ArticlePDF Available

Six Bioabsorbable Polymers: In vitro acute toxicity of accumulated degradation products

Authors:

Abstract and Figures

Bioabsorbable polymer implants may provide a viable alternative to metal implants for internal fracture fixation. One of the potential difficulties with absorbable implants is the possible toxicity of the polymeric degradation products especially if they accumulate and become concentrated. Accordingly, material evaluation must involve dose-response toxicity data as well as mechanical properties and degradation rates. In this study the toxicity and rates of degradation for six polymers were determined, along with the toxicity of their degradation product components. The polymers studied were poly(glycolic acid) (PGA), two samples of poly(L-lactic acid) (PLA) having different molecular weights, poly(ortho ester) (POE), poly(epsilon-caprolactone) (PCL), and poly(hydroxy butyrate valerate) (5% valerate) (PHBV). Polymeric specimens were incubated at 37 degrees C in 0.05 M Tris buffer (pH 7.4 at 37 degrees C) and sterile deionized water. The solutions were not changed during the incubation intervals, providing a worst-case model of the effects of accumulation of degradation products. The pH and acute toxicity of the incubation solutions and the mass loss and logarithmic viscosity number of the polymer samples were measured at 10 days, 4, 8, 12, and 16 weeks. Toxicity was measured using a bioluminescent bacteria, acute toxicity assay system. The acute toxicity of pure PGA, PLA, POE, and PCL degradation product components was also determined. Degradation products for PHBV were not tested. PGA incubation solutions were toxic at 10 days and at all following intervals. The lower molecular weight PLA incubation solutions were not toxic in buffer but were toxic by 4 weeks in water.(ABSTRACT TRUNCATED AT 250 WORDS)
Content may be subject to copyright.
Six Bioabsorbable Polymers:
In
Vitro
Acute Toxicity
of
Accumulated
Degradation Products
M.
S.
Taylor,*
A.
U.
Daniels,*
K.
P.
Andriano,*
and
J.
Hellert
*Division
of
Orthopedic Surgery, University
of
Utah Health Sciences Center, Salt Lake City, Utah and
+SRl
International, Menlo Park, California
Bioabsorbable polymer implants may provide
a
viable alternative to metal implants for internal
fracture fixation. One of the potential difficulties with absorbable implants is the possible tox-
icity of the polymeric degradation products especially if they accumulate and become concen-
trated. Accordingly, material evaluation must involve dose-response toxicity data as well as
mechanical properties and degradation rates.
In
this study the toxicity and rates of degradation
for six polymers were determined, along with the toxicity of their degradation product compo-
nents. The polymers studied were poly(glyco1ic acid) (PGA), two samples of poly(L-lactic acid)
(PLA) having different molecular weights, poly(ortho ester) (POE), poly(t-caprolactone)
(PCL), and poly(hydroxy butyrate valerate)
(5%
valerate) (PHBV). Polymeric specimens were
incubated at 37°C in 0.05
M
Tris buffer (pH 7.4 at 37°C) and sterile deionized water. The
solutions were not changed during the incubation intervals, providing a worst-case model of the
effects of accumulation
of
degradation products. The pH and acute toxicity of the incubation
solutions and the mass
loss
and logarithmic viscosity number of the polymer samples were
measured at
10
days, 4,
8,
12, and 16 weeks. Toxicity
was
measured using a bioluminescent
bacteria, acute toxicity assay system. The acute toxicity of pure PGA, PLA, POE, and PCL
degradation product components was also determined. Degradation products for PHBV were
not tested. PGA incubation solutions were toxic at
10
days and at all following intervals. The
lower molecular weight PLA incubation solutions were not toxic in buffer but were toxic by 4
weeks in water. The other materials did not produce toxic responses during the 16-week
exposures. The degradation products components in order from most toxic to least toxic are:
lactic acid (PLA), c-caproic acid (PCL), glycolic acid (PGA), cyclohexane dimethanol (POE),
propionic acid (POE), 1,6 hexane diol (POE), pentaerythritol dipropionate (POE), and penta-
erythritol (POE).
0
1994
John
Wiley
&
Sons,
Inc.
INTRODUCTION
Metal implants are traditionally used for internal fracture
fixation. Problems with this mode of treatment include in-
complete healing due to stress shielding and the need for
surgical removal of the implant.'
A
potential solution is
implants made from bioabsorbable materials that are less
stiff and should not require surgical removal. An addi-
tional advantage
is
the gradual transfer
of
load from the
implant to bone as the implant loses stiffness, allowing
gradual bone remodeling that decreases the likelihood of
re fracture.
However, there are a number of concerns related to the
use of absorbable implants. One of the major concerns is
the potential for adverse reactions due to accumulated
degradation products. Poly(a-hydroxy acids) are the most
Requests
for
reprints should
be
sent to M.
S.
Taylor, Division
of
Orthopedic
Surgery,
University
of
Utah Health Sciences Center,
50
N.
Medical
Dr.,
Salt Lake
City,
UT
841
32.
Journal
of
Applied Biomaterials,
Vol.
5,
15
1
-
I57
(
1994)
0
1994
John Wiley
&
Sons,
Inc.
CCC
1045-486
1/94/020
I5
1-07
commonly used absorbable polymers for medical applica-
tions, and, much of the literature focuses
on
these materi-
als. Animal studies by Spenelhauer et al.' suggest a high
rate of noninfectious inflammatory response associated
with intermediate molecular weight degradation products
of poly(a-hydroxy acids). A high rate of sterile inflamma-
tory response occurred when poly(L-lactic acid) (PLA)
weight average molecular weight fell below
20,000
D. Use
of a higher initial molecular weight
PLA
delayed but did
not eliminate the response. Additional work by Schaken-
raad and Dijkstra3 demonstrated a sharp increase in solu-
bility for
PLA
and poly(glyco1ic acid) (PGA) as number
average molecular weight drops to about
5,000
to
10,000
Da. The molecular weight determinations were made us-
ing gel permeation chromatography (GPC) analysis, and
solubility by
in vivo
and
in vitro
weight loss. The apparent
difference in these results could be due to the polydisper-
sity of the material. Pistner et aL4 observed a difference in
degradation rate in a comparison of poly(L-lactides) of
three molecular weights. Two of the samples were amor-
phous and the third semicrystalline. The amorphous poly-
mers had a slower initial drop in inherent viscosity than
the semicrystalline polymer, equalizing at about week
20.
Additionally, the amorphous samples had a delayed
151
152
TAYLOR ET
AL.
weight loss. These results indicate the relative importance
of initial molecular weight and morphology. Another
study showed that PLAs exhibit accelerated degradation at
the center of bulk implants. The degraded polymer would
have limited diffusion out initially, possibly causing a large
delayed release of degradation products when the implant
exterior is suddenly breached. Also, both amorphous and
crystalline PLA develop increased crystallinity as degrada-
tion progre~ses.~ This indicates the potential for a change
in degradation rate due to morphology even as the implant
degrades. Other canine experiments showed bone resorp-
tion against PLA surfaces once weight loss from the im-
plant commences.6 When considered together, these re-
sults indicate the need for further evaluation prior to use
of poly(a-hydroxy acids) as bulk implants.
These concerns are amplified by the clinical results ob-
served in Europe with rods of PGA and PGA/PLA copol-
ymers or plates and screws of PLA for fracture fi~ation.~ Jn
this review by Bostman frequency of sterile inflammatory
response ranged from 5.7 to 47% of clinical cases. The
highest response rate was observed for a high molecular
weight (HMW) PLA and occurred 2 to
3
years postopera-
tively. The inflammation may be due in part to a high lo-
calized concentration of the degradation products. Inter-
estingly, the highest response rate for PLA occurred at the
ankle, the site ofthe lowest response rate for PGA. Clearly,
there are contributing factors that are not well understood
at this time, yet PGA and PLA are routinely used in
smaller volumes for suture materials and elicit much
smaller inflammatory responses in such cases. This sug-
gests a dose-response relationship for toxicity.
Little information is available comparing the toxicity
of degradation products as they accumulate or as a func-
tion of their concentration. One of the difficulties in com-
paring the toxicity of materials is the need to perform an
extensive testing battery to completely evaluate any mate-
rial.* The most common
in
vitro cytotoxicity tests rely on
cultures of animal or human cells with exposure intervals
ranging from a few minutes (Trypan Blue exclusion assay)
to 72 h. Discrepancies have been reported for different
time exposures, cell lines, and different methods
of
evalu-
ation. The difficulty in assessing toxicity is one of the indi-
cations for a battery of tests, as one may identify a poten-
tial problem with a material that might not be apparent
from the results of another test. This study relies upon a
single test, the bacterial bioluminescence test (BBT), for
comparison of toxicity. Accordingly, the present study
provides
a
basic comparison
of
the selected materials, but
does not provide definitive answers regarding the relative
toxicity.
The present study was undertaken to compare the deg-
radation and acute toxicity of accumulated degradation
products of six absorbable polymers under uniform
exposure conditions. A second phase of the study mea-
sured the toxicity of known quantities of pure degradation
products for five absorbable polymers.
MATERIALS
AND
METHODS
Bioabsorbable Polymers and
Controls
The polymers studied were PGA, two samples of PLA
having different molecular weights, poly(ortho ester)
(POE), poly( t-caprolactone) (PCL), and poly(hydroxy bu-
tyrate valerate)
(5%
valerate) (PHBV). All of these poly-
mers undergo hydrolysis. Table I lists the source and deg-
radation products for each material. Two sources of PLA
were selected to reflect its broad availability for medical
applications. The
POE
used in this study contained a mix-
ture of rigid
trans-cyclohexanedimethanol
and flexible
1,6
hexane diols present in a
60:40
mole ratio. The initial hy-
drolysis yields pentaerythritol mono- or dipropionate, the
two diols, and propionic acid. The pentaerythritol mono-
or dipropionate is further hydrolyzed to yield pentaeryth-
ritol and propionic acid.' Two potential advantages are
that the degradation is pH sensitive and can be controlled
and that the initial degradation products are pH neutral.
The PHBV used in this study contained 5 mol
%
hydroxy-
valerate to improve ductility. PHBV degrades primarily
by microbial attack, but was included in this study to eval-
uate its degradation in a sterile environment."
A low molecular weight polyethylene (PE-LMW) ob-
tained from Abiomed was used as negative control mate-
rial. The positive control material was the same PE with
40
wt%
trans-cinnamic acid added before hot molding the
specimens. Selection of the positive and negative control
materials was based on Burton et al.
I I
PHBV was provided in a powdered form.
All
other
polymer samples were milled to -40 mesh then freeze
dried for 48 h to minimize moisture and/or static charge
for ease of processing.
For the positive control, trans-cinnamic acid was finely
ground using a mortar and pestle and was slowly stirred
into milled PE. Mixing continued for
15
min at
less
than
100
rpm.
Milled polymers were hot compression molded. The
molding temperatures are given in Table 11. Amorphous
polymers (60:40 POE) were pressed at
50
"C above the
glass transition temperature, pressed at
5000
psi for
5
min,
and cooled to ambient temperature. Semicrystalline poly-
mers (PLA, PGA, PHBV, PCL, PE) were pressed
5"
below
their respective melting temperatures and pressed at 5000
psi for 5 min prior to cooling to ambient temperature. The
resulting samples were
1
-in. diameter disks weighing
1
g.
Thickness varied slightly with density of the polymers.
The disks were quartered using a jeweler's saw and an alu-
minum jig and stored in a vacuum desiccator jar with a
moisture absorber after molding until sterilization.
Each sample was weighed, numbered, and placed in a
vial. The samples in opened vials were sterilized by Nelson
Laboratories using ethylene oxide (ETO). The steriliza-
tion protocol was
1.5
h humidification at
38
"C
and 45%
relative humidity followed by
2.5
h exposure to
12%
ETO
carried in Freon@, and 48 h of degassing at
55
"C.
For PCL,
BIOABSORBABLE POLYMERS:
IN
VITRO
TOXICITY
153
TABLE
1.
Bioabsorbable Polvmers
Material Source, Trade
Name
Degradation Products
PGA DuPont, Medisorb lOOPGA Glycolic acid
PLA-LMW DuPont,
Medisorb
IOOL
Lactic acid
PLA-HM
W
Birmingham
Polymers, BP-0600 PLA Lactic acid
POE
SRI
International, no trade
name
Intermediate stage:
Pentaerythritol dipropionate, trans-cyclo hexane
dimethanol, 1,6 hexane diol
Pentaerythritol,
propionic
acid
Second
stage:
PCL Birmingham Polymers, BP-0800 PCL t-Hydroxy caproic acid
PHBV
ICI
America, no trade name Butyric acid, valeric
acid
the degassing temperature was lowered to
37
"C
and the
time extended to
72
h. This change was necessary due to
the low melting point of PCL.
Once sterilized, samples of each material were tested
for ETO residuals and sterility. All samples tested sterile
and had residual levels well below maximum acceptable
limits. The highest level of ETO was
<10.2
ppm, and the
most stringent FDA guideline for implants is
<25
ppm.
Incubation
Ten milliliters of sterile solution was added to each vial.
The two exposure media were sterile distilled water and
sterile Tris buffer. The Tris buffer solution was selected
because it has little or no effect on the response of the tox-
icity test method described subsequently under the condi-
tions used: pH
7.4
at
37
"C
and
0.05
A4.I'
The samples
were incubated at
37
"C. The removal dates were
10
days,
4,
8,
12,
and
16
weeks. At each removal,
pH
and toxicity
of the incubation solution and mass
loss
and logarithmic
viscosity number of the polymer specimens were mea-
sured. There were three samples per solution per time in-
terval for a total of
30
samples of each polymer for
exposure testing. In addition, there were three positive
control samples and three negative control samples for
each exposure medium as well as two samples for ETO
residuals testing and one sample for measurement of vis-
cosity after processing. The positive and negative controls
were tested at the first time interval only.
TABLE
II.
Hot
Pressing Temperature for Bioabsorbable Polymers
Material Press Temp ("C)
PGA
PLA-LM
W
PLA-HMW
POE
PCL
PHBV
215
165
170
110
50
160
Toxicity
Testing
Toxicity was measured using a commercially available
BBT assay system.
l3
The test procedure involved reconsti-
tuting lyophilized
Photobacterium phosphoreum
that emit
light indicative of their level of metabolic activity. The
bacteria were placed in
2%
saline in five test vials main-
tained at
15
"C.
Light output from the bacteria was mea-
sured in a photometer located adjacent to the refrigerated
cells. After serial dilution, test samples were added to four
of the vials, with the fifth serving as a negative control.
The light levels are then measured at
5
and
15
min after
addition of the test sample. The negative control sample
allowed corrections due to changes in luminescence with
time and volume. The toxicity data presented were given
by percentage light remaining,
(NOIN=,)
X
(
T5/TO),
where
the symbols Nand
T
refer to the light levels for the nega-
tive control and test sample vials, respectively; the sub-
script
0
refers to measurements taken immediately prior
to addition of the sample; and the subscript
5
refers to
measurements taken
5
min after addition of the sample.
A remaining light reading of
50%
or less is considered
toxic, and the concentration necessary to cause a 50%
reading is known as the ECso. For known dilution concen-
trations, the result is an equation relating concentration
and toxicity.
A
study by Burton et
a1.l'
compared this technique to
five standard acute toxicity tests for multiple samples of
eight materials. The tests were minimum Eagle's medium
elution tissue culture (MEM), mouse safety
(MS),
rabbit
intramuscular implantation (RII), rabbit intracutaneous
injection (RI), and mouse systemic injection (MSI). The
MEM was performed on confluent monolayers of mouse
fibroblast cells incubated for
24
to
48
h. The cells were
fixed, stained, and scored for cytotoxicity. As shown in
Table 111, of the
10
1
samples tested with MEM and BBT,
the two tests agreed for
95
samples
(87
nontoxic and
8
toxic) and disagreed
on
six samples
(4
BBT toxic,
2
MEM
toxic). This was the best correlation observed. The
MS
test
was performed on
93
samples of those, none were toxic
with MS and
20
were toxic with BBT. Twenty-six samples
were tested to compare BBT to
in vivo
toxicity screens
(RI,
154
TAYLOR
ET
AL.
TABLE
111.
Bacterial Bioluminescence Test (BBT) Compares Favorably with Other Acute Toxicity Testsll
RII,
MSI,
RI
MS
MEM
Comparison to BBT
(26
Samples Tested)
(93
Samples Tested)
(
10
I
Samples Tested)
Agree
Disagree
~
16
S
10
E
9P
10
S
(BBT toxic)
16
E
(BBT toxic)
17
P
(BBT toxic)
73
20
(BBT toxic)
87
8
(both
toxic)
4
(BBT toxic)
2
(MEM
toxic)
~
S,
E,
and Prefer to extracts
of
saline, ethanol/saline,
or
polyethylene glycol
400,
respectively. Numbers indicate test results that agree or disagreed with
BBT.
Except
as
noted,
test
results were nontoxic.
MSI, and/or
RII).
These showed the greatest dis-
agreement. Autian describes these tests as generally show-
ing a low frequency of response, but worth performing as
part of toxicity screen.l4 The numbers listed in Table
111
indicate the number of samples for which the response of
the tests agreed and disagreed. Tests with toxic results are
listed in parenthesis.
Another study'* determined the response of the BBT to
pH
of
water, buffer pH, and buffer concentration at phys-
iological pH. The pH of sterile water was shifted to basic
or acidic by addition of NaOH or HC1, respectively. Over
the pH range between 4 and
10,
no suppression
of
bacte-
rial light output was observed. The test of buffer pH and
concentration showed that sensitivity might be reduced by
use of a citrate phosphate borate buffer that mildly sup-
pressed the bacterial light output at all pH and concentra-
tions, but that Tris buffer was
a
well-suited medium for an
incubation solution for the studies described here.
Polymer Mass
Loss
The polymer specimen mass
loss
was determined by
weighing the samples after drying in a vacuum oven at
ambient temperature (approximately
23
"C) for at least 48
h. The samples were considered dry when the weight taken
on successive days was unchanged.
Polymer Viscosity Measurement
Reduced viscosity
of
the polymer specimens in solvent
so-
lutions was measured using a Cannon-Ubbelhode vis-
cometer according to ASTM D2857.15 Logarithmic vis-
cosity number was measured for the polymers as received,
as pressed, and at each time exposure. Measurement of the
polymers as received and as pressed identified any change
in molecular weight due to processing. Hexafluoro-2-pro-
panol was the solvent used for PGA, and chloroform was
the solvent used for all other polymers. For POE, triethyl
amine was added to the solvent to stabilize the solution.
The concentration used for all samples was approximately
5
mg/mL. The formula used for calculation of viscosity
was
77
=
[ln(t,/t,)]/c,
where
ta
is the average of three flow
times for the polymer in solution,
t,
is the average of three
flow times for the pure solvent, and
c
is the concentration
in grams/milliliter. To hasten dissolution,
PHBV
samples
were milled to --SO mesh after incubation and prior to vis-
cosity testing.
Toxicity
of
Degradation Product Components
Pentaerythritol dipropionate was synthesized at
SRI
by es-
terifying pentaerythritol with propionic acid. The lactic
acid, glycolic acid, caprolactone, pentaerythritol, propi-
onic acid, cyclohexane dimethanol, and 1,6 hexane diol
were obtained from Aldrich Chemical. Using the BBT,
EC50 values for each of the degradation products were de-
termined. As described previously, POE degrades
in
stages. For acute toxicity evaluation, the intermediate
stage was considered to consist of the two diols and penta-
erythritol dipropionate, and the final stage to consist of the
two diols, pentaerythritol and propionic acid.
ECS0s
were
determined for both combinations. Under actual
in
vivo
conditions,
POE
degradation products are more likely to
consist of a combination of all possible degradation prod-
uct components.
RESULTS
Table
IV
presents the toxicity and pH of the incubation
solutions and weight loss and viscosity for the incubated
polymer specimens at the 16-week time interval and the
interval at which materials first elicited a toxic response.
Only PGA and the PLA-HMW in water incubation solu-
tions elicited toxic responses during the exposure period.
Each material
is
discussed separately below.
Controls
The incubation solutions for the negative and positive
controls were tested for pH and toxicity at
10
days. Weight
loss and viscosity were not determined. The negative con-
trols were all nontoxic, with an average light output for
the strongest concentration tested
of
99%.
The average pH
levels were 7.7 for Tns exposures and 6.8 for the water
exposures. The positive controls all caused a complete
155
BIOABSORBABLE POLYMERS:
IN
VITRO
TOXICITY
TABLE
IV.
Data
for
Final Exposures and First
Toxic
Exposures
Polymer Final Incubation Solution
16-Week Data Mass Loss Viscosity Remaining
Material and Exp. Media
(%I
(% of
original) PH Light
(%)
0.0
t
0.00
PGA
(water)
64.85
t
0.63 18.7
t
2.4 2.49
f
0.06
0.0
k
0.00
PLA-LMW (water)
4.69
-t
1.36 47.0
k
1.7 3.03
f
0.04
6.80
f
0.02
82.1
t
2.88
PLA-LMW
(Tris)
3.62
k
0.09 48.7
k
2.6
PLA-HW (water)
-0.05
k
0.01
19.5
-t
3.4 1.30
k
0.06 96.2
t
6.32
POE (water)
1.29
t
0.12 95.4
t
1.6 6.56
f
0.05 57.2
t
7.95
POE
(Tris)
1.58
f
1.61 89.3
f
8.3 1.75
k
0.05 68.4
f
6.69
PCL (water)
0.49
t
0.04 80.9
k
7.4 6.41
k
0.21 98.3
t
4.62
PCL
(Tris)
0.69
f
0.09 86.8
+.
2.5 7.68
?
0.00
100.5
f
5.58
PHBV
(water)
0.03
f
0.03 84.5
f
1.1
6.40
f
0.14 94.8
f
6.63
PHBV
(Tris)
0.02
k
0.02
80.4
t
0.8 7.75
t
0.03 101.5 k2.51
First Toxic Data
PGA
(Tris)
63.38
t
0.66 25.9
?
2.8 2.85
f
0.04
0.0
k
0.00
PLA-HW
(Tris)
-0.10
t
0.05 77.0
k
4.6 7.72
k
0.02 88.4
f
6.24
PGA (water),
10
day
PGA
(Tris),
10
day
PLA-LMW (water),
4
week
3.25
?
0.3
3.46
t
0.4
0.47
?
0.57
35.4
f
6.7 2.81
t
0.05
0.00
t
0.00
29.3
f
3.3 5.05
t
1.25
0.00
f
0.00
87.9
t
3.6 3.57
?
0.13 7.16t 10.13
Average
values
k
SD.
suppression of the light output, with average pH of
5.5
for
the Tris exposure media and 3.8 for the water.
PGA
For PGA, the pH dropped in both exposure media quite
rapidly to less than
5.
For the 10-day buffered exposure
media, the average pH was
5.05,
which would not cause
the toxic response observed. The incubation solution for
PGA was toxic at all time intervals for both water and
buffer exposures. All tested dilutions of the incubation
so-
lution were toxic. The concentration of the most dilute of
the test solutions was 5.6% of the concentration of the ini-
tial solution, indicating that the exposure solution was
well beyond the threshold of toxicity. The measured
ECsO
for glycolic acid was
100
X
M.
The average mass loss
for the first toxic exposures was
9
mg
(3.5%).
The corre-
sponding calculated concentration would be between
1200
X
Mif degraded to glycolic acid and 180
X
1
0-6
M
if degraded to a molecular weight of
5000.
Either con-
centration would produce
a
toxic response. However, the
most dilute sample (a concentration of 5.6% ofthe original
sample) had no light output suggesting that the latter con-
centration was
a
low estimate. The viscosity decreased
with time indicating a continued drop in average molecu-
lar weight of the polymer specimen.
PLA-LMW
Results for the PLA-LMW were substantially different for
the water exposures than for the buffer exposures.
Water Exposures.
The pH dropped rapidly, to an aver-
age of 3.6 by the 4-week exposure period, and remained
relatively constant afterward. The incubation solution was
toxic at 4 weeks and at all subsequent time intervals. The
average mass loss associated with the first toxic results was
1.2 mg
(0.5%).
The calculated concentration was between
1300
X
M
if degraded to a molecular weight of
5000.
The measured
ECsO
for lactic acid is
100
X
M.
The polymer viscosity
decreased nearly linearly with time to a value of about
50%
of the original value, indicating a large drop in average
molecular weight.
Mif degraded to lactic acid and
24
X
Tris
Exposures.
The pH dropped only slightly, to an
average of 6.8 by the 16-week exposures. The incubation
solution was not toxic at any time interval. The average
mass loss by the sixteenth week was 8.8 mg (3.6%). The
calculated concentration was
9700
X
M
if degraded
to lactic acid or
170
X
1
OP6
M
if degraded to a molecular
weight
of
5000.
The
ECso
for lactic acid was
100
X
1
0-6
M
indicating that the actual concentration in the incubation
solution lies closer to the latter value than to the former.
The polymer viscosity drop was similar in buffered
exposures to that seen for water exposures.
PLA-HMW
The pH for the PLA-HMW remained relatively constant
throughout the test intervals for both exposure media. The
bacterial light output was not suppressed for any of the
tests. PLA-HMW showed a slight weight loss at the initial
156
TAYLOR
ET
AL.
time interval and no weight loss in the subsequent in-
tervals. The actual data for the final exposure indicated a
slight gain in mass
(<O.
1%).
The polymer viscosity drop
was about 25% in both media, indicating that degradation
occurred, although solvation did not.
POE
The pH of the POE incubation solutions remained rela-
tively constant throughout the test in both exposure me-
dia. None of the incubation solutions were toxic, although
there was some light suppression at the later time in-
tervals. Average weight losses for the 16-week samples of
POE were 3.2 mg (1.3%) in water and
4.0
mg (1.6%) in
buffer. The resulting concentration was between 1300
x
1
0-6
Mand
2600
X
M.
The former value represents
degradation to the two diols and pentaerythritol dipropio-
nate, and the latter value represents degradation to the two
diols, pentaerythritol and propionic acid.
PCL
The pH of the PCL incubation solutions remained rela-
tively constant throughout the test in both exposure me-
dia. The average mass losses for the 16-week PCL samples
were 1.2 mg
(0.5%)
in water and 1.7 mg
(0.7%)
in buffer.
The corresponding concentrations of degradation prod-
ucts were
530
X
M
in
buffer. The measured ECSO for caprolactone was
450
X
M.
The calculated concentration was close to the
measured ECSo, yet the bacterial light output was not sup-
pressed. This suggests that the actual concentration was
lower than calculated. One possible explanation is that in-
termediate molecular weight polymer chains become
sol-
uble in a manner similar to that seen with PLA and PGA.
The polymer viscosity for PCL decreased by
25%
during
the exposure period, indicating a drop in average molecu-
lar weight.
M
in water and
750
X
PHBV
The pH of the PHBV incubation solutions remained rela-
tively constant throughout the test in both exposure me-
dia. None of the incubation solutions were toxic. PHBV
showed virtually
no
mass
loss
(<0.05%),
but polymer vis-
cosity did drop, indicating degradation occurred in a ster-
ile environment. One can conclude that microbial attack
is not the only method of degradation for PHBV.
Degradation Products
The EC5o values of the degradation products are shown in
Table
V.
Two values are shown for POE. The two values
represent the difference in toxicity between the first and
final stage products. PHBV is not included because its deg-
radation products were not tested. The tendency for poly-
mers to become soluble at intermediate molecular weights
TABLE
V.
Relative Toxicities
of
Degradation Products
Toxic
Concentration
PGA 150 87
PLA
100
72
POE (intermediate stage)
1000
336
POE (final) 270
91
PCL 120 137
makes it difficult to predict the mass loss that correlates
with toxicity. The value presented assumes degradation to
single repeat units.
DISCUSSION
This study provides insight into the relative acute toxici-
ties of bioabsorbable polymer degradation products. In or-
der from most toxic to least toxic they are: lactic acid
(PLA), c-caproic acid (PCL), glycolic acid (PGA), cyclo-
hexane dimethanol (POE), propionic acid (POE), 1,6 hex-
ane diol (POE), pentaerythritol dipropionate (POE), and
pentaerythritol
(POE).
The difference in concentration
that may be tolerated differs by a factor of 10. The toxicity
is a factor that should be considered in material selection,
but must be combined with degradation rate and local tis-
sue clearance to predict the concentration present in the
tissue, and the resulting response.
in vivo
degradation rates
are affected by not only the polymer crystallinity and mo-
lecular weight but also by the implant size, shape, and im-
plantation site. The order of degradation could be consid-
ered by either mass loss
or
viscosity decrease. Because vis-
cosity decrease represents a loss of mechanical properties,
that is, modulus, it is presented here. The order of viscosity
decrease, from most affected to least affected, is: PGA,
PLA-LMW, PLA-HMW, PHBV, PCL, and POE. Note
that the data presented in Table
4
indicates that PLA-
LMW and PLA-HMW exhibited markedly different be-
haviors, thus demonstrating the role of molecular weight
in bioabsorbables. Similarly, two otherwise identical ma-
terials with different morphologies may exhibit different
degradation rates, but this aspect was not considered in
this study.
The materials tested as degradation product compo-
nents were well controlled uniform solutions. The ex-
pected degradation product
in vivo
probably contain not
only a mixture of molecular weights, but also might exist
in different states of ionization. Both factors could alter
the biologic response.
The results observed are also affected by the particular
processing parameters. For the hot compression molding,
difficulty was encountered with PGA. In the samples used,
granularity was evident. Use
of
a higher processing tem-
perature was attempted without success to reduce the
BIOABSORBABLE
POLYMERS:
IN
VITRO
TOXICITY
157
granularity.
No
doubt
the
granularity
affected
the
rate
at
which
fluid
penetrated the
sample,
thus
affecting
the
deg-
radation
rate.
We
would
like to thank Osteotech
for
support
for
this project and SRI Interna-
tional
for
donation
of
POE.
REFERENCES
1.
Daniels, A.
U.;
Chang, M.
K.
0.;
Andriano,
K.
P.; Heller, J.
Mechanical properties of biodegradable polymers and com-
posites proposed
for
internal fixation of bone. J. Appl. Bio-
mater. 157-78; 1990.
2. Speneihauer, G.; Vert, M.; Benoit,
J.
P.; Boddaert, A.
In
vi-
tro
and
in
vivo
degradation of POIY(D,L lactide/glycolide)
type microspheres made by solvent evaporation method.
Biomaterials
10:
5
57-563; 19 89.
3. Schakenraad, J. M.; Dijkstra, P. J. Biocompatibility of
pOly(D,L-laCtiC acidiglycolide) copolymers. Clin. Mater. 7:
4. Pistner, H.; Bendix, D.; Muhling,
J.;
Reuther, J. Poly(L-lac-
tide): a long-term degradation study
in vivo.
Part
111.
Analyt-
ical characterization. Biomaterials 14:29 1-298; 1993.
5.
Li,
S.;
Garreay, H.; Vert, M. Structure-property relation-
ships in the case of the degradation of massive poly(cY-hy-
droxy acids) in aqueous media. J. Mater. Sci. Mater. Med.
1
:
6. Suganuma, J.; Alexander,
H.
Biological response of intra-
253-269; 1991.
198-206; 1990.
medullary bone to poly-L-lactic acid. J. Biomed. Mater. Res.
(in press).
7. Bostman,
0.
M. Current concepts review: absorbable im-
plants for the fixation of fractures.
J.
Bone Joint Surg.
8. Balls, M.; Clothier, R.
H.
Cytotoxicity assays for intrinsic
toxicity and imtancy. In: Watson, R. R., Ed.
In
vitro
meth-
ods
of
toxicology. Boca Raton,
FL
CRC Press; 1992:37-53.
9. Heller, J. Controlled drug release from poly(ortho esters)-
A
surface eroding polymer. J. Controlled Release
2:
167-
177; 1985.
10.
PHBV biodegradable polyesters. Wilmington, DE: ICI
Americas; 1988.
1
1.
Burton,
S.
A.;
Peterson,
R.
V.;
Dickman,
S.
N.;
Nelson, J. R.
Comparison of
in vitro
bacterial bioluminescence and tissue
culture bioassays and
in
vivo
tests for evaluating acute toxic-
ity. J. Biomed. Mater. Res. 20:827-838; 1986.
12. Daniels, A.
U.;
Andriano,
K.
P.; Felix, B.; Hamber, E. The
effect
of
pH and buffer solutions on the bacterial biolumi-
nescence test for acute toxicity. Trans. SFB
1
10; 199
1.
1
3. Microtox manual. Carlsbad, CA: Microbics Corporation;
1991.
14. Autian, J. Toxicological evaluation of biomaterials: primary
acute toxicity screening program. Artif. Organs 153-60,
1977.
15.
Annual book of standards. Philadelphia, PA: American
So-
ciety for Testing and Materials; 1987.
73A( 1): 148-1 53; 199
I.
Received March
15,
1993
Accepted February
10,
1994
... In rabbit and mouse xenograft models, PGA and PLA copolymers, in conjunction with dental pulp progenitor cells, facilitated the formation of pulp-like tissue. [86][87][88] PLA has successfully formed dental pulp and periodontal tissues. 89 This material can potentially enhance the proliferation of dental pulp stem cells compared to collagen or calcium phosphate scaffolds. ...
... Their main benefit for medical applications resides in the fact that their degradation products are natural metabolites, normally excreted in urine [115]. However, concerns about PLA/PGA biocompatibility have been raised due the accumulation of the degradation products [116]. PLGA is a copolymer composed of two different monomers, the cyclic dimers of glycolic acid and lactic acid. ...
Article
Full-text available
Background and Objectives: Regenerative dentistry aims to regenerate the pulp–dentin complex and restore those of its functions that have become compromised by pulp injury and/or inflammation. Scaffold-based techniques are a regeneration strategy that replicate a biological environment by utilizing a suitable scaffold, which is considered crucial for the successful regeneration of dental pulp. The aim of the present review is to address the main characteristics of the different scaffolds, as well as their application in dentin–pulp complex regeneration. Materials and Methods: A narrative review was conducted by two independent reviewers to answer the research question: What type of scaffolds can be used in dentin–pulp complex regeneration? An electronic search of PubMed, EMBASE and Cochrane library databases was undertaken. Keywords including “pulp-dentin regeneration scaffold” and “pulp-dentin complex regeneration” were used. To locate additional reports, reference mining of the identified papers was undertaken. Results: A wide variety of biomaterials is already available for tissue engineering and can be broadly categorized into two groups: (i) natural, and (ii) synthetic, scaffolds. Natural scaffolds often contain bioactive molecules, growth factors, and signaling cues that can positively influence cell behavior. These signaling molecules can promote specific cellular responses, such as cell proliferation and differentiation, crucial for effective tissue regeneration. Synthetic scaffolds offer flexibility in design and can be tailored to meet specific requirements, such as size, shape, and mechanical properties. Moreover, they can be functionalized with bioactive molecules, growth factors, or signaling cues to enhance their biological properties and the manufacturing process can be standardized, ensuring consistent quality for widespread clinical use. Conclusions: There is still a lack of evidence to determine the optimal scaffold composition that meets the specific requirements and complexities needed for effectively promoting dental pulp tissue engineering and achieving successful clinical outcomes.
Article
Three‐dimensional (3D) bio‐printing of elastic cartilage tissues that are mechanically and structurally comparable to their native counterparts, while exhibiting favorable cellular behavior, is an unmet challenge. A practical solution for this problem is the multi‐material bio‐printing of thermoplastic polymers and cell‐laden hydrogels using multiple nozzles. However, the processing of thermoplastic polymers requires high temperatures, which can damage hydrogel‐encapsulated cells. In this study, we developed waterborne polyurethane (WPU) – polycaprolactone (PCL) composites to allow multi‐material co‐printing with cell‐laden gelatin methacryloyl (GelMA) hydrogels. These composites could be extruded at low temperatures (50−60°C) and high speeds, thereby reducing heat/shear damage to the printed hydrogel‐capsulated cells. Furthermore, their hydrophilic nature improved the cell behavior in vitro. More importantly, the bio‐printed structures exhibited good stiffness and viscoelasticity compared to native elastic cartilage. In summary, our study demonstrated low‐temperature multi‐material bio‐printing of WPU‐PCL‐based constructs with good mechanical properties, degradation time‐frames, and cell viability, showcasing their potential in elastic cartilage bio‐fabrication and regeneration to serve broad biomedical applications in the future. This article is protected by copyright. All rights reserved
Article
Peripheral nerve injury (PNI) is a common condition in orthopedic clinical practice, leading to sensory and motor dysfunction in the affected limbs. The quality of life for patients is drastically affected by this, causing a significant burden on their family and society. Currently, patients with peripheral nerve transection injuries or combined nerve defects requiring autologous nerve transplantation undergo surgical treatment. However, postoperative functional recovery of the affected limbs is often incomplete, and the source of autologous nerve grafts is limited. Therefore, promoting damaged nerve repair and restoration of limb function remains a challenging issue. In recent years, with the continuous advancement of materials science, tissue engineering, and regenerative medicine, the development of biomaterials has provided a new approach for repairing PNI. Biomaterials used for PNI repair include polymer materials, natural materials, and composite materials, with many experimental research results indicating their ability to promote the repair of PNI. This article reviews the application of biomaterials in the repair of PNI, discussing their unique properties and advantages and disadvantages for peripheral nerve restoration. The aim is to provide theoretical support for the further development of novel biomaterials for PNI repair.
Article
Full-text available
As a clinically approved treatment strategy, chemotherapy-mediated tumor suppression has been compromised, and in spite of introducing various kinds of anticancer drugs, cancer eradication with chemotherapy is still impossible. Chemotherapy drugs have been beneficial in improving the prognosis of cancer patients, but after resistance emerged, their potential disappeared. Oxaliplatin (OXA) efficacy in tumor suppression has been compromised by resistance. Due to the dysregulation of pathways and mechanisms in OXA resistance, it is suggested to develop novel strategies for overcoming drug resistance. The targeted delivery of OXA by nanostructures is described here. The targeted delivery of OXA in cancer can be mediated by polymeric, metal, lipid and carbon nanostructures. The advantageous of these nanocarriers is that they enhance the accumulation of OXA in tumor and promote its cytotoxicity. Moreover, (nano)platforms mediate the co-delivery of OXA with drugs and genes in synergistic cancer therapy, overcoming OXA resistance and improving insights in cancer patient treatment in the future. Moreover, smart nanostructures, including pH-, redox-, light-, and thermo-sensitive nanostructures, have been designed for OXA delivery and cancer therapy. The application of nanoparticle-mediated phototherapy can increase OXA's potential in cancer suppression. All of these subjects and their clinical implications are discussed in the current review.
Article
Full-text available
Under diabetic conditions, blood glucose fluctuations and exacerbated immunopathological inflammatory environments pose significant challenges to periosteal regenerative repair strategies. Responsive immune regulation in damaged tissues is critical for the immune microenvironment, osteogenesis, and angiogenesis stabilization. Considering the high‐glucose microenvironment of such acute injury sites, a functional glucose‐responsive immunomodulation‐assisted periosteal regeneration composite material—PLA(Polylactic Acid)/COLI(Collagen I)/Lipo(Liposome)‐APY29 (PCLA)—is constructed. Aside from stimulating osteogenic differentiation, owing to the presence of surface self‐assembled type I collagen in the scaffolds, PCLA can directly respond to focal area high‐glucose microenvironments. The PCLA scaffolds trigger the release of APY29‐loaded liposomes, shifting the macrophages toward the M2 phenotype, inhibiting the release of inflammatory cytokines, improving the bone immune microenvironment, and promoting osteogenic differentiation and angiogenesis. Bioinformatics analyses show that PCLA enhances bone repair by inhibiting the inflammatory signal pathway regulating the polarization direction and promoting osteogenic and angiogenic gene expression. In the calvarial periosteal defect model of diabetic rats, PCLA scaffolds induce M2 macrophage polarization and improve the inflammatory microenvironment, significantly accelerating periosteal repair. Overall, the PCLA scaffold material regulates immunity in fluctuating high‐glucose inflammatory microenvironments, achieves relatively stable and favorable osteogenic microenvironments, and facilitates the effective design of functionalized biomaterials for bone regeneration therapy in patients with diabetes.
Preprint
Full-text available
Composite biomaterials comprising polylactide (PLA) and hydroxyapatite (HA) are applied in bone, cartilage and dental regenerative medicine, where HA confers osteoconductive properties. However, after surgical implantation, adverse immune responses to these composites can occur, which have been attributed to size and morphology of HA particles. Approaches to effectively modulate these adverse immune responses have not been described. PLA degradation products have been shown to alter immune cell metabolism, which drives the inflammatory response. Therefore, we aimed to modulate the inflammatory response to composite biomaterials by regulating glycolytic flux with small molecule inhibitors incorporated into composites comprised of amorphous PLA (aPLA) and HA (aPLA+HA). Inhibition at specific steps in glycolysis reduced proinflammatory (CD86+CD206-) and increased pro-regenerative (CD206+) immune cell populations around implanted aPLA+HA resulting in a pro-regenerative microenvironment. Notably, neutrophil and dendritic cell (DC) numbers along with proinflammatory monocyte and macrophage populations were decreased, and Arginase 1 expression among DCs was increased. Targeting immunometabolism to control the inflammatory response to biomaterial composites, and creating a pro-regenerative microenvironment, is a significant advance in tissue engineering where immunomodulation enhances osseointegration, and angiogenesis, which will lead to improved bone regeneration.
Article
Full-text available
The bacterial bioluminescence test for acute toxicity offers a rapid, uncomplicated, and inexpensive means for preliminary biocompatibility evaluation of materials. Sensitivity is equal to in vitro tissue culture and in vivo mouse safety, mouse systemic injection, rabbit intracutaneous injection, and rabbit intramuscular implantation bioassays. Frequently, in vitro biodegradation evaluation of biomaterials involves exposure to solutions made with a variety of buffers at various pH values. This paper describes investigations to determine the effects of three buffers on the performance of the luminescent bacterial assay.
Article
Bioabsorbable synthetic polymers have been studied for their possible application in absorbable internal fracture fixation devices. The current study examines the biological response of intramedullary bone to PLLA (poly-L-lactic acid). PLLA degrades at a rate sufficiently slow to be useful for fracture fixation and undergoes hydrolytic deesterification to form metabolites normally found in the body. Nevertheless, the lactic-acid-rich degradation products have the potential to significantly lower the local pH in a closed space surrounded by bone. It is hypothesized that this acidity may tend to cause abnormal bone resorption and/or demineralization.
Article
Poly(ortho esters) are a polymer system containing backbone linkages that are stable in base, hydrolyze at very slow rates at the physiological pH of 7.4, and become progressively more labile as the pH is lowered. A major rationale for developing this system was a need for a polymer capable of a wide variety of erosion rates and where erosion could be confined to the surface of a solid device. Surface erosion and variations in delivery rates of therapeutic agents physically incorporated into the polymer can be achieved by either stabilizing the interior of the polymer with a base so that erosion can take place only in the surface layers where the basic excipient is neutralized by the external medium or by using acidic excipients incorporated into the typically highly hydrophobic matrix. In this latter case, surface erosion takes place only in the surface layer where the excipient is exposed to water.
Article
This study focuses on examining the biological response of intramedullary bone to poly-L-lactic acid (PLLA), particularly during the PLLA degradation phase. To study the influence of spherical crystals (spherulites) of PLLA on intramedullary bone response, two different types of PLLA coupon, with and without spherulites but with the same molecular weight, were used. Chambers containing PLLA coupons were implanted into the right femur of eight dogs, four with and four without spherulites; chambers containing stainless steel (SS) coupons (as a control) were implanted in the left femurs of all eight. Two dogs, one with and one without spherulites, were sacrificed at 3, 6, 12, and 24 weeks postoperatively. Histomorphometric evaluation and histophathological assessment were used to compare the response to PLLA and SS. Scanning electron micrographs showed that there were minimal changes in the surface of PLLA coupons at 3 and 6 weeks. But at 12 and 24 weeks, there were many cracks and holes on the surfaces of the coupons, and some parts of the surface were scaling off. The cross-sectional area of PLLA coupons showed no change at 3 and 6 weeks, but started to decrease by 12 weeks. The amount of ingrown bone between PLLA coupons was significantly greater than that between SS coupons at 3 and 6 weeks, but had decreased dramatically by 12 weeks. Extensive bone resorption around PLLA coupons occurred by 12 weeks accompanied by infiltration of inflammatory cells. An abundance of histiocytes, giant cells, and leucocytes were seen, along with a few histiocytes that had phagocytized PLLA particles of less than 2 μm. By contrast, no inflammatory reaction was seen in SS samples at any period up to and including 24 weeks. PLLA demonstrated excellent biocompatibility with intramedullary bone for the first 6 weeks in this model. Once degradation commenced, however, biocompatibility decreased dramatically. Our study detected no difference between coupons with and without spherulites. It thus appears that the existence of relatively large PLLA particles did not influence the response of intramedullary bone to PLLA, but rather that it was the smaller particles (< 2 μm) released from the PLLA that induced foreign-body inflammatory reactions and bone resorption. It is also possible that a local decrease in pH occurred around PLLA coupons, which could have influenced vital kinetics. © 1993 John Wiley & Sons, Inc.
Article
A standard protocol is proposed which has been used to study comparatively the degradation mechanism of bioresorbable poly(-hydroxy acids) with respect to macromolecular structural characteristics and solid-state morphologies. As a first approach, the hydrolytic degradation of poly(dl-lactic acid) (PLA50) parallelepipedic specimens (15 mm10 mm2 mm), processed by compression moulding and machining, was investigated in two aqueous media: iso-osmolar saline and pH 7.4 phosphate buffer. Various techniques (namely weighing, size-exclusion chromatography (SEC), potentiometry, cryometry and enzymatic assay) have been applied to these specimens in order to monitor the degradation. Data show conclusively that the degradation of massive PLA50 specimens proceeds more rapidly in the centre than at the surface. This feature has been related to the formation of an outer layer of slowly degrading polymer, which is caused by surface phenomena and entraps degrading macromolecules. Only oligomers can diffuse and dissolve in the surrounding media. Accordingly, the number of carboxylic groups present in the inner part of the degrading specimens becomes larger than at the surface and accelerates ester bond cleavage. The resultant autocatalytic mechanism explains well the fact that partially degraded PLA50 exhibits bimodal SEC chromatograms although this polymer is amorphous.
Article
In order to assess the effects of morphology on the degradation characteristics of high-molecular weight poly(l-lactic acid) (PLA100), specimens of similar sizes were processed by compression moulding and made either amorphous by quenching (PLA100A) or semicrystalline by annealing (PLA100C). PLA100A specimens were allowed to age in iso-osmolar saline and pH 7.4 phosphate buffer at 37 C for periods up to 2 years, whereas PLA100C specimens were studied in the buffer only. Various techniques were used to monitor comparatively the effects of morphology on the mechanism of hydrolytic degradation for these two types of PLA100 specimens: weighing, enzymatic assay, potentiometry, viscoelasticimetry, size-exclusive chromatography, (SEC), X-ray scattering and differential scanning calorimetry. As in the case of non-crystallizable members of the poly(-hydroxy acid) family, degradation was found to proceed more rapidly in the centre than at the surface for both PLA100A and PLA100C specimens. However, the observed multimodal SEC chromatograms have been assigned primarily to differences of degradation rates in amorphous and crystalline domains, regardless of the initial morphology. Indeed, it was found that initially amorphous PLA100A crystallized as degradation proceeded. Furthermore, PLA100A specimens retained mechanical properties for longer than semicrystalline PLA100C specimens, probably because of the sensitivity of the latter to stress and solvent microcracking. When the integrity of the polymer mass was lost, the residual crystalline matter, initially present or formed during degradation, appeared to be very resistant and was still present in a powdered form after 2 years. It is concluded that the morphology is a critical factor for the degradation of PLA100 and that the degradation of bioresorbable devices derived from this polymer should depend very strongly on both the thermal history and the initial crystallinity. The effects of the morphology do not depend significantly on the nature of the ageing medium, provided that the ionic strength is the same.