3.1. Experiment
Lactic acid is a highly hygroscopic compound hence usually handled in concentrated aqueous solutions (10-40% by weight) [
43]. In our experiments, commercial (L or DL) lactic acid containing about 12% residual water was used as starting material. A lactic acid/toluene mixture was placed in a flask equipped with a Dean-Stark apparatus and a condenser and was refluxed at 140 °C under stirring for 2 hours. Most of the residual water was collected in the Dean-Stark burette and removed (i.e., 1.2 g water per 8.1 g of commercial lactic acid) along with toluene condensed through the azeotropic distillation. During this process, 5% w/w of lactic acid was converted to L-lactide as determined by
1H-NMR (
Figure S1).
The polymerization process was performed in a second step during which, the Dean-Stark apparatus was replaced by a glass tube packed with activated molecular sieves 3 Å and the catalyst tin octanoate catalyst was added. The molecular sieves packed tube was used to remove both residual water and water produced during polymerization as it was previously shown that recycling of the solvent through the molecular sieves may reduce the water content of the mixture to as low as 3 ppm or less [
21,
32]. Additionally, the temperature was slightly reduced (130-137 °C) during polymerization to limit further formation of lactides thus avoiding loss of stereochemistry. Lactides are commonly formed at temperatures of 140 °C or higher [
4,
44], while their polymerization occurs above 180 °C under reduced pressure [
4]. Even though formation of lactides is unlikely to occur under the comparatively milder conditions of the protocol reported herein, the prolonged reaction times might lead in partial reaction of any lactides formed during the first step. In such a case, the optical purity of the resulting PLA would not be guaranteed, since a mixture of D-, L-, and meso-Lactides could form as a result of the racemization of L-lactic acid. Targeting high molecular weight PLAs, the polycondensation was allowed to proceed for specified amounts of time after which, the product was dissolved in THF or chloroform, filtered and studied with
1H-NMR and GPC.
In the
1H-NMR spectrum of PLLA (
Figure 1 and
Figure S2), the double peak at 1.52 ppm is attributed to the methyl groups of the main polymeric chain (
Figure 1 and
Figure S2, protons c) and the quartet at 5.10 ppm to the adjacent methine protons (
Figure 1 and
Figure S2, protons a). In the
13C-NMR spectrum of PLLA three peaks appear at 169.6, 69.0 and 16.6 ppm which can be attributed to the ester carbonyl, methines and methyl groups of the main polymeric chain respectively (
Figure 1 and
Figure S3). To conclude whether the direct polycondensation of L-Lactic acid led to the formation of optical pure PLLA, the reaction was also performed using the racemic mixture of DL-lactic acid. As shown in the spectra acquired for PDLLA (
Figures S4 and S5) the presence of diastereomers causes an easily detectable change in the multiplicity of the characteristic peaks of the polymer. Hence, the NMR spectra of PLA provided a first indication that the direct polycondensation of L-lactic acid at the reported reaction conditions led to the formation of optically pure PLLA. Notably, no formation of terminal alkenes could be detected in either the
1H- or
13C-NMR spectra of the products (
Figure 1,
Figures S2 and S3), which could be an indication of thermal degradation of PLLA [
45].
The polycondensation reaction was optimized in terms of temperature, monomer and catalyst concentration and reaction time (
Table S2). Reaction times prolonged by as much as 5 days were found to be important for the synthesis of high molecular weight PLLA. In particular, the average molecular weight (M
n) of PLA was 41 kDa (
Table S2, Entry 1) after 92 hours while, it almost doubled (81 kDa) when the reaction mixture was further heated for 28 hours (
Table S2, Entry 2). Entry 2 in
Table S2 provides the optimal polycondensation conditions identified in this study based on the targeted increased PLA molecular weights. Alterations either in concentration or temperature led to the formation of lower molecular weight PLAs (
Table S2, Entries 3-5 and 8 respectively). The importance of dehydration was clearly demonstrated when the polycondensation of DL-Lactic acid was performed without stirring and/or solvent recycling yielding only low molecular-weight products under otherwise same reaction conditions (
Figure S6,
Table S2, Entries 9 and 10). Aiming to synthesize high molecular weight PLLA without prolonged reaction times, we accessed the cross-linker 1,4-phenylene diisocyanate. The cross-linker was found to provide fast (20 min) and efficient means to increase the molecular weight of PLLA (i.e. from 43 kDa to 96 kDa within 20 minutes,
Table S2, Entry 6). However, further heating of the reaction mixture in the presence of the crosslinker (from 96 kDa to 70 kDa after 48 hours,
Table S2, Entry 7), led to molecular weight decrease (
Figure S7) which is attributed to depolymerization of PLLA by either cleaving main polymer chain ester bonds [
46] and/or the amide bonds formed after the addition of the crosslinker [
47].
To get insight of the course of polycondensation and provide data for modeling, the progress of lactic acid polycondensation was monitored with GPC at the optimum conditions (
Figure 2 and
Figure S8, Table S2, Entry 2). In the initial stages of polycondensation (20 hours), oligomers with a degree of polymerization (DP) as low as 10 (M
n = 940) and dispersity (Đ) at 1.24 were detected. The formation of low molecular PLLA (M
n = 5766 kDa) was detected after 52.5 hours total polycondensation time (Đ = 2.40). Further heating (68 hours) led to significant increase of the average molecular weight PLLA (23 kDa, Đ = 2.1). The polymerization was terminated after 120 hours, providing high molecular weight PLLA (M
n = 81 kDa, M
w = 149 kDa and Đ = 1.84).
FT-IR spectroscopy revealed all the characteristic vibrations of PLLA, [
24,
48,
49,
50] i.e. the characteristic stretch of the C-H bonds of polymer methines at about 3000 and 2950 cm
-1 along with their deformation vibrations around 1380 and 1360 cm
-1, the strong stretching and bending vibration of ester carbonyl groups at ca. 1755 and 1265 cm
-1 respectively accompanied by the C-O stretching vibrations between 1080 and 1230 cm
-1 (
Table S3). Notably, terminal alkene C-H vibrations of terminal alkenes which could be a result of thermal degradation of PLLA were not detected at any case supporting the findings of NMR spectroscopy [
45]. As shown in
Figure S9, the IR spectra of moderate and high M
w PLLA were found to be almost identical (
Table S2, Entries 1 and 2 respectively), supporting the chromatographic data indicating the formation of oligomers at the early stages of polymerization. Additionally, the absorption peak at 1525.7 cm
-1 in the IR spectrum of PLLA synthesized in the presence of the crosslinker (
Table S2, Entry 7) was attributed to the N-H bending vibration of the amide bond formed after chain extension.
The thermal behavior of three PLLA samples selected on the basis of their molecular weight, i.e. low-, medium- and high-M
n PLLA (41, 70 and 81 kDa,
Table S2, Entries 1, 7 and 2 respectively) is shown in the thermogravimetric (TGA) and derivative mass loss (dTG) curves presented in
Figure S10. A small weight loss of about 0.5 wt. % seen in all samples below 100 °C can most probably be attributed to the evaporation of residual water. A single thermal decomposition step starting at around 170 °C and ending at 280 °C was seen on the TGA of both the low and medium-molecular weight PLLA. TGA showed two decomposition steps with no clear boundary for the high-molecular weight PLLA which was synthesized under optimal conditions (
Table S2, Entry 2) with the prominent step at ca. 276.1 °C and a secondary at 258.2 °C most probably stemming from the presence of a minute amount of lower molecular weight PLLA in the final product as also seen in GPC analyses (
Figure S8). The decomposition temperature of all PLLAs at 5 % and 50% mass loss (T
0.05 and T
0.50, respectively), as well as their maximum decomposition temperature (T
d, max) and rate were recorded (
Figure S10). The temperatures corresponding to 5% and 50 % mass loss were found to be 236.9 °C and 268.4 °C for the high-M
n polymer while the respective temperatures for the lower-M
n PLA’s were found to be 15-20 °C less. The maximum decomposition temperature (T
d, max) for the high-M
n polymer was 296.5 °C and for the low-M
n PLA was 278.1 °C and for the medium-M
n PLA was 283.7 °C. The maximum rate of the main decomposition step for the high-M
n polymer was determined with dTG analysis at 275.7 °C while for the medium and low molecular weight PLLAs was at 261.6 °C and 256.2 °C respectively (
Figure S10).
Differential scanning calorimetry (DSC) measurements of the annealed PLLA samples (
Table S2, Entries 1, 2 and 7) are presented in
Figure S11. The first thermal-cooling cycle was performed to reduce the thermal history of the samples, therefore the data collected by this process were not evaluated (
Figure S10 left,
Table S4) [
51,
52]. The glass transition temperature (T
g) of the medium-M
n PLLA (41 kDa) was 59.9 °C, 5.3 °C less than the T
g of high-M
n PLLA (81 kDa,
Table S4, Entries 1 and 2). The T
g of the PLLA synthesized in the presence of the crosslinker (medium-M
n) was calculated to be 61.1 °C (70 kDa,
Table S4, Entry 7). The polymer with the high-M
n (
Table S4, Entry 2) was found to exhibit the highest cold crystallization temperature (T
cc, 129.1 °C) and the lowest degree of crystallinity (16.9%), in agreement with what has been previously reported in literature [
53]. The degree of crystallinity was calculated using Equation 13 [
54]:
Interestingly, the DSC graphs provided two melting peaks for the polymers with medium molecular weight (40 and 71 kDa,
Table S2 Entries 1, 7), while a single melting point value was found for the higher molecular weight PLLA (81 kDa,
Table S2, Entry 2). The double melting behavior of PLA can be related with the formation of different crystal structures, known as α-form melting at higher temperature, and β-form melting at relatively lower temperatures [
55,
56,
57].
3.2. Modeling
Initial concentrations were determined according to the example shown in
Table S1. Initial mass of dry LA was estimated according to the following reasoning. Removal of residual water (0.93 g confirmed by
1H-NMR) leaves 7.11 g of dry LA. This is actually an upper bound, though, because of further conversion to lactide (by 5% wt.) and oligolactic acid (of known mass evidenced by
1H-NMR). Taking the conversion to lactide into account, we obtain an intermediate estimated dry LA mass of 6.75 g, under the assumptions that lactide does not react under the conditions of our experiments; and no conversion to oligomers has taken place. To determine a lower bound that takes the early formation of oligolactic acid into account, we account for its own share of reactive bonds that will be included as an appropriate correction. We consider an unknown average degree of polymerization, n
p, of the molecules formed according to the reaction
The material balance allows to determine an effective number of LA moles (i.e., ones that have the same number of reactive bonds as the actual mixture of LA and oligolactic acid) in the spirit of our simplified kinetic model. This is equal to nLA,0–nw where nLA,0 is the number of LA moles prior to the early formation of its oligomers; and nw are the moles of water formed (the same reasoning is actually applicable to any ensemble of reactions for all possible values of np). By subtracting the residual water from the 1.2 g total water removed during the initial stage of the experiment, and further considering a small amount of water produced during the partial conversion to lactide (0.0355 g) we end up with 0.2345 g of water produced by oligomerization. Then, we can find the estimated effective initial moles, and finally the corresponding estimated initial mass of dry LA (lower bound) equal to 5.58105 g.
Our calculations based on the measurements taken under optimal experimental conditions for all three estimated dry LA masses, as above, resulted in the rate and water removal constants summarized in
Table 1. The displayed figures represent a proposed range of values that accounts for the uncertainty of the initial dry LA mass associated with the side reactions at the beginning of the experiment. The predictions vary by about 6% to 12%, depending on LA’s initial mass, which, by comparison, was allowed a range of estimated values differing by up to 27%. The predicted number-average molecular weight is compared to the corresponding experimental measurements as well as their weight-average counterparts, in
Figure 3. The model reproduces the measured molecular weights with good accuracy. Calculations (not shown) have also been carried out for non-optimal experimental data sets, leading to equally good agreement between measurements and model predictions (with different, but generally similar optimal values of the rate and water removal parameters). Both experimental data and modelling predictions exhibit an almost steady increase in molecular weight with time following a relatively protracted period of slow growth – a rather counterintuitive finding as one would expect chain growth to gradually slow down and degree of polymerization to converge to a limiting value.
It can be argued that the actual trend, concealed by experimental errors and premature interruption of the process, should look more like a sigmoid function converging to a very high yet finite molecular weight. This scenario sounds more convincing when looking at the weight-average experimental values, which cannot be predicted by the simplified version of our model. Numerical predictions of number-average molecular weight, on the other hand, tend to grow linearly with time, as verified by running the simulation for longer time scales. This is indeed what happens when water concentration falls to negligible levels as we can easily verify by setting w = 0. Then, the set of rate equations is simplified to and , which is trivial to solve analytically yielding the solution and , and a predicted number-average molecular weight .
In other words, chain growth in the absence of water, is predicted to obey a linear trend and proceed uninterrupted until reaching the physical limit of depleting all available pairs of -OH and -COOH groups in the mixture (this highly improbable and idealized termination mechanism could involve the formation of one or more very long ring polymers neutralizing all remaining reactive groups).
In practice, the prevalent presence of very long chains in the mixture would bring to the fore mechanisms related to mass transport (low diffusivity) of the emerging polymer species, affecting the overall kinetics in manners that were not considered when setting up our model. The increasing impact of these mechanisms on the mixture’s kinetics, could indeed lead to sublinear dependence of molecular weight on time. Regardless of the above scenaria that concern time scales even longer than our experiments, it is useful to look at the change of water content with time, as a means of reevaluating the applicability of our assumptions about the constancy of the kinetic parameter governing water removal. Whether the scenario of negligible water concentration holds, can be examined by looking at the concentrations calculated as functions of time.
Concentrations of P-A (and Q-B) bonds, repeat units (equal, in number, to the ester bonds formed during polymerization) and water, as well as water’s mole fraction with time, are shown in
Figure 4. It can be readily seen that the calculated water content is far from negligible during the slow-growth period (roughly, the first 40 hours) but remains at very low levels afterwards, resulting in the predicted quasi-linear trend, which appears to match the experimental data quite well. On the other hand, the wide range of change in water content breaks the assumption of small, relatively stable water concentration made to justify a constant value for the water removal coefficient, k
w. Then, we expect the model to fail during the slow-growth period in one of the following ways: (i) water content in actual experiments remaining high for a prolonged period; in this case, the chemical equilibrium would shift towards depolymerization and our model would overestimate the molecular weight throughout the run; (ii) water content falling rapidly to very low levels; should this be the case, the equilibrium would shift towards polymerization early on, and our model would underestimate the molecular weight throughout the simulation; (iii) water content falling at a rate comparable to the one predicted numerically; in this last case, the model is expected to work well at long enough time scales but would tend to deviate more strongly (in terms of fractional errors) during the slow-growth period. This is indeed the case, as numerical predictions in
Figure 3 vary more smoothly than the measured data, and tend to overestimate molecular weight for up to about the first 50 hours. A more sophisticated model would account for a water removal coefficient, k
w, varying with water content, allowing a scenario in which water content would decrease slowly until falling under some ‘critical’ value, thus shifting equilibrium towards polymerization and accelerating chain growth.
Another interesting observation has to do with the rapid change of concentrations during the first few steps of the simulation. It is reminded that the concentrations refer to bonds that take part in the polymerization-depolymerization reaction; it is only at the beginning of the simulation that the concentration variable a coincides with the initial lactic acid concentration. The curves in
Figure 4 suggest that a large fraction of the lactic acid molecules (more than 90%) react immediately to form dimers and higher oligomers. The average degree of polymerization during this very early stage is about 15 (estimated as c/a, not quite far from the ten-mers measured by GPC during the first 20 hours of experiment under optimal conditions). The concentration of available pairs of reactive -OH and -COOH groups providing P-A and Q-B bonds, falls by that same ratio, and the corresponding amount of water is produced; these two factors immediately restrict chain growth until enough water is removed and the equilibrium is shifted towards polymerization.
Finally, it is worth discussing the role of the equilibrium constant k
eq = k
1/k
–1 as compared to the impact of the magnitude of the rate constants themselves (
Table 1). We multiplied the rate constants, k
1 and k
-1, either by the same factor or by different ones, and calculated three measures of our predictions conforming to the experimental data, i.e., the sum of square errors (our cost function) the sum of fractional errors squared (as percentages of the experimental values) and the sum of absolute fractional differences. It was found that when decreasing the rate constants in a uniform manner even by orders of magnitude (multiplying by down to 10
-3) the cost function remained largely unaffected (increased negligibly) whereas the other two measures improved (decreased) further albeit by a very small amount. In other words, the values of the rate constants themselves play a minor role in reproducing the polymerization data as long as they have the same ratio, i.e., equilibrium constant – which is not a surprising result.
It should be noted that the fractional measures could improve even further by decreasing the polymerization rate constant and/or increasing the depolymerization rate constant; however, the cost function would increase greatly as the model would fail to capture the long-term dynamics of the system (high molecular weights). In view of the above, the optimal values of the rate constants should be understood as upper bounds of a range of values of k1 and k–1, constrained by the condition of an optimal equilibrium constant, keq = k1/k–1; on top of them, comes the optimal value (Table 3) for the water removal parameter, kw, that fills the set of estimated values for the parameters of our model.