma membrane, and form the mycolyl-AG peptidoglycan only have the cis-cyclopropyl group at the proximal position,
complex. The mycobacterial cell wall also contains many while 17% of the cyclopropyl groups at the keto proximal
“free” lipid species, the so-called extractable lipids, that are position are trans.18 Fifty-one percent of the total mycolates
not covalently linked to the AG-peptidoglycan complex and are R-mycolates, 36% are methoxymycolates, and 13% are
are solvent-extractable. The free lipids include glycolipids, ketomycolates.19 All three mycolates contain 24 and 26
phenolic glycolipids, glycopeptidolipids, and other chemical carbon R branches in approximately the ratio of 10:90, with
species.11-17 negligible amounts of 22 carbon R branches. Methoxy-
* Corresponding author. Telephone: 312.996.4816. Fax: 312.413.3479. mycolates and ketomycolates have longer mero chains than
E-mail: hopfingr@uic.edu. R-mycolates. The total carbon numbers for the R, methoxy,
10.1021/bm0345155 CCC: $27.50 © 2004 American Chemical Society
Published on Web 04/06/2004
Mycobacterium tuberculosis Cell Wall Permeability Biomacromolecules, Vol. 5, No. 3, 2004 1067

Figure 1. Structures of mycolic acids found in the M. tuberculosis cell wall. The usual value for k is 23. The sum of n, m, and h is approximately
50. The shorter chain is the R branch, and the longer one is the mero chain. Both R and β carbons have R chirality. (See text for details.)
Reprinted with permission from Hong and Hopfinger (2004). Copyright American Chemical Society.

and keto forms are 76-82, 83-90, and 84-89, respec- from entering the cell, which attributes, at least in part, to
tively.20 the intrinsic drug resistance of mycobacteria.7
AG, the other main component of the mycobacterial cell In the study reported in the preceding paper, conforma-
wall, is a polysaccharide consisting of arabinose (Ara) and tional search and molecular dynamics simulation (MDS)
galactose (Gal).21-23 Within AG, all arabinose and galactose were employed to investigate the conformational profile of
residues are in the furanose form, and mycolic acids are AG and subsequently the conformations and structural
located in clusters of four on the terminal hexaarabinofura- organization of the mycolyl-AG complex.25 An inner leaflet
noside through 1,5 linkages. However, only two-thirds of molecular model of the M. tuberculosis cell wall was
the terminal arabinose residues are mycolated. The linker constructed. (See the summary given in the Methods section.)
disaccharidephosphate connects the galactan region of AG The mycolate hydrocarbon chains were determined to be
to a peptidoglycan. Figure 2 provides an overview of the tightly packed and perpendicular to the “plane” formed by
structure and organization of the mycolyl-AG peptidoglycan the oxygen atoms of the 5-hydroxyl groups of the terminal
complex. arabinose residues to which the mycolic acids bind. The
In the physical organization model for the mycobacterial average packing distance between mycolic acids of M.
cell wall proposed by Minnikin, the mycobacterial cell wall tuberculosis was estimated to be approximately 7.3 Å. This
is composed of an asymmetric lipid bilayer (Figure 3).11 The previous computational study provides support for the
inner leaflet contains mycolic acids covalently linked to AG. Minnikin model and also presents a way to probe the
In this leaflet, mycolic acid hydrocarbon chains are tightly mechanism of low permeability of the cell wall and, in turn,
packed in a parallel fashion, and predominantly oriented in the intrinsic drug resistance of M. tuberculosis.
a direction perpendicular to the cell wall surface. The outer To investigate M. tuberculosis cell wall permeability to
leaflet contains the extractable lipids that accommodate the organic solutes in a quantitive manner and to gain more
uneven surface of the inner leaflet caused by the two insight into the nature of this barrier, a variety of drugs were
branches of mycolic acid that are not equal in length. It has modeled as solutes undergoing transport across the M.
been shown that the outer leaflet is moderately fluid, while tuberculosis cell wall. The properties of the solute-
the inner leaflet has very low fluidity.24 This low fluidity of membrane complexes were studied by the means of MDS,
the inner leaflet may be the result of the unique chemical especially the diffusion coefficients of the solute molecules
structures of mycolic acids and their tight packing, which, inside the cell wall. For the purpose of comparing the
in turn, could account for the low permeability of the overall permeability of the M. tuberculosis cell wall to those of
cell wall. The cell wall permeability barrier prevents drugs bacterial and animal membranes, models for the complexes
1068 Biomacromolecules, Vol. 5, No. 3, 2004 Hong and Hopfinger

Figure 2. Overview of the mycolyl-AG peptidoglycan complex (with permission from the Annual Review of Biochemistry, Volume 64, copyright
1995 by Annual Reviews www.annualreviews.org; minor modifications made).

study.25 As mentioned in the Introduction, the innermost part


of the cell wall, consisting solely of mycolic acids, is believed
to be the most tightly packed membrane region and has the
lowest permeability of the cell wall. Therefore, the mycolic
acid region was the focus in the construction of the M.
tuberculosis cell wall. Pseudo-mycolic acids (PMAs) were
designed to model this cell wall region (Figure 4). There
are 24 carbons in the R branch of a PMA molecule. The
mero chain was shortened in the modeling to the same length
as the R branch. A cis-cyclopropyl group was chosen to be
the functional group at the proximal position. The advantages
of this structural design are (1) the portions of the mycolic
acids that are located in the innermost cell wall, and
Figure 3. Mycobacterial cell wall model proposed by Minnikin. (The presumably are responsible for the cell wall low permeability,
picture is taken from www.niaid.nih.gov/newsroom/focuson/tb02/ are adequately represented by PMA molecules and (2) such
target.htm with modifications.) The funnel-shape structure in the
center of the cell wall represents a porin (for the description of the
a PMA molecule represents all of the major types of mycolic
function of porins, see ref 25). acids present in M. tuberculosis because these molecules have
similar structures to one another except for the distal moieties
of each of the same set of solutes and monolayers of (Figure 1) and a cis-cyclopropyl group is the dominant
dipalmitoylphosphatidylethanolamine (DPPE) and dimyris- functional group at the proximal position.
toylphosphatidylcholine (DMPC) were also constructed and Even though it is known that the mycolic acids are very
studied, respectively. In addition, molecular similarity mea- tightly packed in the inner leaflet, there is very little data
sures among the solutes were computed with the hope of available regarding the actual packing (including distance)
learning more about the solute features that govern perme- between mycolic acids. Therefore, the AG complex, the
ability across the M. tuberculosis cell wall. polysaccharides to which the mycolic acids form ester bonds,
was studied. Conformational analysis was applied. However,
Methods one AG complex is composed of three arabinans and one
galactan, and nearly 100 sugar residues are involved.
1. Construction of a M. tuberculosis Cell Wall Model. Consequently, there are hundreds of torsion angle degrees
The procedures in the construction of a M. tuberculosis cell of conformational freedom. A complete systematic torsion
wall model have been reported in the preceding paper and angle conformational search of such a complex is not
are only summarized here to facilitate understanding of this realistic. Thus, the branch containing the 17 arabinose
Mycobacterium tuberculosis Cell Wall Permeability Biomacromolecules, Vol. 5, No. 3, 2004 1069

Figure 4. Chemical structure of a PMA molecule. Carbon atoms located at position 13 on both the R and the mero chains are pointed out.

residues that are close to the termini of the arabinan chain and fourth columns of the PMs. The primary (largest) inertial
and covalently linked to the mycolic acids was chosen to axis of each solute was aligned parallel to the Z axis of the
represent the structural features of the AG complex (Figure mycolates (i.e., the direction of the mycolate hydrocarbon
2). Exhaustive random conformational sampling was per- chains). The solute was then moved along the Z axis until
formed on this structure. The conformations generated were its center of mass was in the “plane” formed by the carbon-
then ranked according to their conformational potential 13 atoms of the mycolates’ mero chains; see Figure 4.
energies. The low energy conformations were chosen to form The 13 cell wall-solute complexes generated were each
complexes with the PMAs. Energy minimization was per- subjected to extensive MDS. The MOLSIM package,29 with
formed on the resultant PMA-AG complexes, and the an extended MM2 force field,30,31 was used to perform the
complex with the lowest conformational potential energy was MDS. The simulation temperature was set at 311 K and held
used as a starting point in extensive MDS. The average constant during the MDS by coupling the system to an
distance between the oxygen atoms of the 5-OH groups of external fixed temperature bath.32 The initial periodic bound-
the terminal arabinose residues, where the mycolic acids bind, ary conditions were set at 36.5 Å for the X and Y dimensions,
was calculated on the basis of the MD trajectories taken after 80 Å for the Z dimension, and 90° for the surface orientation
the complex had reached equilibrium. Because the mycolic angle, γ. The simulation sampling time was 140 ps with
acids exist as mycolates in the inner leaflet, a pseudo- intervals of 0.001 ps for a total sampling of 140 000
mycolate (PM) was generated by esterifying the carboxyl conformations for each of the complexes. Geometries gener-
group of a PMA molecule with methanol. A PM monolayer ated during the MDS were recorded every 0.1 ps. Three
was then constructed using the average packing distance different random seeds were used to initiate the MDS. The
between the oxygens of the 5-OH groups that, in turn, MDS routine of the MOLSIM software package, as described
represents the inner leaflet of the M. tuberculosis cell wall. above, was applied throughout this study unless stated
2. Construction of the 13 Solutes. Thirteen known drugs otherwise. Methyl ester carbons and the terminal carbons of
were selected as solutes to study the permeability of the M. the PM mero chains were each assigned heavy masses of
tuberculosis cell wall. These drugs include seven first- and 1000 amu. The purpose of the heavy mass assignment is to
second-line anti-TB drugs, five general antimycobacterial take the influence of additional bonded structure into
drugs, and one compound that has no antimycobacterial consideration. For example, restrictions in the movement of
activity. The structures of these 13 “solutes” are given in the terminal carbons would occur in a MDS if the mero
Figure 5. The HyperChem 6.01 software26 was used to build chains had not been artificially shortened.
the three-dimensional structures of the solutes. The MM+ The displacement, ∆di, between two adjacent time steps
molecular mechanics force field, also implemented in Hy- was determined for each atom of each solute from its MD
perChem, was utilized in the solute structure optimizations. trajectory, and the individual diffusion coefficient of every
Each structure was subject to conjugate gradient energy solute atom was calculated using the mean-square displace-
minimization using the Polak-Ribiere first derivative ment method.33 The mean-square displacement is given as
method,27 until a derivative convergence criterion of 0.1 kcal/ n
(mol Å) was reached or until a maximum number of
iterations were performed. The HyperChem software sets the ∑
i)1
∆di2
maximum number of iterations to be 15 times the number χ2 ) (1)
of atoms in the system. After each structure was energy n
minimized in the MM+ force field, it was further energy where n is the total number of time steps in MDS. From the
minimized by the AM1 semiempirical method.28 Partial Einstein diffusion equation, the diffusion coefficient, D, is
atomic charges were estimated by performing a single point given by
AM1 calculation on the minimized structure.
3. Construction and MDS of the Complexes of the M. χ2
tuberculosis Cell Wall Model with the 13 Solutes. On the D) (2)
2∆t
basis of the estimated average packing distance between
mycolic acids from our previous study, the unit cell where ∆t is the size of one time step.34 The diffusion
parameters of the PM monlayer were selected to be a ) 7.3 coefficient of the complete solute molecule is then calculated
Å, b ) 7.3 Å, and γ ) 90°.25 Each solute was inserted into as the average of the diffusion coefficients of all atoms of
the M. tuberculosis cell wall, modeled by the PM monolayer, the solute.
as is illustrated in Figure 6. The solutes were located in the 4. Construction and MDS of the Complexes of the
region defined by the second and third rows and the third DPPE Monolayer Model with the Solutes. A DPPE
1070 Biomacromolecules, Vol. 5, No. 3, 2004 Hong and Hopfinger

Figure 5. Chemical structures of the 13 solute molecules considered in this study.

monolayer model was previously built25 (see Figure 7A for M. tuberculosis cell wall exists in a relatively highly ordered
the structure of DPPE). The complexes of the DPPE state. Hence, the solutes were inserted into an intact PM
monolayer with the thirteen solutes were constructed using monolayer model to mimic this dense and highly ordered
the MI-QSAR software.35 To prevent unfavorable van der state.
Waals interaction between the solute molecule and the Each solute molecule was placed between the regions of
membrane DPPE, one of the “center” DPPE molecules was head groups and aliphatic chains of the monolayer with the
removed from the monolayer model and the solute molecule most polar group of the solute molecule “facing” toward the
was inserted in the space created by the missing DPPE head group region of the monolayer. MDSs were performed
molecule. This procedure was not applied in the construction on the monolayer-solute complex using MOLSIM. The
of the solute-PM complex model. In the previous study,25 periodic boundary conditions, which were used to build the
it was found that the average packing distance of mycolic DPPE monolayer model, were employed (a ) 37.5 Å, b )
acids in a PM monolayer is correspondingly less than that 37.5 Å, c ) 80 Å, and γ ) 92°).25 The monolayer-solute
of DPPEs in a DPPE monolayer, and the inner leaflet of the complexes were simulated for 70 ps using time intervals of
Mycobacterium tuberculosis Cell Wall Permeability Biomacromolecules, Vol. 5, No. 3, 2004 1071

Table 1. IPEs Used in the Molecular Similarity Analyses


IPE code symbol definitions
0 ALL all atoms in the molecule
1 NP nonpolar atoms
2 P+ polar atoms with positive charge
3 P- polar atoms with negative charge
4 HBA hydrogen bond acceptor atoms
5 HBD hydrogen bond donor atoms
6 ARO aromatic atoms
7 HS non-hydrogen atoms (hydrogen suppressed)

were calculated for the solute molecules in the same manner


as in PM and DPPE.
Additional information regarding the construction of a
phospholipid monolayer, and its complexes with solutes,
using the MI-QSAR software can be found in refs 36-38.
6. Molecular Similarity Analysis of the 13 Solutes. The
4D-MS module of the 4D-QSAR software package was used
to perform the four-dimensional molecular similarity study
of the 13 solutes.39,40 4D-QSAR molecular similarity analysis
considers both the three-dimensional molecular structure and
the corresponding complete conformational ensemble of
states of a molecule in estimating molecular similarity. The
formalism measures relative molecular similarity (RMS) that
Figure 6. Illustration of an initial insertion, position, and alignment
of a solute molecule into the M. tuberculosis cell wall model using
is dependent upon an alignment constraint (an external
ethambutol as an example. reference frame) and absolute molecular similarity (AMS)
that is alignment-independent. This method also allows
0.001 ps. The diffusion coefficients of the solutes in the estimation of molecular similarity measures on the basis of
DPPE model membrane were then calculated in the same atom types composing molecules. 4D-QSAR defines eight
manner as just described for the M. tuberculosis cell wall atom types and names them as interaction pharmacophore
model. elements, IPEs (Table 1). In this study, AMS analysis based
5. Construction and MDS of the Complexes of the on the eight IPE types was applied. MDS was performed to
DMPC Monolayer Model with the Solutes. The complexes generate a conformational ensemble profile (CEP) for every
of the DMPC monolayer model with the solute molecules solute molecule. The MOLSIM package was again used to
were constructed in the same manner as in step 4 for DPPE perform the simulations. The simulation sampling time was
(see Figure 7B for the structure of DMPC). The MDSs were 70 ps using time intervals of 0.001 ps. The atomic coordi-
carried out using the MOLSIM package. The periodic nates of all conformations sampled for every solute were
boundary conditions employed are a ) 40 Å, b ) 40 Å, c recorded every 0.1 ps and stored in a trajectory file that, in
) 80 Å, and γ ) 96°. The simulation sampling time was 70 turn, constituted the corresponding CEP of the solute. For
ps using time intervals of 0.001 ps. Diffusion coefficients the detailed methodology of 4D-MS formalism, see ref 40.

Figure 7. Chemical structures of DPPE and DMPC.


1072 Biomacromolecules, Vol. 5, No. 3, 2004 Hong and Hopfinger

Table 2. MW and Calculated log P (calcd log P) Values of the


Solute Molecules
solute MW calcd log Pa
First- and Second-Line Anti-TB Drugs
isoniazid 137.14 0.49
PAS 153.14 0.68
ethionamide 166.24 1.97
ethambutol 204.31 0.29
ciprofloxacin 331.34 0.67
ofloxacin 361.37 0.62
clofazimine 473.40 6.98
General Antimycobacterial Drugs
amithiazone 236.29 1.24
dapsone 248.30 1.31
hydnocarpic acid 252.40 4.78
chaulmoogric acid 280.45 5.57
thiocarlide 400.58 6.72
Non-Antimycobacterial Drug
phencycline 243.39 3.98
a Calcd log P values were calculated from the calcd log P module in
the HyperChem 6.01 package.

Figure 8. Three-dimensional structures of clofazimine, ethionamide,


and ethambutol after structure optimization: H (white), C (grey), N
(blue), Cl (green), S (yellow), and O (red).

Results

1. Permeation of the Solute Molecules through the M.


tuberculosis Cell Wall Model. The 13 solute molecules
considered in this study can be classified into three categories
according to their molecular shapes: (1) bulky, as represented
by clofazimine, and including clofazimine, phencycline, Figure 9. Relationship between mean-square displacement of the
ciprofloxacin, and ofloxacin, where the first two molecules solute (10-20 cm2) and simulation time (ps). The thick green-colored
are much bulkier than the last two (phencycline, the only curve is made of data points. The regression line (in red) and fit are
shown. The diffusion coefficient calculated here corresponds to Dma
non-antimycobacterial drug in the dataset, was chosen for (2) in Table 3.
its bulky and “football”-like molecular shape.); (2) flat, as
represented by ethionamide, and including isoniazide, ethion- dataset of 13 solutes covers a relatively large and diverse
amide, p-aminosalicyclic acid (PAS), amithiazone, dapsone, chemical space regarding molecular shape, size, and polarity.
and thiocarlide; and (3) linear, as represented by ethambutol, Figure 9 illustrates the relationship between mean-square
and including ethambutol, hydnocarpic acid, and chaulmoo- displacements (χ2) of the solutes in the M. tuberculosis cell
gric acid. The three-dimensional structures of the three wall model and simulation time (t), using ethambutol as an
representative molecules, clofazimine, ethionamide, and example. The diffusion coefficients of the solute molecules,
ethambutol, after structure optimization, are shown in Figure calculated as 1/2 of the slope of the linear regression line of
8. χ2 versus time (see eq 2), are given in Table 3 in the form
In addition to steric features, the molecular size, reflected of Dma, where “D” stands for diffusion coefficient and “ma”
by molecular weight (MW), of the solute molecules also stands for mycolic acid. Bulky solutes have a lower Dma than
varies significantly (Table 2). Isoniazid is the smallest flat and linear solutes, while there is no significant difference
molecule in the dataset and has a MW of 137 amu, while between solutes in the latter two classes. Moreover, it is
clofazimine has the highest MW of 473 amu. The solute difficult to find a straightforward relationship between Dma
molecules also possess a considerable range in lipophilicity values of the solutes and their molecular shapes and
(Table 2) and can be grouped into two classes based on their polarities. The geometry of a M. tuberculosis cell wall-solute
log P values being greater (class 1) or less (class 2) than 2. complex, at the end of MDS, is exemplified using ethambutol
Clofazimine, hydnocarpic acid, chaulmoogric acid, thiocar- (Figure 10). The hydrocarbon chains of the mycolates remain
lide, and phencycline belong to the first class, and the rest tightly packed and perpendicular to the cell surface through-
of the compounds fall into the second class. Overall, this out the MDS.
Mycobacterium tuberculosis Cell Wall Permeability Biomacromolecules, Vol. 5, No. 3, 2004 1073

Table 3. Diffusion Coefficients (10-8 cm2/s) of the Solute


Molecules in the M. tuberculosis Cell Wall (Dma)a
solute 〈Dma〉 Dma (1) Dma (2) Dma (3)
isoniazid 1.0 1.1 0.9 1.0
PAS 1.0 1.0 1.0 1.0
ethionamide 1.0 0.9 1.0 1.0
ethambutol 1.3 1.2 1.3 1.3
ciprofloxacin 0.8 0.8 0.7 0.8
ofloxacin 0.7 0.7 0.7 0.7
clofazimine 0.7 0.8 0.6 0.7
amithiazone 0.9 0.8 1.0 0.9
dapsone 1.3 1.6 1.1 1.1
hydnocarpic acid 1.1 1.1 1.0 1.2
chaulmoogric acid 1.1 1.1 1.1 1.1
thiocarlide 1.1 1.2 1.1 1.1
phencycline 0.8 0.7 0.7 0.9
a The D
ma (1-3) were calculated from the molecular dynamics trajec-
tories using three different random seeds to inititate the MDS. The 〈Dma〉
values reported in the second column are the average values from the
three MDSs.

Figure 11. Axial ∆d2 (10-20 cm2) components of clofazimine,


ethambutol, and chaulmoogric acid in the membrane models over
the course of the MDS. The unit of simulation time is ps. The color
coding for the X, Y (lateral diffusion), and Z (transverse diffusion) is
given in the clofazimine plot.
are referred to as ∆dX2, ∆dY2, and ∆dZ2, respectively.
According to the different patterns of movement exhibited
by the solutes inside the cell wall model, the solute molecules
can be classified into three groups. Solutes in the first group
Figure 10. Structure of the membrane-solute complex of the M.
tuberculosis cell wall with ethambutol at the end of MDS. The move in small ranges relative to each of the three axes. These
hydrogen atoms of the cell wall are shown in white, the carbon atoms solutes are clofazimine, ciprofloxacin, ofloxacin, and phen-
in gray, and the oxygen atoms in red. The solute, ethambutol, is cycline. The second group of solutes also moves uniformly
shown in CPK form. along all three axes, but the overall movement is larger than
Some solutes during the MDS diffusion experiments that of solutes in the first group. Ethambutol, dapsone,
predominantly move inside the membrane in directions other isoniazid, ethionamide, amithiazone, PAS, and thiocarlide
than through the cell. These molecules may have high exhibit group 2 behavior. Chaulmoogric acid and hydno-
diffusion coefficients, but their actual uptake into the interior carpic acid have considerably different movements from the
of a cell is low. The measurement of a diffusion coefficient solutes of groups 1 and 2. These two solutes do not move
cannot detect this lateral component to net diffusion. Thus, equally along each axis, instead they undergo much more
the observed diffusion coefficient can, perhaps, be misleading movement in the membrane, the X and Y directions, than
of the cellular uptake of the compound. through the membrane, the Z direction. Chaulmoogric acid
For the solutes in Figure 5, the individual movements and hydnocarpic acid, thus, make up the third group of
along the X, Y, and Z axes of each solute were calculated solutes.
from the distance that the molecule moves parallel to each Clofazimine, ethambutol, and chaulmoogric acid are
axis during every single MDS time step. These distances representative of the three solute groupings with respect to
1074 Biomacromolecules, Vol. 5, No. 3, 2004 Hong and Hopfinger

Table 4. Components of the Diffusion Coefficients with Respect to the Individual Axes (10-8 cm2/s) for the Solute Molecules in the M.
tuberculosis Cell Walla
〈Dma〉 Dma (1) Dma (2) Dma (3)
solute 〈Dx〉 〈Dy〉 〈Dz〉 Dx Dy Dz Dx Dy Dz Dx Dy Dz
isoniazid 0.4 0.4 0.3 0.4 0.4 0.3 0.3 0.4 0.3 0.4 0.4 0.3
PAS 0.4 0.4 0.3 0.4 0.4 0.4 0.4 0.4 0.3 0.4 0.4 0.3
ethionamide 0.3 0.4 0.3 0.4 0.3 0.3 0.3 0.4 0.3 0.3 0.4 0.3
ethambutol 0.5 0.5 0.3 0.5 0.5 0.3 0.5 0.4 0.4 0.5 0.5 0.4
ciprofloxacin 0.3 0.3 0.2 0.3 0.3 0.3 0.3 0.2 0.2 0.3 0.3 0.2
ofloxacin 0.3 0.3 0.2 0.3 0.3 0.2 0.3 0.3 0.2 0.2 0.3 0.2
clofazimine 0.3 0.3 0.2 0.3 0.3 0.3 0.2 0.2 0.2 0.3 0.3 0.2
amithiazone 0.3 0.4 0.3 0.3 0.3 0.2 0.3 0.4 0.3 0.3 0.4 0.3
dapsone 0.5 0.5 0.4 0.6 0.5 0.6 0.4 0.4 0.4 0.4 0.4 0.3
hydnocarpic acid 0.4 0.5 0.2 0.4 0.5 0.2 0.4 0.5 0.2 0.5 0.5 0.2
chaulmoogric acid 0.5 0.5 0.2 0.5 0.4 0.2 0.4 0.6 0.2 0.4 0.5 0.2
thiocarlide 0.3 0.2 0.2 0.3 0.3 0.2 0.3 0.2 0.2 0.2 0.2 0.2
phencycline 0.3 0.3 0.2 0.3 0.3 0.2 0.3 0.3 0.2 0.2 0.3 0.2
a The indices, (1-3), represent the MDS experiments initiated by three different random seeds. The diffusion coefficients reported in the second to

fourth columns are the average values from the three MDS experiments. The diffusion coefficients of hydnocarpic acid and chaulmoogric acid, which
exhibit significant lateral diffusion, are shown in bold.

Table 5. Two-Dimensional and Four-Dimensional AMS Molecular


diffusion and their movement patterns are shown in Figure Similarity Measurements between Pairs of Solute Moleculesa
11. These movement patterns, especially those of the third
Part A. Two-Dimensional Molecular Similarity
group, can also be seen in Table 4, where Dma is partitioned chaulmoogric acid
into Dma-x, Dma-y and Dma-z. The Dma-z of hydnocarpic acid hydnocarpic acid
and chaulmoogric acid (in bold) are significantly smaller than 2D 0.906
the corresponding Dma-x and Dma-y (in bold). There is a
Part B. Four-Dimensional AMS
correlation between the steric shape of a molecule and its amithiazone isoniaid ciprofloxacin
movement pattern inside the M. tuberculosis cell wall. All dapsone PAS ofloxacin
the spherical and bulky molecules are in the first group, while ALL 0.938 0.935 0.901
linear and flat molecules compose the second and third isoniaid
groups. For every solute in the first and second groups, the PAS
extent of its movement relative to each reference axis is NP 0.961
proportional to its Dma value. However, the large Dma values chaulmoogric acid isoniazid isoniazid PAS
of the third group of solutes do not reflect their penetration hydnocarpic acid ofloxacin ethionamide amithiazone
rates. Chaulmoogric acid and hydnocarpic acid both resemble P+ 1.000 0.942 0.939 0.917
the shape, charge distribution, and polar/nonpolar asymmetry PAS isoniazid
of a PM molecule, which may promote their preferred lateral ethambutol ethambutol
movements within, as opposed to through, the membrane. P- 0.957 0.907
In fact, these two molecules are readily taken up into the clofazimine PAS PAS isoniazid
PM monolayer, and by the end of MDS the polar head groups ethambutol ethambutol clofazimine ethambutol
of these acid solutes are located in the same region as the HBA 0.961 0.957 0.937 0.907
head groups of mycolates. Lateral movement is, in fact, quite chaulmoogric acid ciprofloxacin amithiazone PAS
common for molecules composing membranes and includes hydnocarpic acid thiocarlide ethambutol amithiazone
phospholipids and membrane proteins.41,42 The first observa- HBD 1.000 0.954 0.925 0.917
tion of this event using MDS has been reported by Moore ofloxacin phencycline ciprofloxacin ciprofloxacin
and co-workers recently.43 PAS PAS ofloxacin phencycline
Table 5 lists the two-dimensional and four-dimensional ARO 1.000 0.999 0.999 0.999
(alignment independent) AMSs for the solute molecules. The ofloxacin isoniazid amithiazone amithiazone
AMS consists of measurements based on the eight IPE types. phencycline ethionamide PAS ciprofloxacin
Only molecule pairs having their similarity measures greater ARO 0.999 0.999 0.999 0.999
than 0.900 are considered similar in this study and are given amithiazone amithiazone PAS
in Table 5. Most spherical and bulky compounds are similar ofloxacin phencycline ciprofloxacin
to one another, such as ciprofloxacin and ofloxacin, having ARO 0.999 0.999 0.999
the ALL-AMS, ARO-AMS and HS-AMS measures PAS ciprofloxacin
greater than 0.900. Some flat-shaped molecules are also ethionamide ofloxacin
similar, including isoniazid and PAS, and amithiazone and HS 0.941 0.922
dapsone, and so forth. Chaulmoogric acid and hydnocarpic a A value of 0.000 means there is no similarity and a value of 1.000

acid are similar, having a two-dimensional molecular simi- implies identical similarity.
larity measure of 0.906 and the AMS P+-AMS and HBD- It is a core working hypothesis in medicinal chemistry that
AMS have measured similarities of 1.000. structurally similar molecules are likely to express similar
Mycobacterium tuberculosis Cell Wall Permeability Biomacromolecules, Vol. 5, No. 3, 2004 1075

Table 6. Diffusion Coefficients (10-8 cm2/s) of the Solute Molecules in the DPPE and DMPC Monolayers (DDPPE, DDMPC)
solute 〈DDPPE〉 DDPPE (1) DDPPE (2) DDPPE (3) 〈DDMPC〉 DDMPC (1) DDMPC (2) DDMPC (3)
isoniazid 1.5 1.6 1.5 1.5 1.3 1.4 1.2 1.3
PAS 1.6 1.7 1.5 1.5 1.2 1.3 1.3 1.1
ethionamide 1.5 1.6 1.5 1.5 1.3 1.3 1.3 1.4
ethambutol 1.7 1.8 1.6 1.6 1.6 1.6 1.6 1.7
ciprofloxacin 1.2 1.2 1.2 1.1 1.0 1.0 1.0 1.0
ofloxacin 1.0 1.0 1.1 0.9 0.9 0.9 0.9 1.0
clofazimine 1.0 1.0 0.9 1.0 0.8 0.8 0.9 0.8
amithiazone 1.7 1.8 1.5 1.7 1.4 1.3 1.4 1.3
dapsone 1.7 1.7 1.8 1.7 1.4 1.5 1.4 1.4
hydnocarpic acid 1.6 1.5 1.6 1.6 1.4 1.4 1.4 1.3
chaulmoogric acid 1.4 1.4 1.4 1.5 1.5 1.4 1.5 1.5
thiocarlide 1.3 1.3 1.4 1.3 1.2 1.1 1.2 1.2
phencycline 1.2 1.2 1.2 1.2 0.9 0.9 0.9 1.0

biological behavior profiles including absorption distribution, and the corresponding absorption of solute inside the cell
metabolism, elimination, and toxicity (ADMET) properties. are both low even though the diffusion coefficient may be
In this case, it would appear that similar compounds have large.
similar permeation behavior within a common cell wall Lateral diffusion may also affect cell wall function. In fact,
environment. This conclusion is supported by several ex- lipid analysis of M. Vaccae grown in the presence of
amples, such as chaulmoogric acid and hydnocarpic acid, chaulmoogric acid has demonstrated that chaulmoogric acid
isoniazid and PAS, amithiazone and dapsone, and the is taken up by the organism and incorporated into cell wall
spherical and bulky solutes. components.44 It is also observed that cell growth is retarded
2. Comparison of Diffusion Coefficients from the DPPE by the addition of chaulmoogric acid to the growth medium.44
and the DMPC Monolayer Models. There is no significant Therefore, it is interesting to postulate that the antimyco-
difference between the DDPPE and the DDMPC values for the bacterial properties of chaulmoogric acid result from its
solute molecules studied (Table 6). But, as expected, solutes uptake into or distortion of the cell wall structure. In the
in both phospholipid membrane models have correspondingly bacterial and animal membranes modeled by DPPE and
larger diffusion coefficients than the Dma of the M. tuber- DMPC monolayers, respectively, chaulmoogric acid under-
culosis cell wall model. Most solutes retain three-dimensional goes transverse diffusion across the membrane equally well
movement patterns in both the DPPE and DMPC monolayers compared to lateral diffusion. Thus, this compound is not
such as those observed in the M. tuberculosis cell wall model, readily retained inside the phospholipid membranes and its
except for chaulmoogric acid. This molecule moves unifor- cellular absorption and distribution properties are likely to
mally with respect to the three reference axes in the DPPE reflect its diffusion coefficients in these membranes. Al-
and DMPC monolayers. Only hydnocarpic acid from the though chaulmoogric acid is traditionally used to treat
solute dataset retains its principal movement within, as leprosy, this molecule might also provide a promising anti-
opposed to through, the membrane in the DPPE and DMPC TB mechanism: the uptake into or disruption of the M.
monolayers. That is, hydnocarpic acid undergoes significant tuberculosis cell wall structure. Thus, chaulmoogric acid
lateral diffusion in both DPPE and DMPC membranes. could be an important lead for future anti-TB drug develop-
ment. Hydnocarpic acid, similar to chaulmoogric acid, is also
Discussion used to treat leprosy, and its antimycobacterial effects may
also be due to the uptake into or distortion of the mycobac-
The molecular shape of a solute is found to be an important terial cell wall structure.
factor for the permeation behavior of a solute through the The resolution of solute diffusion into transverse and lateral
M. tuberculosis cell wall. For example, clofazimine and components in membranes might be very important in
thiocarlide have similar MWs. However, bulky-shaped studying ADMET properties of drug candidates. It is shown
clofazimine has a Dma value significantly lower than that of that lateral diffusion plays a significant role in the transport
the flat-shaped thiocarlide. Moreover, phencycline and clo- of molecules across tissues, including the blood-brain
fazimine have the same Dma values even though the MW of barrier, corneal membrane, and intestinal membranes (Figure
phencycline is only half that of clofazimine. 12).45 However, to get inside a target cell and exert its action,
A major finding of this MDS study is the observation for the drug candidate has to have good transverse diffusion
some solutes of predominant lateral diffusion within, as properties. Among the solute molecules currently studied,
opposed to transverse diffusion across, the M. tuberculosis hydnocarpic acid is the only molecule that displays pre-
cell wall model. Chaulmoogric acid is one of the solutes that dominant lateral diffusion in the bacterial and animal
prefers to move “parallel” to the cell surface inside the cell membrane models, while the rest of the solutes exhibit,
wall structure (lateral diffusion) as opposed to moving along essentially, equal lateral and transverse diffusion. Considering
the hydrocarbon chains of mycolates (transverse diffusion). that these solute molecules are actual drugs and have
As a result of preferred lateral diffusion, such solutes remain clinically acceptable ADMET properties, it is not clear from
inside the cell wall. Overall, the apparent penetration rate the results of this study what the delicate balance between
1076 Biomacromolecules, Vol. 5, No. 3, 2004 Hong and Hopfinger

Figure 12. Illustration of lateral diffusion and transverse diffusion


processes across three cells of a tissue. The green circles represent
the plasma membrane. The red arrows represent nonpolar solutes,
and the red circle stands for the lateral diffusion of the solute
molecules inside the membrane. The diffusion processes are driven
by the concentration gradient of the solute molecules.

lateral and transverse diffusion should be to realize an


effective drug for a particular therapy.
The molecular similarity analysis suggests that compounds
having markedly high features of molecular similarity will
display similar transport behavior in a common membrane/
cell wall environment. What is not clear is if a set of 13
diverse compounds is sufficient to extend the similar
structure-similar transport behavior observation into a design
rule. Moreover, there remains the nagging problem that, in Figure 13. Structure of the membrane-solute complex of the M.
tuberculosis cell wall with clofazimine at the end of MDS. Molecules
general, building in favorable ADMET properties in a drug are represented in the same manner as in Figure 10.
candidate correspondingly diminishes biological potency.
Dma is generally smaller than both DDPPE and DDMPC, fluid outer leaflet where the solute first enters the cell wall.
indicating it is more difficult for a solute to diffuse through This “pushing back” likely involves some kind of collabora-
a tightly packed M. tuberculosis cell wall inner leaflet than tive movement between the AG backbone and the mycolate
to pass through a more liquid-crystal-like phospholipid chains. As a result, the permeation rates of the solutes will
membrane structure. This observation is consistent with the be reduced. However, in the MDS experiments conducted
results from the M. tuberculosis cell wall construction study.25 in this study, the impact of cell wall structure deformation
However, the calculated variations in diffusion coefficients on solute diffusion processes may not be adequately modeled.
of the same solute in the different membrane/cell wall Another possible reason for the small variation in com-
environments and the computed diffusion coefficients of the puted diffusion coefficients may arise from neglect of the
different solutes in the same membrane/cell wall environment concentration gradients of the solute molecules in the MDS
are markedly small (1 order of magnitude) in terms of both experiments. That is, the solubility of the solute in the
the different membrane environments and the diversity in membrane (a component of permeation) has not been
chemical structures of the solutes. The role AG plays in considered. However, the variations in the computed diffu-
formation and stabilization of the M. tuberculosis cell wall sion coefficients, although small, are still consistent with the
may not have been adequately modeled and could be observed trends in the biological behavior of the different
responsible for this small range, overall, in the computed solute molecules in the different membrane/cell wall envi-
diffusion coefficients (Figures 2 and 3).25 When a bulky ronments. For example, the differences in the Dma values of
solute is inserted into the cell wall model, it introduces clofazimine, ethambutol, and chaulmoogric acid reveal quite
significant deformations to the structure of the M. tubercu- different behaviors among these solutes in the M. tuberculosis
losis model cell wall. For example, in the cell wall- cell wall model (Tables 3 and 4, Figure 11).
clofazimine complex, the PM molecules in the center of the In future work, a model for the entire mycolyl-AG
cell wall are “squeezed” out by clofazimine, and as a result, complex will be constructed and used to simulate the
the “surface” formed by the head groups of the PM molecules permeability of the M. tuberculosis cell wall. A means to
is no longer flat (Figure 13). The “chain reaction” induced include the concentration gradient of the solutes in the
by this structural deformation might disrupt the extensive membrane/cell wall (solute solubility in the membrane)
inter-residue and intra-residue hydrogen bonds among the should also be developed to actually represent permeation,
hydroxyl groups of arabinose and galactose residues of AG.25 as opposed to diffusion, processes for the solutes.
In turn, changes in this hydrogen bonding pattern might alter
the stability of the original mycolyl-AG complex structure. Acknowledgment. Resources of the Laboratory of Mo-
Consequently, the mycolyl-AG complex likely tries to keep lecular Modeling and Design were used in performing this
the solute molecule out and “push” it back to the relatively study. We gratefully appreciate financial support from The
Mycobacterium tuberculosis Cell Wall Permeability Biomacromolecules, Vol. 5, No. 3, 2004 1077

Chem21 Group, Inc., and X.H. acknowledges a University (22) McNeil, M.; Daffe, M.; Brennan, P. J. Evidence for the nature of
Fellowship from UIC. We also thank Mr. Jianzhong Liu of the link between the arabinogalactan and peptidoglycan of myco-
bacterial cell walls. J. Biol. Chem. 1991, 265, 18200-18206.
our research group and Professor Scott G. Franzblau of the (23) McNeil, M.; Daffe, M.; Brennan, P. J. Location of the mycolyl ester
Institute for Tuberculosis Research at UIC for their helpful substituents in the cell walls of mycobacteria. J. Biol. Chem. 1991,
comments and discussions over the course of this work. 266, 13217-13223.
(24) Liu, J.; Rosenberg, E. Y.; Nikaido, H. Fluidity of the lipid domain
References and Notes of cell wall from Mycobacterium chelonae. Proc. Natl. Acad. Sci.
U.S.A. 1995, 92, 11254-11258.
(1) WHO Tuberculosis Fact Sheet. http://www.who.int/mediacentre/
factsheets/who104/en/index.html (accessed May 2003). (25) Hong, X.; Hopfinger, A. J. Construction, Molecular Modeling, and
(2) WHO report 2003. http://www.who.int/gtb/publications/globrep/ Simulation of Mycobacterium tuberculosis Cell Walls. Biomacro-
index.html (accessed June 2003). molecules, 2004, 5, 1052-1065.
(3) Dolin, P. J.; Raviglione, M. C.; Kochi, A. Global tuberculosis (26) HyperChem Program Release 6.01 for Windows; Hypercube, Inc.:
incidence and mortality during 1990-2000. Bull. W. H. O. 1994, Gainesville, FL, 2000.
72, 213-220. (27) Polak, E. Computational Methods in Optimization; Academic Press:
(4) Jarlier, V.; Nikaido, H. Mycobacterial cell wall: structure and role New York, 1971.
in natural resistance to antibiotics. FEMS Microbiol. Lett. 1994, 123, (28) Dewar, M. J. S.; Zoebisch, E. G.; Healy, E. F.; Stewart, J. J. P.
11-18. AM1: a new general purpose quantum mechanical model. J. Am.
(5) Brennan, P. J.; Nikaido, H. The envelope of mycobacteria. Annu.
Chem. Soc. 1985, 107, 3902-3909.
ReV. Biochem. 1995, 64, 29-63.
(6) Barry, C. E., III; Mdluli, K. Drug sensitivity and environmental (29) Doherty, D. C. MOLSIM User’s Guide; The Chem21 Group, Inc.;
adaptation of mycobacterial cell wall components. Trends Microbiol. Lake Forest, IL, 1997.
1996, 4, 275-281. (30) Allinger, N. L. Conformational analysis. 130. MM2. A hydrocarbon
(7) Jarlier, V.; Nikaido, H. Permeability barrier to hydrophilic solutes force field utilizing V1 and V2 torsional terms. J. Am. Chem. Soc.
in Mycobacterium chelonei. J. Bacteriol. 1990, 172, 1418-1423. 1977, 99, 8127-8134.
(8) Trias, J.; Benz, R. Permeability of the cell wall of Mycobacterium (31) Hopfinger, A. J.; Pearlstein, R. A. Molecular mechanics force-field
smegmatis. Mol. Microbiol. 1994, 14, 283-290. parametrization precedures. J. Comput. Chem. 1984, 5, 486-492.
(9) Chambers, H. F.; Moreau, D.; Yajko, D.; Miick, C.; Wagner, C.; (32) Berendsen, H. J. C.; Postman, J. P. M.; Gunsteren, W. F. V.; Nola,
Hackbarth, C.; Kocagoz, S.; Rosenberg, E.; Hadley, W. K.; Nikaido, A. D.; Haak, J. R. Molecular dynamics with coupling to an external
H. Can penicillins and other beta-lactam antibiotics be used to treat
bath. J. Chem. Phys. 1984, 81, 3684-3690.
tuberculosis? Antimicrob. Agents Chemother. 1995, 39, 2620-2624.
(10) Jarlier, V.; Gutmann, L.; Nikaido, H. Interplay of cell wall barrier (33) Tinoco, I.; Sauer, K.; Wang, J. Physical Chemistry. Prentice-Hall:
and beta-lactamase activity determines high resistance to beta-lactam New York, 1985.
antibiotics in Mycobacterium chelonae. Antimicrob. Agents Chemoth- (34) Einstein, A. InVestigations on the theory of the Brownian moVement;
er. 1991, 35, 1937-1939. Methuen & Co., Ltd.: London, 1926.
(11) Minnikin, D. E. In The biology of the Mycobacteria; Ratledge, C., (35) MI-QSAR User’s Manual, version 2.0; The Chem21 Group, Inc.; Lake
Standford, J. L., Eds.; Academic: London, 1982; Vol. 1, pp 95- Forest, IL, 1997.
184. (36) Kulkarni, A.; Hopfinger, A. J. Membrane-interaction QSAR analy-
(12) Minnikin, D. E.; Goodfellow, M. In Microbiological classification sis: application to the estimation of eye irritation by organic
and identification; Goodfellow, M., Board, R. G., Eds.; Academic:
compounds. Pharm. Res. 1999, 16 (8), 1245-1253.
London, 1980; pp 189-256.
(13) Dobson, G.; Minnikin, D. E.; Minnikin, S. M.; Parlett, J. H.; (37) Kulkarni, A.; Han, Y.; Hopfinger, A. J. Predicting caco-2 cell
Goodfellow, M. In Chemical methods in bacterial systematics; permeation coefficients of organic molecules using membrane-
Goodfellow, M., Minnikin, D. E., Eds.; Academic: London, 1985; interaction QSAR analysis. J. Chem. Inf. Comput. Sci. 2002, 42, 331-
pp 237-265. 342.
(14) Brennan, P. J. In Microbial lipids; Ratledge, C., Wilkinson, S. G., (38) Iyer, M.; Mishru, R.; Han, Y.; Hopfinger, A. J. Predicting blood-
Eds.; Academic: London, 1988; Vol. 1, pp 203-298. brain barrier partitioning of organic molecules using membrane-
(15) Brennan, P. J. Structure of mycobacteria: recent developments in interaction QSAR analysis. Pharm. Res. 2002, 19 (11), 1611-1621.
defining cell wall carbohydrates and proteins. ReV. Infect. Dis. 1989, (39) Hopfinger, A. J.; Wang, S.; Tokarski, J. S.; Jin, B.; Albuquerque,
11 (Suppl.), 420-430. M.; Madhav, P. J.; Duraiswami, C. Construction of 3D-QSAR models
(16) Hunter, S. W.; Murphy, R. C.; Clay, K.; Goren, M. B.; Brennan, P.
using the 4D-QSAR analysis formalism. J. Am. Chem. Soc. 1997,
J. Trehalose-containing lipooligosaccharides. A new class of species-
specific antigens from mycobacterium. J. Biol. Chem. 1983, 258, 119, 10509-10524.
10481-10487. (40) Duca, J. S.; Hopfinger, A. J. Estimation of molecular similarity based
(17) Hunter, S. W.; Gaylord, H.; Brennan, P. J. Structure and antigenicity on 4D-QSAR analysis: formalism and validation. J. Chem. Inf.
of the phosphorylated lipopolysaccharide antigens from the leprosy Comput. Sci. 2001, 41, 1367-1387.
and tubercle bacilli. J. Biol. Chem. 1986, 261, 12345-12351. (41) Elson, E. L. Membrane dynamics studied by fluorescence correlation
(18) Liu, J.; Barry, C. E., III; Besra, G. S.; Nikaido, H. Mycolic acid spectroscopy and photobleaching recovery. Soc. Gen. Physiol. Ser.
structure determines the fluidity of the mycobacterial cell wall. J. 1986, 40, 367-383.
Biol. Chem. 1996, 271 (47), 29545-29551. (42) Frye, C. D.; Edidin, M. The rapid intermixing of cell surface antigens
(19) Yuan, Y.; Crane, D. C.; Musser, J. M.; Sreevatsan, S. Barry, C. E., after formation of mouse-human heterokaryons. J. Cell Sci. 1970, 7,
III. MMAS-1, the branch point between cis- and trans-cyclopropane-
319-335.
containing oxygenated mycolates in Mycobacterium tuberculosis. J.
Biol. Chem. 1997, 272 (15), 10041-10049. (43) Moore, P. B.; Lopez, C. F.; Klein, M. L. Dynamical properties of a
(20) Barry, C. E., III; Lee, R. E.; Mdluli, K.; Sampson, A. E.; Schroeder, hydrated lipid bilayer from a multinanosecond molecular dynamics
B. G.; Slayden, R. A.; Yuan, Y. Mycolic acids: structure, biosynthesis simulation. Biophys. J. 2001, 81, 2484-2494.
and physiological functions. Prog. Lipid Res. 1998, 37, 143-179. (44) Cabot, M. C.; Goucher, C. R. Chaulmoogric acid: assimilation into
(21) Daffe, M.; Brennan, P. J.; McNeil, M. Predominant structural features the complex lipids of mycobacteria. Lipids 1981, 16 (2), 146-148.
of the cell wall arabinogalactan of Mycobacterium tuberculosis as (45) Edwards, A.; Prausnitz, M. R. Predicted permeability of the cornea
revealed through characterization of oligoglycosyl alditol fragments to topical drugs. Pharm. Res. 2001, 18 (11), 1497-1508.
by gas chromatography/mass spectrometry and by 1H and 13C NMR
analysis. J. Biol. Chem. 1990, 265, 6734-6743. BM0345155

You might also like