ArticlePDF Available

Nanoconjugate Platforms Development Based in Poly(β,L-Malic Acid) Methyl Esters for Tumor Drug Delivery

Authors:

Abstract and Figures

New copolyesters derived from poly(β,L-malic acid) have been designed to serve as nanoconjugate platforms in drug delivery. 25% and 50% methylated derivatives (coPMLA-Me(25)H(75) and coPMLA-Me(50)H(50)) with absolute molecular weights of 32 600 Da and 33 100 Da, hydrodynamic diameters of 3.0 nm and 5.2 nm and zeta potential of -15mV and -8.25mV, respectively, were found to destabilize membranes of liposomes at pH 5.0 and pH 7.5 at concentrations above 0.05mg/mL. The copolymers were soluble in PBS (half life of 40 hours) and in human plasma (half life of 15 hours) but they showed tendency to aggregate at high levels of methylation. Fluorescence-labeled copolymers were internalized into MDA-MB-231 breast cancer cells with increased efficiency for the higher methylated copolymer. Viability of cultured brain and breast cancer cell lines indicated moderate toxicity that increased with methylation. The conclusion of the present work is that partially methylated poly(β,L-malic acid) copolyesters are suitable as nanoconjugate platforms for drug delivery.
Content may be subject to copyright.
Hindawi Publishing Corporation
Journal of Nanomaterials
Volume 2010, Article ID 825363, 8pages
doi:10.1155/2010/825363
Research Article
Nanoconjugate Platforms Development Based in
Poly(β,L-Malic Acid) Methyl Esters for Tumor Drug Delivery
Jos ´
ePortilla-Arias,
1Rameshwar Patil,1Jinwei Hu,1Hui Ding,1
Keith L. Black,1Montserrat Garc´
ıa-Alvarez,2Sebasti´
an Mu˜
noz-Guerra,2
Julia Y. Ljubimova,1and Eggehard Holler1, 3
1Department of Neurosurgery, Maxine Dunitz Neurosurgical Institute, Cedars-Sinai Medical Center, 8631 W. Third Street, Suite 800E,
Los Angeles, CA 90048, USA
2Departament d’Enginyeria Qu´
ımica, Universitat Polit`
ecnica de Catalunya, ETSEIB, Diagonal 647, 08028 Barcelona, Spain
3Institut f¨
ur Biophysik und Physikalische Biochemie, Universit¨
at Regensburg, D-93040 Regensburg, Germany
Correspondence should be addressed to Jos´
ePortilla-Arias,portillaj@cshs.org
Received 29 October 2009; Accepted 21 January 2010
Academic Editor: Chao Lin
Copyright © 2010 Jos´
e Portilla-Arias et al. This is an open access article distributed under the Creative Commons Attribution
License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly
cited.
New copolyesters derived from poly(β,L-malic acid) have been designed to serve as nanoconjugate platforms in drug delivery.
25% and 50% methylated derivatives (coPMLA-Me25H75 and coPMLA-Me50H50) with absolute molecular weights of 32 600 Da
and 33 100 Da, hydrodynamic diameters of 3.0nm and 5.2nm and zeta potential of 15 mV and 8.25 mV, respectively, were
found to destabilize membranes of liposomes at pH 5.0 and pH 7.5 at concentrations above 0.05mg/mL. The copolymers were
soluble in PBS (half life of 40 hours) and in human plasma (half life of 15 hours) but they showed tendency to aggregate at high
levels of methylation. Fluorescence-labeled copolymers were internalized into MDA-MB-231 breast cancer cells with increased
eciency for the higher methylated copolymer. Viability of cultured brain and breast cancer cell lines indicated moderate toxicity
that increased with methylation. The conclusion of the present work is that partially methylated poly(β,L-malic acid) copolyesters
are suitable as nanoconjugate platforms for drug delivery.
1. Introduction
Biodegradable polymers are suitable materials for the man-
ufacture of various devices which, because of their biocom-
patibility, are widely applied in medicine and pharmacology.
Examples are poly(lactic acid) and (PLA), poly(glycolic acid)
and their copolymers, with a successful 30 years history in
surgical settings [1].
The relatively new biopolymer, poly (β,L-malic acid)
(PMLA), is a carboxylic-functionalized polyester that can be
produced by either chemical synthesis or biological fermen-
tation from the slime mold Physarum polycephalum.Both
α-andβ-structures, either racemic or optically pure, may
be obtained by chemical methods whereas microorganisms
exclusively generate PMLA of extremely high optical purity
[2, and references therein]. The attractive properties in
nanobiotechnology and biomedicine are the lack of in vitro
and in vivo toxicity and non-immunogenicity. PMLA is
a completely biodegradable polymer that is metabolized
to water and carbon dioxide in the citric acid cycle. It is
biocompatible with regard to its limited stability in the
bloodstream thus prohibiting physiologically adverse host
responses after injection. It is chemically convenient to
handle due to its solubility in water and certain organic
solvents and to the reactivity of its pendant carboxylic groups
[29].
So far, the most advanced application reported for
PMLA, is the development of “Polycefin, a nanoconjugate
prototype vehicle for the cellular delivery of antisense
oligonucleotides. It consists of PMLA that harbors antisense
morpholino oligonucleotides and several biochemically
functional units which are chemically conjugated with the
carboxyl groups. These functional units are tumor-specific
antibodies that promote receptor-mediated endosomal
2Journal of Nanomaterials
uptake by the tumor cells, a unit that allows the vehicle
to escape from the endosomal vesicles into the cytoplasm
and another unit that, after cleavage, releases free oligonu-
cleotides. This nanoconjugate can carry several antisense
oligonucleotides or other drugs at the same time and also
simultaneously several antibodies, which can bind specifi-
cally to cell surface antigens during drug delivery [59].
Both synthetic and biosynthetic polymalic acids are
spontaneously or enzymatically degraded in aqueous envi-
ronment (see review [2]). Regardless chirality, degradation is
moderate at physiological pH and proceeds rapidly in acidic
(pH <5) and basic (pH >9) solutions [2] by random scission
of the main chain ester bonds to yield malic acid as the
final degradation product [10]. By chemical blocking of the
pendant carboxylic groups, the properties of PMLA could
be changed to slow down its hydrolysis rate. Methylation
with diazomethane has proven to be an ecient method
that produced poly(α-methyl β,L-malic acid) (PMLA-Me)
without significant cleavage of the polyester backbone [11,
12]. While low degrees of methylation resulted in water sol-
uble products, 75% methylation of pendant carboxyl groups
allowed formulation of stable, water insoluble nanoparticles
that could be loaded with proteins for delivery applications
[13].
The soluble PMLA methyl esters containing 25% and
50% of methylated carboxylic side groups (coPMLA-
Me25H75 and coPMLA-Me50H50 )arecharacterizedhereby
investigating their light scattering properties, their degra-
dation in human plasma and their cytotoxicity for several
human cancer cell lines to be used as nanoplatforms for drug
delivery.
2. Materials and Methods
2.1. PMLA Production. PMLA of microbial origin was used
in this work. It was obtained by cultivation of Physarum
polycephalum and subsequent purification as described in
detail elsewhere [14]. The polyacid of NMR purity had
aMw=34 300 Da and a polydispersity Mw/Mn=1.1.
All chemicals of highest purity, including human plasma,
were bought from Merck (Germany) and Sigma-Aldrich
(Germany). Organic solvents were of analytical grade and
used without further purification. Water used for buer
preparation was double distilled and deionized in a “Milli-
Q” system.
2.2. Esterification. Partial esterification of PMLA was per-
formed as described recently by Portilla-Arias et al. [11]. In
brief, a solution of diazomethane in ether (12.5 meq) was
added to a solution of PMLA in dry acetone (4.3 meq with
regard to malic acid units) in dierent ratios according to
the esterification degree to be obtained, and the mixture
was left under stirring at room temperature for 1 hour.
Thereactionmixturewasthenevaporatedundervacuum
and the residue was dissolved in a small amount of N-
methyl-2-pyrrolidone and precipitated with cold diethyl
ether. The copolymer was recovered by filtration as a white
powder. Yields for coPMLA-Me25H75 and coPMLA-Me50H50
were 97% and 92%, respectively. The 1H NMR analysis
in deuterated water revealed that the conversion actually
attained in the copolyesters was 20.2, 46.5, which are pretty
close to the nominal values.
2.3. Absolute Molecular Weight, Hydrodynamic Diameter and
Zeta Potential Measurements. The copolymers were charac-
terized with respect to their absolute molecular weight (Mw),
size and ζpotential using a Malvern Zetasizer Nano (Malvern
Instruments, UK). Absolute weight average molecular weight
was calculated with a modification of the Rayleigh equation
that can be used to generate a Debye plot, which is a linear
fit of KC/Rθversus concentration according to the equation
Kc/Rθ=1/Mw+2A2cRθis the Rayleigh ratio of scattered to
incident light intensity, Kis a constant defined by the solvent
and analyte dependent refractive index increment (dn/dc=
0.169 mL/g for PMLA), Avogadro’s number and the solvent
refractive index. cis the particle concentration and A2is the
second virial coecient [15]. The intercept obtained from
the Debye plot is equal to the inverse of the molecular weight
and the slope is twice the second virial coecient.
The size was calculated on the basis of noninvasive back-
scattering (NIBS) measurements using the Stokes-Einstein
equation, d(H) =kT/3πηD.d(H) is the hydrodynamic diam-
eter, Dthe translational diusion coecient, kBoltzmann’s
constant, Tabsolute temperature and ηthe viscosity. The
diameter that is measured in DLS (Dynamic Light Scattering)
refers to the particle diusion within a fluid and is referred
to as the hydrodynamic diameter corresponding to the
diameter of a sphere that has the same translational diusion
coecient as the particle [16]. The ζpotential was calculated
from the electrophoretic mobility based on the Helmholtz-
Smoluchowski formula, using electrophoresis M3-PALS. All
calculations were carried out by the Zetasizer 6.0 software.
For the molecular weight determination, 5 solutions of the
copolymers in phosphate buered saline (PBS, pH 7.4) were
generated by serial dilution starting with 4 mg/mL. For the
measurement of the ζpotential, the concentration of the
sample was 2 mg/mL dissolved in water containing 10 mM
NaCl, and the voltage applied was 150V. For the particle
size measurements, the solutions were prepared in PBS
at a concentration of 2 mg/mL, filtered through a 0.2 μm
pore membrane. All the copolymer solutions were prepared
immediately before analysis at 25C. Data represent the
mean ±standard deviation obtained for three measure-
ments.
2.4. Copolyesters Stability in PBS and Human Plasma. The
degradation essays in human plasma were carried out at
37C with a polymer concentration of 1 mg/mL. The sample
vials were sealed to avoid evaporation and stored at 37Cin
an incubator. For the isolation from the plasma, aliquots of
1 mL were extracted with 5 mL of chloroform/ethyl acetate
(1 : 1 v/v). The copolyester contained in the organic phase
was dried and redissolved in PBS and the Mwmeasured by
sec-HPLC (Calibrated with polystyrene sulphonate-sodium
salt standards). Sample preparation with the polymers of
known Mwverified that the isolation had no eect on
Journal of Nanomaterials 3
O
O
O
O
OO
O
O
Me H
xy
coPMLA-Me25H75
coPMLA-Me50H50
x=20.2; y=79.8
x=46.5; y=53.5
Figure 1: Chemical structure of copolyesters studied in the present
work with indication of their contents in methylated and free-
carboxyl malic units.
molecular weights. For comparison, the degradation study
was performed in PBS (pH 7.4) at a concentration of
1 mg/mL for each copolymer. Chromatography was per-
formed on a Hitachi analytical Elite LaChrom HPLC-
UV system and size exclusion column BioSep-SEC-S 3000
column (300 ×7.80 mm) following the elution at 220 nm
wavelength. Molecular weights Mw(t) were plotted as a
function of degradation time with reference to Mw(t=0) at
zero incubation time.
2.5. Cell Lines and Culture Conditions. Primary glioma cell
lines—U-87 MG and T98G—and invasive breast carci-
noma cell lines—MDA-MB-231 and MDA-MB-468—were
obtained from American Type Culture Collection (ATCC)
USA. U-87 MG and T98G cells were cultured in MEM
supplemented with the following ingredients (final concen-
trations): 10% fetal bovine serum, 1% MEM NEAA, 1 mM
sodium pyruvate and 2 mM L-glutamine. For MDA-MB-231
and MDA-MB-468, Leibovitz’s L-15 medium with 10% final
concentration fetal bovine serum was used. Cells were seeded
at 10,000 cells per well (0.1 mL) in 96-well flat-bottomed
plates and incubated overnight at 37C in humid atmosphere
with 5% CO2(breast cancer cell lines MDA-MB-231 and
MDA-MB-468 were incubated without CO2).
2.6. Cytotoxicity Test. The copolymers (1 mg/mL and serial
dilutions) were dissolved in culture media and incubated
with cells for 24 hours. Cell viability was measured using the
CellTiter 96 Aqueous One Solution Cell Proliferation Assay
kit (Promega Corporation, Cat. No.PR-G3580). Yellow [3-
(4,5-dimethylthiazol-2-yl)-5-(3-carboxymethoxyphenyl)-
2-(4-sulfophenyl)-2H-tetrazolium, inner salt] (MTS) is
bioreduced by cells into formazan that is soluble in the tissue
culture medium. The absorbance reading at 490 nm from
the 96-well plates was directly proportional to the number
of living cells [17]. The viability of the untreated cells was
referenced as 100%. The results shown are the means and
deviations standard of three independent measurements,
calculated with the statistical software GraphPad PRISM 3.0.
2.7. Liposome Leakage Assay. The capability of the copoly-
mers to escape from endosome was measured by the
liposome leakage assay. This method generally assumed to
represent main features of the endosome membrane and
Tab le 1: Physical properties of copolymers.
coPMLA-Me25 coPMLA-Me50
H75 H50
Methylation(a) (%) 20.2 46.5
Mw(b) 32,600 33,100
Mn(b) 26,100 24,300
Tm(c) (C) 174 172
Td(d) (C) 186 198
Microstructure(e)
Contiguous patches nfMe 3.9 11.8
Contiguous patches ngCOOH 11.5 14.5
R0.3 0.2
(a)Copolymer composition determined by 1H NMR; (b)Weight- and
number-average molecular weight measured by sec-HPLC; (c)Melting tem-
perature measured by dierential scanning calorimetry; (d) Onset decom-
position temperature measured at 5% of loss of initial weight; (e)nfand
ngrefer to averaged numbers of methylated and free carboxyl group
within homogeneous sequences, respectively, and (R) refers the randomness
determined by 13C NMR analysis [11].
0 20 40 60 80 100 120 140 160
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
PMLA
coPMLA-(Me25H75)coPMLA-(Me50 H50)coPMLA-(Me25H75)
coPMLA-(Me50H50)
Time (h)
Mw(t)/Mw(0)
Human plasma, 37C
PBS (pH 7.4), 37C
Figure 2: Degradation at 37CinPBS(pH7.4)andhumanplasma
for PMLA and the copolyesters.
gives similar results as the red blood cell lysis method,
because it is less biased by adverse eects of proteins con-
tained in erythrocytes/membranes. Liposome suspensions
were prepared by the extrusion method. Briefly, the mixture
of egg phosphatidylcholine and cholesterol (molar ratio 2 : 1)
dissolved in CHCl3/MeOH (v/v, 2 : 1) was dried under a
stream of nitrogen. The lipid mixture was hydrated with HBS
buer (5 mM HEPES, 150 mM NaCl, pH 7.4) containing
90 mM calcein, followed by 19 extrusions through 0.1 μm
polycarbonate membrane using mini-extruder (Avanti Polar
Lipids). Serial dilutions were carried out using 95 μLoftwo
buers of dierent pH, 137 mM HEPES buer pH 7.4 and
137 mM citrate buer pH 5.0. Liposome 5 μL (lipid concen-
tration 200 μM) was added to each sample and the plate was
incubated at room temperature for 1hour. Complete leakage
of calcein (100% reference) was achieved with the addition
of 0.25% (v/v) Triton-X100 solution of respective buers.
4Journal of Nanomaterials
8 9 10 11 12 13 14 15 16
ab
Retention time (min)
Absorbance
Figure 3: Elution profiles obtained by sec-HPLC for coPMLA-
Me50H50 at time zero (a) and after 24 hours of incubation
(b) in human plasma at 37C. Samples were extracted with
chloroform/ethylacetate before chromatography (see Section 2).
Fluorescence of released calcein release was measured using
excitation wavelength 488nm and an emission wavelength
535 nm. The results shown are the means and deviations
standard of three independent measurements, calculated
with the statistical software GraphPad PRISM 3.0.
2.8. Method of Fluorescence Labeling. The copolymers were
fluorescence labeled with rhodamine as follows: Pendant
carboxyl groups of copolymer (1 mmol malyl residues) in
1 mL of dimethyl formamide (DMF) were activated with
amixtureofN-hydroxysuccinimide (NHS, 1 mmol) and
dicyclohexylcarbodiimide (DCC, 1 mmol) dissolved in 2 ml
ofDMFatroomtemperaturefor3-4hoursundervigorous
stirring. 2-Mercapto-1-ethylamine (0.1 mmol) and 0.1 mmol
of dithiothreitol (DTT). were added and the reaction was
completed in 30–40 minutes. After 30 minutes of stirring
in 6 mL of water at room temperature, the mixture was
centrifuged and the clear supernatant passed over Sephadex
G10 columns in water. The product containing fractions
were freeze dryed yielding a white powder. Rhodamine Red
C2maleimide (40 μLof1mg/mLsolutioninDMF)was
added to 2 mg of activated copolymer dissolved in 2 mL of
PBS of pH 5.5 and stirred for 3 hours at room temperature.
The fluorescent polymers were purified over Sephadex PD-
10 columns pre-equilibrated with PBS (pH 7.4).
2.9. Fluorescence Microscopy. MDA-MB-231 cells were
seeded into the microscopic chamber slide. For fluorescence
microscopy study, cells were incubated with rhodamine-
labeled-polymer (1 mg/mL) in fetal bovine serum (FBS) free
medium for 3 hours. The stained cells were washed with PBS
and fixed in 4% paraformaldehyde (PFA) for 15 minutes.
Then the cells were counterstaining with DAPI to visualize
the nuclei and observed under a DM6000 Leica fluorescence
microscope (Wetzlar, Germany). The amount of polymer
0.00001 0.0001 0.001 0.01 0.1 1 10
0
5
10
15
20
25
30
coPMLA-(Me25H75)
coPMLA-(Me50H50)
Concentration (mg/mL)
Leakage (%)
pH 7.5
pH 5
Figure 4: Liposome leakage at pH 5.0 and 7.5 and 37C. 100%
leakage is obtained after the addition of 0.25% Triton X100.
uptake was calculated from the rhodamine fluorescence
intensity, with subtracted background and divided by the
area of the cell under study using the Image J1.43c software
from NIH (values averaged over >50 cells). The experiment
was done by triplicate, with three independent preparations
of cells. Mean and deviation standard were calculated with
the statistical software GraphPad PRISM 3.0.
3. Results and Discussion
3.1. Chemical Characterization. Partially methylated poly(β-
l-malic acid) copolymers were synthesized in high yield and
purity by the diazomethane in acetone method. The two
products are denoted as coPMLA-Me25H75 and coPMLA-
Me50H50 , their chemical formulae are depicted in Figure 1
and their properties are shown in Tabl e 1 .Thedegreeof
methylation attained in the two cases (20.2 and 46.5%) in
the reaction was close to the input ratio of diazomethane
to total carboxylic acid (25 and 50%). Molecular weights
of Mw(weight-average) =32,600 Da and 33,100 Da with
polydispersity indexes (Mw/Mn) of 1.3 and 1.1, respectively,
were determined by sec-HPLC for the two copolyesters which
are values comparable to those measured for the original
polymalic acid (Mw=34 300, Mw/Mn=1.1). The slightly
decrease in Mwand increase in polydispersity observed after
methylation suggest partial cleavage during the chemical syn-
thesis. The 13C-NMR analysis revealed that the distribution
of methyl groups along the copolyesters chain is not random.
The average number of contiguous methylated carboxylic
groups indicated hydrophobic patches as opposed to regions
with contiguous free carboxyl groups and interdispersed
free/methylated carboxyl groups. As expected, the length of
hydrophobic sequences was larger for the polymer with the
higher degree of methylation.
3.2. Absolute Molecular Weight, Hydrodynamic Diameter and
Zeta Potential. The values for absolute weight averaged
molecular weight, particle size and zeta potential of the
resulting copolyesters together with PMLA are shown in
Tabl e 2 . The values of molecular weight and polydispersity
Journal of Nanomaterials 5
Tab le 2: Light scattering and zeta potential measurements.
PMLA coPMLA-Me25H75 coPMLA-Me50H50
Molecular weight (Da) 34,200 30,100 31,900
Polydispersity index 1.1 1.22 1.41
2nd virial coecient A2(mL ·mol/g2) 6.50E-05 2.85E-05 3.50E-08
Hydrodynamic diameter (nm) 3.4 (±0.1) 3.0 (±0.1) 5.2 (±0.1)
Zeta potential (mV) 22.9 (±1.7) 15 (±1.1) 8.25 (±1.3)
0 200 400 600 800 1000
40
50
60
70
80
90
100
110
PMLA
coPMLA-(Me25H75)
coPMLA-(Me50H50)
Concentration (mg/mL)
U-87 MG
Viability (%)
(a)
0 200 400 600 800 1000 1200
40
50
60
70
80
90
100
110
PMLA
coPMLA-
(Me25H75 )
coPMLA-(Me50H50)
Concentration (mg/mL)
T98G
Viability (%)
(b)
0 200 400 600 800 1000
40
50
60
70
80
90
100
110 PMLA
coPMLA-(Me25H75)
coPMLA-(Me50H50)
Concentration (mg/mL)
MDA-MB-231
Viability (%)
(c)
0 200 400 600 800 1000 1200
40
50
60
70
80
90
100
110
PMLA
coPMLA-(Me25H75)
coPMLA-(Me50H50)
Concentration (mg/mL)
MDA-MB-468
Viability (%)
(d)
Figure 5: Cell viability of cultured cells after 24 hours incubation at 37Cfordierent cell lines in the presence of the indicated polyesters.
found by DLS are similar to those obtained by GPC (see
Tabl e 1 ) and corroborate the notion that some cleavage
occurred during synthesis. While the hydrodynamic diam-
eter does not follow a clear trend upon methylation, the
second virial coecient A2and the ζ-potential correlate well
with the degree of methylation. The second virial coecient
is a parameter describing the interaction strength between
the molecule and the solvent [16]. The relatively high values
obtained for PMLA and coPMLA-Me25H75 indicate the
ability of these polymers to stay in solution, whereas the
much lower value obtained for coPMLA-Me50H50 indicates
that this copolymer has some tendency for aggregation. The
conclusion is corroborated by the progressive decrease of
negative ζ-potential observed for increasing methylation,
that is, less repulsion and thus higher tendency for aggrega-
tion.
The eects of esterification on solubility are both an
increase in hydrophobicity and a decrease in electrostatic
mutual repulsion between polymers at higher degrees of
methylation. The blocky microstructure of the copolymer
harboring highly hydrophobic domains of contiguously
methylated units could contribute to aggregation by favoring
intermolecular hydrophobic contacts. The polyacid is highly
soluble in water and acetone, coPMLA-Me50H50 is less
soluble in water than coPMLA-Me25H75,andcoPMLA-
Me75H25 (not studied here) has been reported to be water
insoluble [11].
3.3. Copolyesters Degradation in PBS and Human Plasma.
Copolyester degradation was followed by measuring molecu-
lar weights Mwby sec-HPLC as a function of incubation time
(Figure 2). During incubation in PBS (pH 7.4) at 37C the
polymers were slowly degraded as indicated by an increase in
the retention time [11]. The measurements for degradation
in serum at 37C were complicated by the presence of
proteins which co-eluted and interfered with the polymers
in sec-HPLC, rendering polymer detection impossible. The
degraded polyesters except PMLA could be separated from
6Journal of Nanomaterials
proteins by extraction of the plasma mixture with chloro-
form/ethyl acetate. As an example the chromatograms of
coPMLA-Me50H50 incubated in serum at time zero (a) and
after incubation (b) are compared in Figure 3. This case
once more demonstrates the degradability of PMLA and
its derivatives in physiological conditions. Specific PMLA
hydrolases have been reported for microorganisms [18,19].
As in the case of degradation in PBS (not shown), the elution
profile indicated a single peak. This suggested cleavage
from the ends and not fragmentation at internal cleavage
sites.
Comparison of kinetics in Figure 2 revealed that (i)
degradation in human plasma was faster than in PBS with
half lives of 13 h in plasma for coPMLA-Me25H75 and 20
hours for coPMLA-Me50H50 compared with 45 hours and
50 hours in PBS, respectively and (ii) Half life was lowest
for PMLA and increased with higher levels of methylation in
agreement with the notion that alkylation of the α-carboxyl
group stabilized the main chain ester bond. Hydrolysis was
significantly enhanced by constituents of human plasma,
either by general catalysis or by hydrolytic enzymes, most
likely esterases, such as plasma lipases or cholinesterases [20].
3.4. Membrane Disruption. The membrane disruption activ-
ity of coPMLA-Me25H75 and coPMLA-Me50H50 was mea-
sured by the phosphatidylcholine liposome leakage assay.
The liposomes are filled with calcein, which leaks out if
the liposome becomes destabilized or disrupted. Leakage
was measured at pH 7.5, resembling physiological pH, and
at pH 5.0, resembling pH of late endosomes/lysosomes.
While coPMLA-Me25H75 did not show leakage activity,
coPMLA-Me50H50 was active at concentrations above 0.1
mg/mL and the activity was pH independent (Figure 4).
The finding suggested that membrane leakage was inferred
by methylation of pendant carboxyl groups that increased
hydrophobicity and neutralized negative charges (decrease
in ζ-potential, Tabl e 2). The absence of eect of pH change
indicated that the carboxylic groups of the methylated PMLA
did not protonate in this pH range, or if they did, they had
no eect on the leakage activity.
The role of hydrophobicity together with charge neutral-
ization has been considered in drug delivery to be the mech-
anism for membrane disruption by a variety of molecular
devices [2125]. In the case of our methylated PMLA we
think that the methylation dependent leakage refers mainly
to the formation of the contiguous, electrically neutral
hydrophobic patches (Tab le 2) which are prone to intrude
into the lipid bilayer and cause the membrane damage. The
membrane disruption activity especially coPMLA-Me50H50
is thought to be useful in the design of nanoconjugates that
can deliver drugs to intracellular targets.
3.5. Cytotoxicity. Partially methylated PMLA contains car-
boxylic groups that can be conjugated to several prodrugs
and in addition to a variety of biologically active units such as
antibodies for targeting or PEG for protection against enzy-
matic degradation and resorption by the reticuloendothelial
system (RES). The pro-drug is activated within the targeted
cell compartment and only then unfolds its cytotoxic or any
other activity.
It is desirable to know whether methylated PMLA as the
platform is itself not toxic. To this end, the in vitro toxicity
of the copolymers was tested. Figure 5 shows the viability
of cultured brain and breast cancer cells: T98G, U-87 MG,
MDA-MB-231 and MDA-MB-468 cells after 24 hours of
incubation as a function of copolymer concentrations.
While PMLA only marginally decreased cell viability, the
copolymers showed toxicity that increased with the degree
of methylation, but the decrease in percentage cell viability
was moderate and above 50% at concentrations 1mg/mL.
Eects on cell viability depended also on the type of cell line,
glioma U-87 MG cells, and breast cancer MDA-MB-231 cells
being more aected than glioma T98G cells and breast cancer
MDA-MB-468 cells.
There are two main toxicity mechanisms commonly
considered: (i) toxicity due to physical damage such as
destabilization of membranes and (ii) toxicity resulting from
the degradation products after intracellular uptake. The
possibility of physical damage due to membrane disrupture
of the kind as seen by the liposome leakage assay (Figure 5)
may not be significant since the eect of copolymers on cell
viability is not manifested in the time scale of minutes or a
few hours (results not shown). It is highly likely that toxicity
was the result of methanol formation during degradation. It
is known that degradation in pure deuterated water generates
as main products methanol and L-malic acid [11]. While
l-malic acid is converted into water and carbon dioxide
in the tricarboxylic acid cycle, methanol is known to be
toxic for living cells. In the same line it has been reported
that benzylesters of PMLA were toxic due to release of
free benzylalcohol during degradation [26]. However, for
consideration of the copolymer application in drug delivery,
the short residence times of only a few hours before complete
clearance through the renal system relatives their toxicity in
vivo.
3.6. Cellular Uptake Study. Since we have found that
the copolymers coPMLA-Me25H75 and coPMLA-Me50H50
contained hydrophobic patches that could be active in
membrane disruption, it was important to test whether
the polymers could enter cells in vitro. Rhodamine labeled
copolymers were incubated with MDA-MB-231 cells for
3 hours. The cell’s uptake of the fluorescent polymers
was seen under the fluorescence microscope (Figure 6).
Rhodamine alone did not stain the cells (picture not
shown). The distribution of copolymers-rhodamine in all
cells, appeared to be homogeneous and probably involved
nuclei. The fluorescence intensity was semi quantitatively
measured by triplicate in >50cells,andappearedtobe
1.8 (Standard Deviation=±0.3) folds more intensive for
coPMLA-Me50H50 than for coPMLA-Me25H75 accordingly
to the Image J1.43c software, thus correlating with the
higher degree of hydrophobicity of the higher methylated
polymer. The results are evidences that the copolymers
are highly versatile materials and suitable for various
applications.
Journal of Nanomaterials 7
coPMLA-(Me50H50)coPMLA-(Me25H75)
(a) (b) (c)
Figure 6: Localization of rhodamine tagged copolymers in MDA-MB-231 cells. Panels (a) and (b) separated copolymer-rhodamine uptake
and DAPI staining. Panel (c) The copolymers are indicated by red fluorescence co-localizing with nuclei stained in blue with DAPI.
4. Conclusion
New designs of nanoconjugate drug delivery systems are
introduced here. The systems involve a polymer platform
containing pendant chemically reactive groups to be con-
jugated with molecular units that function in pro-drug
attachment, cell recognition, membrane penetration and
protection. The copolymers coPMLA-Me25H75 and coPMLA-
Me50H50 are biodegradable and biocompatible and its half-
life is limited, so they may minimize adverse host’s responses
and development of liver storage diseases. The copolymers
are endowed with membrane disrupting/penetrating activi-
ties which allow them to deliver drugs directly to intracellular
targets by passing the plasma membrane. By raising the
degree of methylation above 50%, the copolymers become
insoluble and can be used as drug delivering nanoparticles.
These results support the fact that poly (β,L-malic acid) is
a highly versatile material and can be used for the design
of a variety of nanoconjugate platforms for drug delivery
systems.
Acknowledgments
The project was funded by grants from NIH/NCI R01
CA123495 to JYL; MAT2006-13209-C02-02 from CICYT
(Comisi´
on Interministerial de Ciencia y Tecnolog´
ıa) of Spain
to SM and CONACYT-M´
exico (Concejo Nacional de Ciencia
y Tecnolog´
ıa) for fellow granted to J. Portilla-Arias.
References
[1] D.Peer,J.M.Karp,S.Hong,O.C.Farokhzad,R.Margalit,and
R. Langer, “Nanocarriers as an emerging platform for cancer
therapy,Nature Nanotechnology, vol. 2, no. 12, pp. 751–760,
2007.
[2] B.-S. Lee, M. Vert, and E. Holler, “Water-soluble aliphatic
polyesters: poly(malic acid)s,” in Biopolymers,Y.Doiand
A. Steinb¨
uchel, Eds., pp. 75–103, Wiley-VCH, Weinheim,
Germany, 2002.
[3] B. Gasslmaier, C. M. Krell, D. Seebach, and E. Holler, “Syn-
thetic substrates and inhibitors of β-poly(L-malate)-hydrolase
(polymalatase),European Journal of Biochemistry, vol. 267,
no. 16, pp. 5101–5105, 2000.
[4] B. Gasslmaier and E. Holler, “Specificity and direction of
depolymerization of β-poly(L-malate) catalysed by poly-
malatase from Physarum polycephalum—fluorescence label-
ing at the carboxy-terminus of β-poly(L-malate),European
Journal of Biochemistry, vol. 250, no. 2, pp. 308–314, 1997.
[5] B.-S. Lee, M. Fujita, N. M. Khazenzon, et al., “Polycefin,
a new prototype of a multifunctional nanoconjugate based
on poly(β-L-malic acid) for drug delivery,Bioconjugate
Chemistry, vol. 17, no. 2, pp. 317–326, 2006.
[6] M. Fujita, N. M. Khazenzon, A. V. Ljubimov, et al., “Inhibition
of laminin-8 in vivo using a novel poly(malic acid)-based
carrier reduces glioma angiogenesis,” Angiogenesis, vol. 9, no.
4, pp. 183–191, 2006.
[7] M. Fujita, B.-S. Lee, N. M. Khazenzon, et al., “Brain
tumor tandem targeting using a combination of monoclonal
antibodies attached to biopoly(β-L-malic acid),Journal of
Controlled Release, vol. 122, no. 3, pp. 356–363, 2007.
8Journal of Nanomaterials
[8] J. Y. Ljubimova, M. Fujita, N. M. Khazenzon, et al., “Nanocon-
jugate based on polymalic acid for tumor targeting,Chemico-
Biological Interactions, vol. 171, no. 2, pp. 195–203, 2008.
[9] J. Y. Ljubimova, M. Fujita, A. V. Ljubimov, V. P. Torchilin,
K. L. Black, and E. Holler, “Poly(malic acid) nanoconjugates
containing various antibodies and oligonucleotides for multi-
targeting drug delivery,Nanomedicine, vol. 3, no. 2, pp. 247–
265, 2008.
[10] C. Braud and M. Vert, “Degradation of poly(β-malic acid)-
monitoring of oligomers formation by aqueous SEC and
HPCE,Polymer Bulletin, vol. 29, no. 1-2, pp. 177–183, 1992.
[11] J. A. Portilla-Arias, M. Garc´
ıa- ´
Alvarez, A. M. de Ilarduya,
E. Holler, J. A. Galbis, and S. Mu˜
noz-Guerra, “Synthesis,
degradability, and drug releasing properties of methyl esters of
fungal poly(β, L-malic acid),Macromolecular Bioscience, vol.
8, no. 6, pp. 540–550, 2008.
[12] C.E.Fernandez,M.Mancera,E.Holler,J.J.Bou,J.A.Galbis,
and S. Mu˜
noz-Guerra, “Low-molecular-weight poly(α-methyl
β,L-malate) of microbial origin: synthesis and crystallization,
Macromolecular Bioscience, vol. 5, no. 2, pp. 172–176, 2005.
[13] J. A. Portilla-Arias, M. Garc´
ıa- ´
Alvarez, J. A. Galbis, and
S. Mu˜
noz-Guerra, “Biodegradable nanoparticles of partially
methylated fungal poly(β-L-malic acid) as a novel protein
delivery carrier,Macromolecular Bioscience,vol.8,no.6,pp.
551–559, 2008.
[14] E. Holler, Handbook of Engineering Polymeric Materials, vol.
997, Marcel Dekker, New York, NY, USA, 1997.
[15] P. C. Hiemenz, “Light scattering by polymer solutions,” in
Polymer Chemistry: The Basic Concepts, p. 659, Marcel Decker,
New York, NY, USA, 1984.
[16] (ISO), I. O. f. S., “Methods for Determination of Particle
Size Distribution Part 8: Photon Correlation Spectroscopy,
International Standard ISO13321, 1996.
[17] T. Mosmann, “Rapid colorimetric assay for cellular growth
and survival: application to proliferation and cytotoxicity
assays,Journal of Immunological Methods,vol.65,no.1-2,pp.
55–63, 1983.
[18] C. Korherr, M. Roth, and E. Holler, “Poly(β-L-malate) hydro-
lase from plasmodia of Physarum polycephalum, Canadian
Journal of Microbiology, vol. 41, supplement 1, pp. 192–199,
1995.
[19] K. Rathberger, H. Reisner, B. Willibald, H.-P. Molitoris, and
E. Holler, “Comparative synthesis and hydrolytic degradation
of poly (L-malate) by myxomycetes and fungi,Mycological
Research, vol. 103, no. 5, pp. 513–520, 1999.
[20] F. M. Williams, “Clinical significance of esterases in man,
Clinical Pharmacokinetics, vol. 10, no. 5, pp. 392–403, 1985.
[21] S. R. Tonge and B. J. Tighe, “Responsive hydrophobically
associating polymers: a review of structure and properties,”
Advanced Drug Delivery Reviews, vol. 53, no. 1, pp. 109–122,
2001.
[22] C. Kusonwiriyawong, P. van de Wetering, J. A. Hubbell,
H. P. Merkle, and E. Walter, “Evaluation of pH-dependent
membrane-disruptive properties of poly(acrylic acid) derived
polymers,European Journal of Pharmaceutics and Biopharma-
ceutics, vol. 56, no. 2, pp. 237–246, 2003.
[23] M.-A. Yessine, M. Lafleur, C. Meier, H.-U. Petereit, and J.-
C. Leroux, “Characterization of the membrane-destabilizing
properties of dierent pH-sensitive methacrylic acid copoly-
mers,Biochimica et Biophysica Acta, vol. 1613, no. 1-2, pp.
28–38, 2003.
[24] M.-A. Yessine and J.-C. Leroux, “Membrane-destabilizing
polyanions: interaction with lipid bilayers and endoso-
mal escape of biomacromolecules,Advanced Drug Delivery
Reviews, vol. 56, no. 7, pp. 999–1021, 2004.
[25]R.Chen,S.Khormaee,M.E.Eccleston,andN.K.H.
Slater, “The role of hydrophobic amino acid grafts in
the enhancement of membrane-disruptive activity of pH-
responsive pseudo-peptides,Biomaterials, vol. 30, no. 10, pp.
1954–1961, 2009.
[26] M. E. Martinez Barbosa, S. Cammas, M. Appel, and G.
Ponchel, “Investigation of the degradation mechanisms of
poly(malic acid) esters in vitro and their related cytotoxicities
on J774 macrophages,Biomacromolecules,vol.5,no.1,pp.
137–143, 2004.
... In view of its hydrophilicity and negative charge at physiological pH, PMLA is a polyanion that has little affinity for negatively charged lipid bilayers and does not translocate through cell membranes. To increase the interaction between the biopolymer and the plasma membrane, methylation of carboxylic acid groups with different levels of diazomethane was used to generate a PMLA-Me x H 100−x copolymer (where x is the percentage of methyl units) [77,78]. As x increases, the hydrophobicity of the molecule increases in the order PMLA < PMLA-Me 25 H 75 < PMLA-Me 50 H 50 < PMLA-Me 75 H 25 < PMLA-Me. ...
... Both PMLA-Me 75 H 25 and PMLA-Me were completely insoluble in water, similarly to PMLA benzyl esterification. The same order was observed for hydrolysis in saline and plasma, rupture of liposome membranes, and cytotoxicity [77]. In addition, hydrophobic amino acids or peptides can be conjugated to PMLA side-chain carboxylic acids to modulate its hydrophilicity and net charge, thus tuning the interplay with cellular/subcellular membranes in a pH-responsive manner. ...
Article
Poly(β-L-malic acid) (PMLA) is a natural polyester produced by numerous microorganisms. Regarding its biosynthetic machinery, a nonribosomal peptide synthetase (NRPS) is proposed to direct polymerization of L-malic acid in vivo. Chemically versatile and biologically compatible, PMLA can be used as an ideal carrier for several molecules, including nucleotides, proteins, chemotherapeutic drugs, and imaging agents, and can deliver multimodal theranostics through biological barriers such as the blood–brain barrier. We focus on PMLA biosynthesis in microorganisms, summarize the physicochemical and physiochemical characteristics of PMLA as a naturally derived polymeric delivery platform at nanoscale, and highlight the attachment of functional groups to enhance cancer detection and treatment.
... Polymalate is a secondary metabolite produced by A. pullulans. A polymer consisting of L-malic acid monomers that can potentially be used in drug delivery or tissue scaffolding (Portilla-Arias et al., 2010;Qiang et al., 2012;Kövilein et al., 2020). A. pullulans also belongs to the melanin-producing fungi and is thus considered a black yeast (Campana et al., 2022). ...
Article
Full-text available
Polyol lipids (a.k.a. liamocins) produced by the polyextremotolerant, yeast-like fungus Aureobasidium pullulans are amphiphilic molecules with high potential to serve as biosurfactants. So far, cultivations of A. pullulans have been performed in media with complex components, which complicates further process optimization due to their undefined composition. In this study, we developed and optimized a minimal medium, focusing on biosurfactant production. Firstly, we replaced yeast extract and peptone in the best-performing polyol lipid production medium to date with a vitamin solution, a trace-element solution, and a nitrogen source. We employed a design of experiments approach with a factor screening using a two-level-factorial design, followed by a central composite design. The polyol lipid titer was increased by 56% to 48 g L ⁻¹ , and the space-time yield from 0.13 to 0.20 g L ⁻¹ h ⁻¹ in microtiter plate cultivations. This was followed by a successful transfer to a 1 L bioreactor, reaching a polyol lipid concentration of 41 g L ⁻¹ . The final minimal medium allows the investigation of alternative carbon sources and the metabolic pathways involved, to pinpoint targets for genetic modifications. The results are discussed in the context of the industrial applicability of this robust and versatile fungus.
... PMA has attracted attention from researchers for its potential use in biomedical applications as a result of its good biocompatibility, and biodegradability and lack of toxicity and nonimmunogenic (Portilla-Arias et al. 2008). PMA, with multiple free carboxyl groups, can be used to chemically bind various molecules of interest, such as drugs and targeting moieties (Francis et al. 2017;Portilla-Arias et al. 2010). Moreover, PMA can easily chelate with positive inions, which has potential applications in environmental research fields. ...
Article
Full-text available
Polymalic acid (PMA) is a biodegradable polymer produced by the polyextremotolerant fungi Aureobasidium pullulans and has been shown to have potential applications in environmental fields. In this work, a high PMA yield mutant FJ-D2 was screened from T-DNA-based mutant libraries and showed a 12.9% increase in PMA titers, which was attributed to decreased the expression of a glycosyltransferase gene (celA), resulting in a 39.5% reduction in cellulose biosynthesis. Untreated waste xylose mother liquor (WXML), an environmental waste generated from the xylitol industry, can be directly used as an economical substrate for PMA production. Using batch-fermentation of FJ-D2, the PMA titer of 57.1 ± 0.02 g/L was produced in a 5-L fermentor, with the highest MA yield of 0.77 g/g mixed sugar. Furthermore, compared with ethylenediaminetetraacetic acid (EDTA), PMA had a comparable cadmium (Cd) removal efficiency (88.7% for EDTA versus 86.0% for PMA), which was not found in the monomer of L-malic acid (MA) monomers. These findings indicated that PMA was an environmentally friendly and biodegradable chelator for soil remediation. Moreover, our results provided an economically competitive process for PMA production from renewable environmental wastes.
... Due to its metabolic flexibility, A. pullulans is a useful source for a number of bioproducts . This fungus is, for example, capable of producing polymalic acid, a molecule discussed to be used as a new functional material for drug delivery, tissue scaffolds, and food materials (Portilla-Arias et al. 2010;Qiang et al. 2012). Similar to pullulan, polymalic acid is also produced as a secondary metabolite, preferentially under nitrogen limitation (Cheng et al. 2017;Wang et al. 2016). ...
Article
Full-text available
Liamocins are biosurfactants produced by the fungus Aureobasidium pullulans. A. pullulans belongs to the black yeasts and is known for its ability to produce pullulan and melanin. However, the production of liamocins has not been investigated intensively. Initially, HPLC methods for the quantification of liamocin and the identification of liamocin congeners were established. Eleven congeners could be detected, differing in the polyol head groups arabitol or mannitol. In addition, headless molecules, so-called exophilins, were also identified. The HPLC method reported here allows quick and reliable quantification of all identified congeners, an often-overlooked prerequisite for the investigation of valuable product formation. Liamocin synthesis was optimized during cultivation in lab-scale fermenters. While the pH can be kept constant, the best strategy for liamocin synthesis consists of a growth phase at neutral pH and a subsequent production phase induced by a manual pH shift to pH 3.5. Finally, combining increased nitrogen availability with a pulsed fed-batch fermentation, cell growth, and liamocin titers could be enhanced. Here, the maximal titers of above 10 g/L that were reached are the highest reported to date for liamocin synthesis using A. pullulans in lab-scale fermenters.
... PMLA as a carrier matrix lacks toxicity in vitro and in vivo, is characterized by nonimmunogenicity, biodegradability, and versatility for drug loading (Ljubimova et al., 2008a). With these properties, nanocomposites based on PMLA demonstrated inhibition activities relating to tumor angiogenesis (Ljubimova et al., 2008a) and tumor targeting of human breast cancer cells (Inoue et al., 2011(Inoue et al., , 2012 and human glioma cells Ding et al., 2013;Lee et al., 2006;Ljubimova et al., 2008bLjubimova et al., , 2013Portilla-Arias et al., 2010). In addition, methanol extracts from the plasmodia and sclerotia of F. septica have been shown to display an inhibitory effect on the B16 F1 murine melanoma cells (Jiang, 2013). ...
Chapter
The life cycle of myxomycetes involves three morphologically distinct stages. These are the amoeboflagellate stage (which consists of naked amoebae or flagellated cells), the plasmodium, and the fruiting body containing spores. The biology of myxomycetes has been studied for more than 300 years, and the life cycle of more than 100 species of myxomycetes has been examined in hanging-drop cultures, moist chamber cultures and agar cultures. As a result, a considerable body of knowledge now exists relating to the physiological characteristics of the three stages. More recently, various metabolites obtained from myxomycetes and the bioactivities of these metabolites have been subjected to study. In this chapter, a preliminary understanding of the chemical composition and structure of examples of metabolites from myxomycetes will be discussed, and the relationships that appear to exist between biological activity, physiological function, and their specific structure described. Some of those metabolites have been demonstrated to exhibit antibacterial activity, antitumor activity, cytotoxic activity, and antioxidant activity.
Article
Full-text available
With rich carboxyl groups in the side chain, biodegradable polymalic acid (PMLA) is an ideal delivery platform for multifunctional purposes, including imaging diagnosis and targeting therapy. This polymeric material can be obtained via chemical synthesis, or biological production where L-malic acids are polymerized in the presence of PMLA synthetase inside a variety of microorganisms. Fermentative methods have been employed to produce PMLAs from biological sources, and analytical assessments have been established to characterize this natural biopolymer. Further functionalized, PMLA serves as a versatile carrier of pharmaceutically active molecules at nano scale. In this review, we first delineate biosynthesis of PMLA in different microorganisms and compare with its chemical synthesis. We then introduce the biodegradation mechanism PMLA, its upscaled bioproduction together with characterization. After discussing advantages and disadvantages of PMLA as a suitable delivery carrier, and strategies used to functionalize PMLA for disease diagnosis and therapy, we finally summarize the current challenges in the biomedical applications of PMLA and envisage the future role of PMLA in clinical nanomedicine. Graphical Abstract
Chapter
The life cycle of myxomycetes involves three morphologically distinct stages. These are the amoeboflagellate stage (which consists of naked amoebae or flagellated cells), the plasmodium, and the fruiting body containing spores. The biology of myxomycetes has been studied for more than 300 years, and the life cycle of more than 100 species of myxomycetes has been examined in hanging-drop cultures, moist chamber cultures, and agar cultures. As a result, a considerable body of knowledge now exists relating to the physiological characteristics of the three stages. More recently, various metabolites obtained from myxomycetes and the bioactivities of these metabolites have been subjected to study. In this chapter a preliminary understanding of the chemical composition and structure of examples of metabolites from myxomycetes will be discussed, and the relationships that appear to exist between biological activity, physiological function, and their specific structure described. Some of those metabolites have been demonstrated to exhibit antibacterial activity, antitumor activity, cytotoxic activity, and antioxidant activity.
Article
Polymalic acid (PMA), a homopolymer of L-malic acid (MA) generated from a yeast-like fungus Aureobasidium pullulans, has unique properties and many applications in food, biomedical, and environmental fields. Acid hydrolysis of PMA, releasing the monomer MA, has become a novel process for the production of bio-based MA, which currently is produced by chemical synthesis using petroleum-derived feedstocks. Recently, current researches attempted to develop economically competitive process for PMA and MA production from renewable biomass feedstocks. Compared to lignocellulosic biomass, PMA and MA production from low-value food processing wastes or by-products, generated from corn, sugarcane, or soybean refinery industries, showed more economical and sustainable for developing a MA derivatives platform from biomass biorefinery to chemical conversion. In the review, we compared the process feasibility for PMA fermentation with lignocellulosic biomass and food process wastes. Some useful strategies for metabolic engineering are summarized. Its changeable applicability and future prospects in food and biomedical fields are also discussed.
Article
Polymer nanomedicines (macromolecular therapeutics, polymer-drug conjugates, drug-free macromolecular therapeutics) are a group of biologically active compounds that are characterized by their large molecular weight. This review focuses on bioconjugates of water-soluble macromolecules with low molecular weight drugs and selected proteins. After analyzing the design principles, different structures of polymer carriers are discussed followed by the examination of the efficacy of the conjugates in animal models and challenges for their translation into the clinic. Two innovative directions in macromolecular therapeutics that depend on receptor crosslinking are highlighted: a) Combination chemotherapy of backbone degradable polymer-drug conjugates with immune checkpoint blockade by multivalent polymer peptide antagonists; and b) Drug-free macromolecular therapeutics, a new paradigm in drug delivery.
Article
Some microorganisms naturally produce β-poly(l-malic acid) (PMA), which has excellent water solubility, biodegradability, and biocompatibility properties. PMA has broad prospective applications as novel biopolymeric materials and carriers in the drug, food, and biomedical fields. Malic acid, a four-carbon dicarboxylic acid, is widely used in foods and pharmaceuticals, as a platform chemical. Currently, malic acid produced through chemical synthesis and is available as a racemic mixture of l- and d-forms. The d-form malic acid exhibits safety concerns for human consumption. There is extensive interest to develop economical bioprocesses for l-malic acid and PMA production from renewable biomass feedstocks. In this review, we focus on PMA biosynthesis by Aureobasidium pullulans, a black yeast with a large genome containing genes encoding many hydrolases capable of degrading various plant materials. The metabolic and regulatory pathways for PMA biosynthesis, metabolic engineering strategies for strain development, process factors affecting fermentation kinetics and PMA production, and downstream processing for PMA recovery and purification are discussed. Prospects of microbial PMA and malic acid production are also considered.
Article
Full-text available
Nanotechnology has the potential to revolutionize cancer diagnosis and therapy. Advances in protein engineering and materials science have contributed to novel nanoscale targeting approaches that may bring new hope to cancer patients. Several therapeutic nanocarriers have been approved for clinical use. However, to date, there are only a few clinically approved nanocarriers that incorporate molecules to selectively bind and target cancer cells. This review examines some of the approved formulations and discusses the challenges in translating basic research to the clinic. We detail the arsenal of nanocarriers and molecules available for selective tumour targeting, and emphasize the challenges in cancer treatment.
Article
Polymalatase from Physarum polycephalum calalysed the hydrolysis of β-poly[l-malate] and of the synthetic compounds β-di(l-malate), β-tetra(l-malate), β-tetra(l-malate) β-propylester, andl-malate β-methylester. Cyclic β-tri(l-malate), cyclic β-tetra(l-malate), andd-malate β-methylester were not cleaved, but were competitive inhibitors. The O-terminal acetate of β-tetra(l-malate) was neither a substrate nor an inhibitor.l-Malate was liberated; the Km,Ki and Vmax values were measured. The appearance of comparable amounts of β-tri(l-malate), and β-di(l-malate) during the cleavage of β-tetra(l-malate) indicated a distributive mechanism for small substrates. The accumulation of a series of oligomers, peaking with the 11-mer and 12-mer in the absence of higher intermediates, indicated that the depolymerization of β-poly(l-malate) was processive. The results indicate that β-poly(l-malate) is anchored at its OH-terminus by the highly specific binding of the penultimate malyl residue. The malyl moieties beyond 12 residues downstream from the OH-terminus extend into a diffuse second, electrostatic binding site. The catalytic site joins the first binding site, accounting for the cleavage of the polymer into malate residues. It is proposed that the enzyme does not dissociate from β-poly(l-malate) during hydrolysis, when both sites are filled with the polymer. When only the first binding site is filled, the reaction partitions at each oligomer between hydrolysis and dissociation.
Article
A 68-kDa extracellular glycoprotein from Physarum polycephalum that hydrolyses specifically poly(β-L-malic acid) by removing monomers of L-malic acid in an exolytic manner has been purified and characterized. The enzyme was purified 1740-fold from the culture medium by ammonium sulfate precipitation, hydrophobic interaction chromatography on butyl-Toyopearl, and gel permeation chromatography on Superdex 200 to a specific activity of 9.0 μmol∙min−1∙mg−1. The hydrolase was also purified from the cytosol, which contained 1 mg in 43 g cells in contrast to 1 mg extracellular enzyme in 28 L of culture medium. The pH optimum was pH 3.5 as a result of the effect of an acidic side chain on Vmax and the preferred binding of poly(β-L-malate) in the ionized form. Intracellular hydrolase was only marginally active on [14C]poly(β-L-malate) that had been injected into plasmodia. Poly(L-aspartate), poly(L-glutamate), poly(vinyl sulfate), and poly(acrylate) were neither bound nor degraded by the hydrolase. Poly(β-hydroxybutyric acid), which was considered the reduced form of poly(β-L-malate), was not a substrate. The enzyme is neither a metallo- nor a serine-esterase, and is distinct from poly(3-hydroxybutyric acid) depolymerases. It is related to a glucosidase with respect to hydrophobic interaction chromatography, the pH-activity dependence, and its inhibition with mercuribenzoate, N-bromosuccinimide, and D-gluconolactone, but not the use of the substrates.Key words: poly(β-L-malate), polymalatase, Physarum polycephalum, biodegradative polymer.
Chapter
Introduction Historical Outline Poly(malic acid) from Natural Sources Poly(malic acid) by Chemical Synthesis Chemical Structures Occurrence Poly(malic acid) from Natural Sources Poly(β‐malic acid) by Chemical Synthesis Functions Analysis Physicochemical Properties Chemistry Physiology Biochemistry Degradation and Biodegradation Production Applications Outlook and Perspectives Patents
Article
Photon correlation spectroscopy (PCS) has been applied to various systems of particles suspended in a fluid. Computer simulations were used to analyse CONTIN, a program for the numerical inversion of the experimental data. Measurements were performed on a spectrometer equipped with a digital correlator. Various suspensions of polysterene latex spheres differing both in size and size distribution were prepared in double distilled water. Further some samples provided by an external institution without prior specification were analysed. Diffusivity measurements on submicron DES (di-2-ethylhexyl sebacate) particles in nitrogen were also performed. The results obtained demonstrate the strengths and the limitations of PCS as a method of determining particle size distributions in particle fluid systems.
Article
Poly (-malic acid) is a functional aliphatic polyester which can serve as parent compound to make systems aimed at temporary therapeutic applications. Degradation of the polyester backbone of poly(-malic acid) leads to the formation of oligomers which can be hardly separated by conventional Size Exclusion Chromatography. In order to monitor the formation of oligomeric degradation by-products and to quantify their relative amount, the potential of two separation techniques has been investigated, namely Aqueous Size Exclusion Chromatography using a Bio-Gel P2 gel column and High Performance Capillary Electrophoresis. Aqueous SEC allowed separation of oligomers smaller than hexamer whereas HPCE successfully separated oligomers smaller than pentadecamers. These techniques were used to control the efficiency of oligomer fractionation by the dissolution/precipitation method.applied to the sodium salt of poly(-malic acid).
Article
Various fungi and myxomycetes have been tested for their synthesis and hydrolytic degradation of β-poly(L-malate) (PMLA), a highly anionic polyester of L-malate. The PMLA content of the culture medium, and also of some cell extracts, has been assayed by L-malate dehydrogenase (EC 1.1.1. 83)-catalysed reduction of NAD+ after the alkaline hydrolysis of the polymer. All eight myxomycete isolates were producers. Eight producers were Mitosporic fungi but only one Ascomyete was productive. Aureobasidium pullulans contained PMLA in yeast-like cells but not in hyphae. The polymer was bound to membrane fractions from where it was slowly released during preparation for analysis. Only 12 of the 237 fungi and myxomycetes showed PMLA-hydrolase activity; and two of these were producers of PMLA. Fungi with hydrolase activity were mostly different from those releasing L-malate. The Physarum polycephalum-type unbranched polymer of high molecular mass was not recognized in fungi. Fungi, typically A. pullulans, produced polymer of low molecular mass which seemed to be covalently bound to polysaccharide.