Next Article in Journal
Preparation of pH-Responsive PET TeMs by Controlled Graft Block Copolymerisation of Styrene and Methacrylic Acid for the Separation of Water–Oil Emulsions
Previous Article in Journal
Influence of the Molar Mass and Concentration of the Polyvinylpyrrolidone on the Physical–Mechanical Properties of Polylactic Acid for Food Packaging
Previous Article in Special Issue
Encapsulation of Perfluoroalkyl Carboxylic Acids (PFCAs) Within Polymer Microspheres for Storage in Supercritical Carbon Dioxide: A Strategy Using Dispersion Polymerization of PFCA-Loaded Monomers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Towards Greener Polymers: Poly(octamethylene itaconate-co-succinate) Synthesis Parameters

by
Magdalena Miętus
,
Tomasz Gołofit
and
Agnieszka Gadomska-Gajadhur
*
Faculty of Chemistry, Warsaw University of Technology, Noakowskiego 3 Street, 00-664 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Polymers 2025, 17(16), 2220; https://doi.org/10.3390/polym17162220
Submission received: 14 July 2025 / Revised: 6 August 2025 / Accepted: 12 August 2025 / Published: 14 August 2025
(This article belongs to the Special Issue New Progress of Green Sustainable Polymer Materials)

Abstract

A group of renewable, unsaturated resins from itaconic acid, 1,8-octanediol, and succinic anhydride were synthesized in a non-solvent and non-catalyst melt polycondensation reaction. The study addresses the need for sustainable polymeric materials suitable for additive manufacturing by investigating the influence of synthesis parameters—namely itaconic acid content, reaction time, and temperature—on the properties of poly(octamethylene itaconate-co-succinate) (POItcSc). The Box-Behnken mathematical planning method was applied to optimize the reaction conditions. The optimal synthesis conditions of POItcSc were achieved with an itaconic acid molar fraction = 0.50:0.50, reaction time t = 7 h, and reaction temperature T = 150 °C. The conversion of the carboxyl group (by titration) was 83.3%, and the maintenance of C=C bonds (by NMR) was 88.7%. Structural characterization confirmed the formation of the desired polymer through FTIR and 1H NMR analyses. Molecular weight (Mn = 1001 g/mol for an optimal product), thermal behavior (DSC, TG, DTG), and rheological properties (η = 14.4 and 3.6 Pa∙s for an optimal product at 25 and 36.6 °C) were systematically evaluated. The synthesized POItcSc resins were transparent and exhibited physicochemical properties favorable for extrusion-based 3D printing techniques such as Direct Ink Writing, offering a promising alternative to conventional petrochemical-based inks.

Graphical Abstract

1. Introduction

Most organic molecules are currently synthesized by petrochemical methods [1,2,3,4,5]. However, according to Anastas and Warner’s Green Chemistry principles, renewable materials should be used in syntheses, and toxic solvents and substrates should be limited [4]. Renewable carboxylic acids and their derivatives are commonly used as a source for building blocks with desired properties, to minimize the contribution of toxic materials in industrial applications [3,5]. These primarily include adipic acid, sebacic acid, succinic anhydride, and itaconic acid [6]. These are mainly used in the medical field [3,7]. One of the latest developments is the use of 3D printing in medicine. Three-dimensional printing relies on UV crosslinking of the printed model using mostly toxic and non-renewable acrylic compounds. These are structurally similar to itaconic ones (Figure 1).
Itaconic acid (IA) is an unsaturated dicarboxylic acid obtained from renewable resources [8]. It is commercially produced using an inexpensive method of fermentation of carbohydrates (for instance, glucose, sucrose, and starch) [8,9,10]. It is performed by the Aspergillus terreus and Aspergillus itaconicus fungi [8,11]. The US Department of Energy describes IA as one of the 12 most valuable organic chemicals [5,12]. This is because of its crucial properties: non-toxicity, biodegradability, and biocompatibility [9]. Furthermore, itaconic compounds exhibit antibacterial and anticancer properties. Because of its advantages, the global demand for IA in 2026 is expected to be more than $110 million [11]. The most crucial feature of itaconic compounds is the existence of a C=C multiple bond in their lateral chain. Its presence makes it possible to perform post-polymerization reactions—that is, the Michael addition or UV-activated photopolymerization reactions [13]. Such a structure of itaconic compounds allows them to be used to synthesize resin-like polymers (for Direct Ink Writing (DIW) 3D printing methods) and produce drugs with anticancer and antiviral properties [14,15,16,17,18]. However, the sterically uncrowded multiple bond can cause problems in synthesizing macromolecular itaconic compounds. During the reaction of obtaining polyesters based on itaconic compounds and compounds having a hydroxyl group in their structure, a Michael reaction can occur (in other words, an oxo-Michael or Ordelt reaction) (Figure 2) [1,19].
Itaconic compounds can isomerize to less reactive isomers: mesaconic and citraconic ones. This, and the radical polymerization reaction of the itaconic unit (Figure 2), are the remaining undesirable reactions of itaconic compounds [17]. The contribution of the isomerization reaction can be reduced by running reactions at temperatures not higher than 150 °C [17]. To reduce the contribution of the undesirable radical polymerization reaction, inhibitors can be incorporated into the reaction system. The most commonly used are 4-methoxyphenol (MEHQ), butylated hydroxytoluene (BHT), phenothiazine, and hydroquinone [20,21,22,23]. However, they are not biocompatible, which makes their use in medicine impossible.
As mentioned, itaconic compounds can be mixed with diols to obtain polyesters. The synthesized aliphatic polyesters exhibit biocompatibility [10]. One such diol is the terminal 1,8-octanediol (1,8-OD). It is the longest water-soluble aliphatic diol [24]. It is commonly used in medicine as a part of medical devices for vascular purposes and to obtain scaffolds [1,25,26,27,28,29,30]. However, as it is challenging to efficiently conduct the synthesis reaction between the itaconic compound and 1,8-octanediol, we decided to enrich the product with succinic anhydride (SAn) [31]. Using saturated succinic anhydride, it is possible to reduce the proportion of the undesirable Ordelt reaction and, consequently, the polycondensation product’s crosslinking degree [32]. Succinic anhydride is a derivative of succinic acid (SA) [6]. It can be produced in the biological fermentation conducted by microorganisms [33]. It is commonly used to obtain biodegradable polymers—for instance, polybutyrate succinate (PBS) and polyamides (Nylon®x,4) [33]. As SA and its derivatives are saturated organic compounds, their use makes it possible to extend the polymer chain without the risk of side reactions [34].
In this study, to meet today’s environmental demands, we developed the synthesis of a fully biodegradable and renewable itaconic copolyester—poly(octamethylene itaconate-co-succinate) (POItcSc)-copolymer from itaconic acid, succinic anhydride, and 1,8-octanediol, for prospective application in additive manufacturing (similar to poly(butylene adipate-co-terephthalate) (PBAT)) [3].

2. Materials and Methods

2.1. Materials

The following materials were used: itaconic acid (≥99%, Sigma Aldrich, St. Louis, MO, USA), 1,8-octanediol (98%, Angene, Nanjing, China), succinic anhydride (≥99, Gdańsk, Poland), methanol (Chempur, Piekary Śląskie, Poland), 1M aqueous sodium hydroxide solution (Chempur, Piekary Śląskie, Poland), 1M hydrochloric acid solution (Chempur, Piekary Śląskie, Poland), chloroform (Chempur, Piekary Śląskie, Poland), Hanus reagent (Hempur, Piekary Śląskie, Poland), 10% KI solution (Chempur, Piekary Śląskie, Poland), 0.1 M sodium thiosulfate (Chempur, Piekary Śląskie, Poland), starch indicator (Chempur, Piekary Śląskie, Poland), deuterated DMSO (Deuteron GmbH, Kastellaun, Germany), tert-butanol, dichloromethane (HPLC grade, Sigma Aldrich, St. Louis, MO, USA), n-hexane (POCH, Gliwice, Poland), toluene (Chempur, Piekary Śląskie, Poland), diethyl ether (Chempur, Piekary Śląskie, Poland), ethyl alcohol (POCH, Gliwice, Poland), dichloromethane (POCH, Gliwice, Poland), THF (POCH, Gliwice, Poland), chloroform (Chempur, Piekary Śląskie, Poland), ethyl acetate (POCH, Gliwice, Poland), 1,4-dioxane (POCH, Gliwice, Poland), methanol (Chempur, Piekary Śląskie, Poland), acetone (POCH, Gliwice, Poland), acetonitrile (POCH, Gliwice, Poland), 1-butanol (POCH, Gliwice, Poland), DMF (POCH, Gliwice, Poland), and DMSO (Chempur, Piekary Śląskie, Poland).

2.2. Polyester Synthesis

The syntheses were carried out in the Mettler Toledo MultiMax parallel reactor system (Schwerzenbach, Switzerland) in the Hastelloy reactors. The reactants 1,8-octanediol, itaconic acid, and succinic anhydride were used as supplied, without prior preparation.
To obtain the IA fraction to SAn of 0.35, 0.50, and 0.65, the used weight of the reactants was as follows: 1,8-octanediol (22.78 g; 22.38 g; 22.00 g), itaconic acid (7.09 g; 9.96 g; 12.73 g), and succinic anhydride (10.13 g; 7.66 g; 5.27 g). The molar ratio of IA + SAn:1,8-OD was 1:1. The substrates were weighed and placed in the metal and non-transparent reactor, and the used reactants weighed 40.00 g. The reactor was equipped with a mechanical stirrer, a temperature sensor, and a Dean–Stark apparatus.
The reaction procedure was as follows: In the first stage, the reactants were heated for 15 min to x3 temperature. The temperature was held constant for x2 h. Then, 30 min after the start of this phase, the reduced pressure was switched on (200 mbar). Finally, the mixture was cooled down to 40 °C for 15 min. The reaction system was stirred throughout the reaction (200 rpm).

2.3. Titration Analysis

To calculate the Acid Number (ANtit) value, about 0.50–1.00 g of the sample was weighed and dissolved in 25.00 mL of methanol. Five drops of indicator, thymol blue, was added to each sample. After dissolution, each sample was titrated with 1 M NaOHaq until the change in color from yellow to blue was observed. The obtained ANtit is a mean of three determinations for each sample of POItcSc. A blank test was performed under the same conditions. To calculate the acid number, the following formula was used:
ANtit [mgKOH/gsample] = ((VV0) × MNaOH × 56.1)/m
where
V—the volume of 1 M NaOH solution used to titrate the investigated sample [cm3];
V0—the volume of 1 M NaOH solution used for blank titration [cm3];
MNaOH—the titer of the solution for the titration (1 M);
56.1—the molar mass of KOH [g/mol];
m—the weight of the investigated sample [g].
To calculate the conversion of carboxyl groups (%convCOOH tit) in the structure of the synthesized product, the following formula was used:
%convCOOH tit = (2 × nIA − ((ANtit/1000)/56.1 × w)/(2 × nIA)) × 100%
where
nIA —the amount of itaconic acid used in the synthesis [mol];
w—the weight of the substrates in the reaction system [g].
To calculate the Ester Number (ENtit), about 0.20–0.50 g of the sample was weighed and dissolved in a solution of 15.00 mL of methanol and 20.00 mL of 1 M NaOHaq. Then, the solutions were refluxed for one hour at around 120 °C. Next, the solution was cooled down to room temperature. Five drops of phenolphthalein were added to each trial. The samples were titrated with 1 M HClaq until discoloration of the pink solution. A blank test was performed under the same conditions. The obtained ENtit is a mean of three determinations for each sample of POItcSc. To calculate the acid number, the following formula was used:
ENtit [mgKOH/gsample] = (((V0V) × MHCl × 56.1)/m) − ANtit
where
V—the volume of aqueous 1 M HCl solution used to titrate the investigated sample [cm3];
V0—the volume of aqueous 1 M HCl solution used for blank titration [cm3];
56.1—the molar mass of KOH [g/mol];
m—the weight of the investigated sample [g].
The final result is the average of three determinations.
To calculate the esterification degree (EDtit), the following formula was used:
EDtit [%] = ENtit/(ENtit + ANtit) × 100%
where
ENtit—the ester number from titration;
ANtit—the acid number from titration.
To calculate the Iodine Number (INtit), approximately 0.50 g of the investigated sample was weighed and dissolved in a solution of 10.00 mL of chloroform and 15.00 mL of Hanus reagent. After mixing the flask contents for 30 min in a dark place, 15.00 mL of 10% KI solution and 50 mL of distilled water were added. Then, the solution was titrated with 0.1 M sodium thiosulfate solution until a bright orange color was observed. Next, 5 mL of starch indicator was added to the flask to obtain a dark blue color in the mixture. Then, the solution was titrated with 0.1 M thiosulfate solution until it discolored. The obtained INtit is a mean of two determinations for each sample of POItcSc. To calculate the iodine number, the following formula was used:
INtit = 1.269 × ((a − b)/c))
where
a—the volume of sodium thiosulfate solution (0.1 M) used for blank titration [cm3];
b—the volume of sodium thiosulfate solution (0.1 M) used to titrate the sample [cm3];
c—the weight of the investigated sample [g].
To calculate the percentage of unreacted C=C double bonds (%C=C IN tit) in the structure of the POItcSc, the following formula was used:
%C=C IN tit = ((((INtit/100)/253.81) × w)/nIA) × 100%
where
253.81—molar mass of the molecular iodine (I2) [g/mol].

2.4. Fourier Transform Infrared (FT-IR) Analysis

FT-IR spectra were recorded using an ALPHA spectrometer (Bruker, Berlin, Germany), ranging from 400 to 4000 cm−1, with 32 scans for each sample.

2.5. Nuclear Magnetic Resonance (NMR) Analysis

1H NMR and 13C NMR measurements of the obtained polyesters were performed on an Agilent 400 MHz NMR spectrometer with deuterated dimethyl sulfoxide (DMSO-d6) as a solvent and tert-butanol (t-BuOH) as the internal chemical shift standard.

2.6. Gel Permeation Chromatography (GPC) Analysis

The number-average molecular weight (Mn), weight-average molecular weight (Mw), and dispersity index (DI) of the synthesized products were determined by Size Exclusion Chromatography (SEC) carried out on an Agilent 1260 Infinity System (Santa Clara, CA, USA). The setup included an isocratic pump, an autosampler, a degasser, a thermostated column, and a differential refractometer (MDS RI Detector). Data acquisition and analysis were performed using Addon Rev. software (version B.01.02, Agilent Technologies, Santa Clara, CA, USA). Mw values were calculated based on calibration with linear polystyrene standards (580–128,900 g/mol). The separation process incorporated a pre-column guard (3 μm, 50 × 7.5 mm) along with two analytical columns: PLgel MIXED-D (5 μm, 300 × 7.5 mm) and PLgel MIXED-E (3 μm, 300 × 7.5 mm). The analyses were conducted at 30 °C using dichloromethane (HPLC grade, Sigma Aldrich, St. Louis, MO, USA) as the mobile phase, with a 0.8 mL/min flow rate.

2.7. Thermal Analysis

DSC (Differential Scanning Calorimetry) measurements were conducted on the Q2000 DSC analyzer (TA Instruments, Eschborn, Germany). The DSC procedure was as follows. In the first step, the sample (weighing approximately 10 mg) was weighed and placed in the crucibles, which were closed with lids with holes. Then, the chamber was sealed, and the measurement was performed. At first, the sample was cooled to −90 °C. Then, it was heated to 250 °C (10 °C/min step). In the next step, the sample was cooled to −90 °C. In the final stage, the sample was heated again to 250 °C. DSC thermograms were analyzed using TA Instruments Universal Analysis 2000 software. The glass transition temperature was determined as a midpoint temperature. The cold crystallization temperature was defined as the peak temperature. DSC analyses were conducted in the nitrogen flow (50 mL/min).
TG (Thermogravimetry) measurements were conducted on the SDT Q600 analyzer (TA Instruments, Eschborn, Germany). The weight loss of the samples (weighing approximately 10 mg) was analyzed using the temperature range from room temperature to 500 °C (10 °C/min step). TG analyses were conducted in the nitrogen flow (100 mL/min).

2.8. Rheological Analysis

The MCR 301 rheometer (Anton Paar, Graz, Austria) was used to measure the viscosity of POItcSc products. The plate–plate method was applied. The diameter of the moving plate was 25 mm. Every experiment lasted for 8 min (4 min/stage). In the first stage, the shear rate increased from 0.01 cm−1 to 10 cm−1. The shear rate decreased from 10 cm−1 to 0.01 cm−1 in the second and final stage. The number of measurement points per decade was 20. To approximate the measurement curve, the Casson model was used:
τ 1 / 2   =   τ 1 / 2 0   +   ( η p   ×   γ ˙ ) 1 / 2
where
τ—shear stress [Pa];
τ0—shear limit (yield stress) [Pa];
ηp—rheological parameter (plastic viscosity) [Pa∙s];
γ ˙ —shear rate [s−1].
The graphs of the relationship between the tangential stress’s square root and the shear rate’s square root were obtained using the data. For each graph, the regression equation was calculated. The straight line’s directional coefficient was the plastic viscosity’s square root raised to the power of 2. The estimated value was the plastic viscosity.
Amplitude sweep testing of the POItcSc was conducted at 25 °C and 36.6 °C. Three different frequencies were investigated: 0.1 Hz, 1 Hz, and 10 Hz. Each measurement preceded a phase in which the sample was subjected to a constant, lowest strain (γ = 0.005%). Then, measurements were carried out at increasing strain (γ = 100%). The range of linear viscoelasticity of the POItcSc was determined.
Frequency sweep testing of the POItcSc was carried out at 25 °C and 36.6 °C in the range from 0.01 Hz to 100 Hz, with a constant strain of 0.1%. The measurement phase was preceded by a pre-shearing phase, in which the test sample was subjected to sweeping at a continuous frequency of 0.01 Hz for 120 s at a constant strain of 1%.
To determine the dependence of the form modulus (G′) and loss modulus (G″) as a function of temperature, POItcSc was heated at a rate of 3 °C/min from 15 °C to 195 °C, at a constant strain of 0.1% and a constant frequency of 1 Hz.

2.9. Viscosity-Visual-Utility (VVU) Analysis

The numerical interval scale characterized every synthesis product from the mathematical model for its specified properties (Table 1) at two temperatures: Troom and T = 36.6 °C. The numerical interval scale was converted into a percentage scale. The highest number of possible points, 32, was defined as 100%, and the lowest number of possible points, 10, was described as 0%.

2.10. Statistical Analysis

All results were analyzed statistically using Statistica software (Version 13.3). Comparisons between groups were performed using one-way ANOVA. Differences were considered statistically significant when p-values < 0.05.

3. Results and Discussion

The main objective of the performed experiments is an in-depth investigation of the poly(octamethylene itaconate-co-succinate) synthesis products obtained under different conditions in the polyesterification process [35]. In the research, no catalyst, inhibitor, or solvent was used to obtain the final product. We obtained poly(octamethylene itaconate-co-succinate) in the polycondensation reaction between itaconic acid, succinic anhydride, and 1,8-octanediol (Figure 3). The synthesis was performed without the use of a catalyst, as it can cause difficulties in the intended final application and can contribute to the presence of an undesired Ordelt reaction [16,17,36].

3.1. Statistical Analysis

To optimize the synthesis of POItcSc, the Box–Behnken mathematical plan was used, as it allows avoiding experiments in extreme conditions (only +1 or −1 values) and reduces research time (fewer experiments) [37]. Fifteen experiments were conducted. Three of those fifteen experiments were conducted under the same conditions to examine the repeatability of the results and the experimenter’s reliability.
The three chosen input variables (x) were as follows:
x1—molar fraction of itaconic acid in the reaction system (IA molar fraction) (primarily for optimization of molecular weight, rheological and thermal properties, transparency and homogeneity of the product, and contribution of side reactions);
x2—time of the POItcSc synthesis (t), [h] (to define the variability of carboxyl group conversion, the contribution of side reactions, and the consistency of the product);
x3—temperature of the POItcSc synthesis (T), [°C] (to determine whether it affects the contribution of side reactions and the consistency of the product).
The three output variables (y), most relevant regarding the subsequent use of the obtained macromolecular product, were selected. These were, in sequence,
y1—percentage conversion of carboxyl groups -COOH (%convCOOH tit) (calculated from the ANtit), [%];
y2—percentage of unreacted unsaturated C=C double bonds (%C=C 1H NMR) (calculated from the 1H NMR spectra analysis), [%];
y3—Viscosity-Visual-Utility analysis (%VVU), [%].
The selection of these parameters as output variables was clarified in a previous article [37].
To facilitate the understanding of the use of input/output variables, the POItcSc synthesis model is presented as a “black box” in Figure 4.
Table S1 presents the coded values of the input variables used in the Box–Behnken plan. A temperature of 150 °C was selected as the maximum temperature for POItcSc syntheses to minimize the contribution of undesirable side reactions [38,39,40,41]. A summary of the coded values of the input variables and the experimental and calculated values of the output variables is presented in Table 2.
According to the NMR (Nuclear Magnetic Resonance) analysis results (Table S2), the contribution of the mesaconic compound in the reaction system is mild (from 0.5 to 3.7%), thanks to the use of a synthesis temperature no higher than 150 °C [16,21]. However, not all received products had the desired consistency (Figure S1). They were in the form of wax, which resulted from a synthesis temperature that was too low and a low value of the esterification degree (Table S2). It should be noted that a slight increase in the temperature of the sample (to 36.6 °C) contributed to obtaining a product with the desired resin consistency. This, and the results of the viscosity (Table S2) for some of the products (9, 11, 12, 14, 15), lead to the conclusion that POItcSc is a suitable macromolecular compound for use as, for instance, 3D printing ink in the DIW method.
Every product obtained in the conducted experiments was investigated for its molecular weight by GPC-SEC (Gel Permeation Chromatography-Size Exclusion Chromatography) experiments (Table S2). The resultant elugrams are usually used for polycondensation products (Gaussian distribution) [41]. The obtained Mw (weight-average molecular weight) of the products is in the range of 821 to 3370 (g/mol), which contributes to the number of repeating units from 4 to 16 (assuming that the repeating unit consists of one itaconate, succinate, and diol unit). The dispersity index (DI) ranges from 1.7 to 4.0, which is anticipated for the polycondensation reactions [42]. Compared with other macromolecular compounds with itaconate segments for additive manufacturing purposes, the obtained POItcSc is characterized by similar values of molecular weights determined by GPC analyses [1,42,43,44,45]. Structural differences cause the main differences in the assigned values.
To define which regression equation coefficients are significant, the t-Student test was used (∣tcalculated∣ > tcritical). The probability level (p-value) was 5%, and the tcritical = 4.303. The F-Snedecor test (Fcritical (0.05; 3; 2) = 19.16) was performed to determine the adequacy of the regression equations.
A Pareto Chart (Figures S3–S5) was presented as a graphical representation for the Analysis of Variance (ANOVA) for every output variable. Below, all regression equations, which describe every output variable, are presented (where the green color corresponds to the relevant coefficients):
y1 = 73.48.56 × x1 + 11.1 × x2 + 9.97 × x3 + 3.95 × x1 × x2 + 0.184 × x1 × x3 + 0.398 × x2 × x3 + 1.51 × x12 − 0.603 × x22 + 9.97 × x32
y2 = 79.2 − 0.743 × x1 − 1.14 × x2 + 6.53 × x3 + 0.558 × x1 × x2 − 2.30 × x1 × x3 + 4.13 × x2 × x3 − 4.31 × x12 − 2.36 × x22 + 0.103 × x32
y3 = 80.2 + 5.47 × x1 + 0.391 × x2 + 0.391 × x3 + 3.13 × x1 × x2 − 1.56 × x1 × x3 + 0.781 × x2 × x33.00 × x12 + 4.04 × x225.34 × x32
The graphical representations of the above regression equations are shown in Figure 5.
For the y1 variable, every input variable is significant (∣tcalculated x1∣ = 8.31; (∣tcalculated x2∣ = 10.80; (∣tcalculated x3∣ = 9.68), as is the linear relationship between the x1 and x2 (∣tcalculated x1x2∣ = 2.71) input variables. However, the inclusion of other investigated variables, their powers, and their relations in the model increase the coefficient of determination R2 value from 0.96 to 0.98. This means that approximately 98% of the variability of the %convCOOH tit originates from the variability of the investigated input variables. As the Fcalculated for y1 (17.66) variable is smaller than Fcritical, the regression equation for the y1 variable can be defined as sufficient. The experimental and calculated values of the y1 variable differ only by ±4.1 percentage points (Table 2), which shows the perfect fit for the used model. Reaction time and temperature play a significant role in the conversion of carboxyl groups. The longer the reaction time, and with higher temperature, the greater the %convCOOH (>80%).
For the y2 variable, only the x3 input variable plays a significant role (∣tcalculated x3∣ = 3.14) in the regression equation. Using only this coefficient for the model, the coefficient of determination R2 has a low value (0.48). This means many other factors not included in the research affect the %C=C 1H NMR value (for instance, pressure in the reaction system or factors beyond the researcher’s control). The inclusion of insignificant coefficients increases the R2 value to 0.75. The Fcalculated for the y2 variable is 6.80, meaning there is no evidence to reject the hypothesis that the presented equation is adequate. However, it should be noted that other factors affecting the value of the y2 variable were not considered. Similar to the %convCOOH tit values, temperature has the most crucial role. Although side reactions involving multiple bonds are expected to occur more frequently at higher temperatures, the opposite trend was observed. This may be due to the use of two acidic monomers (itaconic acid and succinic anhydride), which contributed to the reduction of the unfavorable effect of temperature on the occurrence of reactions by multiple bonds of the itaconic unit. This was confirmed by the result of the t-Student test, in which the power of the input variable x1 shows almost the highest significance, right after the variable x3. The experimental and calculated values of the y2 variable differ only by ±5.3 percentage points (Table 1), which represents a perfect fit for the used model.
For the y3 variable, the IA molar fraction (∣tcalculated x1∣ = 9.30), squared temperature (∣tcalculated x32∣ = 6.17), time (∣tcalculated x22∣ = 4.66), IA molar fraction (∣tcalculated x12∣ = 3.46), and IA molar fraction with time linear relationship (∣tcalculated x1x2∣ = 3.76) have a crucial role in obtaining a product with desired properties for 3D printing. Because the incorporation of other coefficients into the model increases the R2 value (from 0.94 to 0.97), they were included in the model. For the y3 variable, Fcalculated = 0.75, meaning that there is no evidence to reject the hypothesis that the equation is adequate. The experimental and calculated values of the y3 variable differ only from -2.1 to 1.2 percentage points (Table 2), which indicates a perfect fit for the applied model. The product with the highest %VVU can be obtained in the reaction with the highest IA molar fraction to the SAn (0.65:0.35), using the longest examined reaction time (t = 7 h). Then, the %VVU can exceed 85.0%. The higher the IA molar fraction, the higher the %VVU values.
The least squares method (Statistica software) was used to specify the optimal conditions of the POItcSc synthesis. For this, the utility profile function software was used (Figure S6), and the output variables were programmed as variables with low, medium, and high utility (Table S3). To obtain a product with the highest values of investigated output variables (Figure S7), the reaction has to be performed under the following conditions: IA molar fraction 0.50:0.50 (x1), reaction time = 7 h (x2), and temperature = 150 °C (x3). The output variables’ values for the optimal conditions of the product are shown in Table 3.
Considering that the experimental and determined values of the output variables vary negligibly, and the model’s utility is 84.1%, a remarkable fit is demonstrated between the statistical model and reality. The optimal product is characterized by one of the higher molecular weights obtained in the performed experiments—the Mw was 3129 g/mol, with a dispersity index of 3.1. More extensive product characteristics are shown in Table S2.

3.2. FT-IR Spectroscopy and NMR Analysis

The structure of the substrates and POItcSc product was confirmed by FT-IR (Fourier Transform Infrared) (Figure 6), 1H NMR (Figure 7), and 13C-NOE NMR (Figure S8).
The O-H bond stretching vibrations can be seen in the 3550–3000 cm−1 range. These correspond to the substrates—1,8-octanediol (strong and sharp signal) and itaconic acid (strong and wide signal). For the POItcSc product, there is no visible signal from the O-H group, meaning that the product is not mostly terminated with diol. Furthermore, only a small proportion of unreacted substrates are present in the final product. The 29400–2850 cm−1 bands correspond to the methylene groups’ stretching vibrations. The most important signals can be observed in the range below 2000 cm−1. The bands in the 1780–1715 cm−1 range correspond to the stretching vibrations of the carbonyl groups from α and β-unsaturated esters. In the 1680–1620 cm−1 range, stretching vibrations of the C=C double bond can be seen. The presence of polyester can be confirmed by the signals in the range of 1220–1140 cm−1 (stretching vibrations of the acyl groups) and 1050–1030 cm−1 (stretching vibrations of the acyl groups). In the 1000–980 cm−1 range, a signal corresponds to the C-CO-O-CO-C stretching vibrations from the substrate, succinic anhydride.
The 1H NMR (Figure 7) was used in calculations (see Supporting Information). Figure S9 presents the assignment of H and C atoms to the signals on NMR spectra.
In each obtained spectra, the characteristic resonance signals corresponding to the C=CH2 protons of itaconic acid in the polycondensation products can be observed at 6.1–6.4 ppm (e′1) and 5.7–6.1 ppm (e′2). This means that there are unreacted double bonds, which are available for subsequent photo-crosslinking. The signals corresponding to the mesaconic isomer are visible on every 1H NMR. However, they are weak, and the contribution of the mesaconic isomer is in the range of 0.5% to 3.7%. Signals that correspond to the protons corresponding to the occurrence of undesired radical polymerization (in the range of 2.4% to 8.8%) and Ordelt reactions (in the range of 8.4% to 34.4%) were also observed.
Similar to the 1H NMR spectra, the 13C NMR-NOE spectra (Figure S8) confirm the successful polycondensation reaction between IA, SAn, and 1,8-OD. The presence of the signals from the carbon atoms corresponding to the C=C double bonds in the polycondensation product (signal E′) confirm that the double bonds of IA were not significantly affected by side reactions.

3.3. Thermal Analysis

DSC (Differential Scanning Calorimetry) was used to examine the thermal properties of the synthesized poly(octamethylene itaconate-co-succinate). Figure 8 shows the characteristic thermal transitions for the POItcSc polymer.
Figure 8 shows some characteristic temperature transitions for POItcSc (Table S6). The glass transition temperature of POItcSc can be seen during the first (Tgh1, A) and second (Tgh2, F) heating cycle. Furthermore, the cold crystallization temperature (Tcch1, B), crystallization temperature (Tcc1, E), and melting temperature (Tmh1, C and Tmh2, G) of POItcSc can also be observed. The determined Tg values are lower than for the previously characterized poly(tetramethylene itaconate) (PBItc) [38]. Using a diol with a greater proportion of -CH2- groups contributes to higher mobility of POItcSc chains [42,46]. The presence of Tcc and Tm indicates a semicrystalline polymer nature, mainly due to the use of 1,8-OD. The presence of two endothermic peaks labeled as C indicates the melting of the crystalline phase of POItcSc fractions of different molecular weights. The presence of POItcSc fractions with varying chain lengths was confirmed by GPC analysis, where DI = 3.1. An endothermic peak D is also visible during the first heating cycle, corresponding to the melting of unreacted 1,8-OD. The absence of such a peak during the second heating cycle indicates that 1,8-OD underwent the reaction during the first heating cycle.
The thermal stability of the POItcSc polymer was analyzed by thermogravimetric analysis (TG) (Figure 9).
The characteristic degradation temperatures are summarized in Table S6. As shown on the DTG (Derivative Thermogravimetry) curve, POItcSc exhibits two-stage weight loss, similar to other macromolecular compounds within itaconate units [47]. A minor weight loss at about 100 °C is due to water removal from the reaction system [48]. The first significant mass loss occurs at about 270 °C, which is associated with the degradation of intermediate decay products [48]. The maximum weight loss is observed at 405.5 °C (Td max). At this temperature, the dissociation of ester bonds occurs [49,50,51]. Then, the dehydrogenation and dissociation of ashes formed at lower temperatures occur [52]. Such high degradation temperatures indicate that a polymer with high thermal stability has been synthesized [53]. Furthermore, the obtained POItcSc polycondensation product shows higher thermal stability than other previously synthesized itaconate macromolecular compounds, namely BItc [38]. This is due to the use of a longer aliphatic diol than 1,4-butanediol. As the thermal stability is higher, the POItcSc final product for application will tend to maintain its physical and mechanical properties at high temperatures for a longer time than PBItc [54].
Considering the applicability of the obtained POItcSc resin, besides the possibility of its use as an ink for 3D printing by DIW or DLP (Digital Light Processing), based on thermal analysis, it is possible to consider it for obtaining thermoplastic polyester films. This is due to a low melting temperature range, rapid crystallization, and high thermal stability, which suggests good processability [55].

3.4. Rheology Analysis

The resin produced in the optimal conditions was analyzed for its rheological properties. A shear-thinning product was obtained (Figure 10A) because the product’s viscosity decreases as the shear rate increases. As the shear rate increases, the polymer aggregates disintegrate, and the polymer chains orient and align parallel to the shear direction. From this, it can be specified that the received resin will be suitable for 3D printing with DIW methods, for example. It will easily pass through the printer’s nozzle under pressure, and once it is placed on the table, it will retain its shape. Furthermore, with the higher measuring temperature, the viscosity of the investigated macromolecular product decreases (at 25.0 °C = 14.4 Pa∙s, at 36.6 °C = 3.6 Pa∙s). As the temperature rises, the movement of the polymer chains is facilitated. Thus, the obtained product can again be considered suitable for applications requiring material injection molding, e.g., for DIW 3D printing methods. The obtained product has higher viscosity than the other resins with itaconic fragments (η = 0.27–1.66 Pa∙s) [42]. Its reduced viscosity was due to the use of diluents. In the case of POItcSc, its addition was avoided because of the toxicity of those most commonly used ones [42,56]. In the case of PIDDOL (poly(ester thioether) based on 1,12-dodecanediyl bis(methyl itaconate) and linalol), a slightly higher viscosity value for the same measurement conditions was found (η = 16.7 ± 1.8 Pa∙s), which was the effect of the use of somewhat longer diol-1,12-dodecanodiol [44].
Thermorheological analysis of POItcSc (Figure 10B) confirmed that the glass transition is observed at a temperature lower than the minimum range of the used measurement apparatus. The intersection of the curves corresponding to the elastic modulus (G′) and loss modulus (G″) occurs at 86.2 °C. The material transitions to an elastic (flexible) state at this temperature. Then, the crosslinking process of POItcSc begins. Considering the potential application of the obtained POItcSc resin, it should be noted that when printing with it, the head temperature must not be greater than the limiting temperature of 86.2 °C.
We performed amplitude sweep tests to determine the range of linear viscoelasticity (LVE) for the obtained POItcSc (Figure 11A,B). Regardless of the experimental temperature, there is an oscillation of G′ and G″ values for low strain values. This indicates the occurrence of the reorganization of polymer chains. With higher temperatures, it occurs more easily. For the tested strain range, the value of G′ < G″ indicates that the viscous values of POItcSc are dominant (compared to elastic properties) [57]. Such behavior is desirable for DIW 3D-printed materials. For the tested product, the range of LVE is present for a strain of 0.1%.
A frequency sweep study was conducted for POItcSc (Figure 11C,D). Frequency sweep tests confirmed that viscous properties dominate over elastic properties (G′ < G″) for POItcSc. This indicates that a non-crosslinked or low-crosslinked product was obtained. The intersection of the curves occurs at the higher measurement temperature (36.6). In other words, in the case of 3D printing POItcSc resin with a heated printer extruder, the resin would pass through the printer nozzle more easily and hold its shape better immediately after printing. This means that carrying out 3D printing of POItcSc at an elevated temperature is beneficial.

4. Conclusions

We successfully synthesized a set of biobased poly(octamethylene itaconate-co-succinate) polyesters from itaconic acid, 1,8-octanediol, and succinic anhydride in a catalyst-free and solvent-free melt polycondensation reaction. For this, we used the Box–Behnken mathematical model to optimize the synthesis of POItcSc. We determined how the itaconic acid molar fraction in combination with succinic anhydride, the reaction time, and temperature influence the properties of the obtained polycondensation product. The results show that the presence of succinic anhydride as a co-monomer reduced the percentage of the undesired side reactions. The Mn of the obtained products was defined using GPC. Every product was examined for its rheological properties. The product obtained in optimal conditions was additionally characterized using DSC, TG (DTG), and rheological (oscillation) methods. The performed statistical analysis revealed that to receive a product with the best properties in terms of potential use, the reaction should be conducted under the following conditions: IA molar ratio = 0.50:0.50, reaction time = 7 h, and reaction temperature T = 150 °C. The optimal POItcSc product is a semi-crystalline polyester with proper viscosity and consistency (resin) for 3D printing.
As for now, POItcSc may be considered as a potential ink for 3D printing purposes (for instance, in medicine). However, extensive studies on the UV-crosslinking of POItcSc and the properties of the UV-crosslinked films need to be performed in the future.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym17162220/s1, Figure S1: Consistency of the obtained POItcSc products at 25 °C; Figure S2: Consistency of the obtained POItcSc products at 36.6 °C; Figure S3: Pareto Chart of Standardized Effects for the %convCOOH tit (y1) variable (the red line refers to the limit, beyond which the coefficient of the regression equation becomes significant); Figure S4: Pareto Chart of Standardized Effects for the %C=C 1H NMR (y2) variable; Figure S5: Pareto Chart of Standardized Effects for the %VVU (y3) variable; Figure S6: Profile of the approximated values of input variables and utility of the used mathematical model; Figure S7: Consistency of the optimal product; Figure S8: 13C NMR spectra of poly(octamethylene itaconate-co-succinate); Figure S9: Assignment of protons and carbon atoms to the corresponding signals on the 1H NMR and 13C NMR spectra; Table S1: Coded input variables in the Box–Behnken plan for the PISO synthesis optimization; Table S2: Characterization of the polycondensation products; Table S3: Values of the output variables to generate the response utility profile; Table S4: Significance test of regression equation coefficients for the investigated output variables (numbers are written to four significant digits); Table S5: Model adequacy test for the investigated output variables; Table S6: Thermal analysis of POItcSc polycondensation product.

Author Contributions

Conceptualization, M.M.; methodology, M.M.; validation, M.M. and A.G.-G.; formal analysis, M.M. and A.G.-G.; investigation, M.M.; resources, T.G. and A.G.-G.; data curation, M.M.; writing—original draft preparation, M.M.; writing—review and editing, M.M.; visualization, M.M.; supervision, A.G.-G.; project administration, A.G.-G.; funding acquisition, A.G.-G. All authors have read and agreed to the published version of the manuscript.

Funding

(1) This research was financed from the budgetary funds of the YOUNG PW II program—“Photo-crosslinked polyesters of itaconic acid and selected dihydroxyl diols enriched with magnetic microparticles for medical applications (MagnetItac)” (504/04496/1020/45.180003); (2) This research was funded by the Warsaw University of Technology within the Excellence Initiative: Research University (IDUB) program as research titled “Assessment of the Application Potential of Biodegradable Materials Containing Cysteine in Regenerative Medicine” (504/04496/1020/45.010045).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors would like to thank the Faculty of Chemistry, Warsaw University of Technology, for providing the laboratory equipment.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
%convCOOH titConversion of carboxyl groups (by titration)
%C=C IN titPercentage of unreacted C=C double bonds (by titration)
%C=C 1H NMRPercentage of unreacted C=C double bonds (by 1H NMR)
1,8-OD1,8-octanediol
ANOVAAnalysis of variance
ANtitAcid number (by titration)
BHTButylated hydroxytoluene
CACitraconic acid
DIDispersity index
DIWDirect ink writing
DLPDigital light processing
DMSO-d6Dimethyl sulfoxide
DSCDifferential scanning calorimetry
DTGDerivative thermogravimetry
EDtitEsterification degree (by titration)
EDNMREsterification degree (by NMR)
ENtitEster number (by titration)
FT-IRFourier transform infrared
G′Elastic modulus
G″Loss modulus
GPCGel permeation chromatography
IAItaconic acid
INtitIodine number (by titration)
LVELinear viscoelasticity
MAMesaconic acid
MEHQ4-metoxyphenol
MnNumber-average molecular weight
MwWeight-average molecular weight
NMRNuclear magnetic resonance
PBATpoly(butylene adipate-co-terephthalate)
PBItcPoly(tetramethylene itaconate)
PBSPolybutyrate succinate
PIDDOLpoly(ester thioether) based on 1,12-dodecanediyl bis(methyl itaconate) and linalool
POItcScPoly(octamethylene itaconate-co-succinate)
SASuccinic acid
SAnSuccinic anhydride
SECSize exclusion chromatography
t-BuOHTert-butanol
TGThermogravimetry
Tcc1Crystallization temperature
Tcch1Cold crystallization temperature during the first heating
Td5%5% decomposition temperature
Td30%30% decomposition temperature
Td50%50% decomposition temperature
Td85%85% decomposition temperature
Tgh1Glass transition temperature during the first heating
Tgh2Glass transition temperature during second heating
Tmh1Melting temperature during first heating
Tmh2Melting temperature during second heating
VVUViscosity-Visual-Utility

References

  1. Shou, Y.; Campbell, S.B.; Lam, A.; Lausch, A.J.; Santerre, J.P.; Radisic, M.; Davenport Huyer, L. Toward Renewable and Functional Biomedical Polymers with Tunable Degradation Rates Based on Itaconic Acid and 1,8-Octanediol-SI. ACS Appl. Polym. Mater. 2021, 3, 1943–1955. [Google Scholar] [CrossRef]
  2. Papadopoulos, L.; Pezzana, L.; Malitowski, N.M.; Sangermano, M.; Bikiaris, D.N.; Robert, T. UV-Curing Additive Manufacturing of Bio-Based Thermosets: Effect of Diluent Concentration on Printing and Material Properties of Itaconic Acid-Based Materials. ACS Omega 2023, 8, 31009–31020. [Google Scholar] [CrossRef]
  3. Karimi, A.; Rahmatabadi, D.; Baghani, M. Direct Pellet Three-Dimensional Printing of Polybutylene Adipate-co-Terephthalate for a Greener Future. Polymers 2024, 16, 267. [Google Scholar] [CrossRef]
  4. Sheldon, R.A. Green and Sustainable Manufacture of Chemicals from Biomass: State of the Art. Green Chem. 2014, 16, 950–963. [Google Scholar] [CrossRef]
  5. Renganathan, R.R.A.; Rai, V.R. Molecular Docking and Molecular Dynamics Evaluation of Aspergillus sp., Itaconic Acid Isolated from Garcinia Indica for Anticancer Potential. Biointerface Res. Appl. Chem. 2023, 13, 247. [Google Scholar] [CrossRef]
  6. Chongcharoenchaikul, T.; Thamyongkit, P.; Poompradub, S. Synthesis, Characterization and Properties of a Bio-Based Poly(Glycerol Azelate) Polyester. Mater. Chem. Phys. 2016, 177, 485–495. [Google Scholar] [CrossRef]
  7. Wei, B.T.; Lei, L.; Kang, H.; Qiao, B.; Wang, Z.; Zhang, L.; Coates, P.; Hua, K.; Kulig, J. Tough Bio-Based Elastomer Nanocomposites with High Performance for Engineering Applications. Adv. Eng. Mater. 2012, 14, 112–118. [Google Scholar] [CrossRef]
  8. Brännström, S.; Malmström, E.; Johansson, M. Biobased UV-Curable Coatings Based on Itaconic Acid. J. Coat. Technol. Res. 2017, 14, 851–861. [Google Scholar] [CrossRef]
  9. Gowsika, J.; Nanthini, R. Synthesis, Characterization and In Vitro Anticancer Evaluation of Itaconic Acid Based Random Copolyester. J. Chem. 2014, 2014, 173814. [Google Scholar] [CrossRef]
  10. Birajdar, M.S.; Joo, H.; Koh, W.G.; Park, H. Natural Bio-Based Monomers for Biomedical Applications: A Review. Biomater. Res. 2021, 25, 8. [Google Scholar] [CrossRef]
  11. Sindhu, M.; Mehta, S.; Kumar, S.; Saharan, B.S.; Malik, K.; Kayasth, M.; Nagar, S. Itaconic Acid: Microbial Production Using Organic Wastes as Cost-Effective Substrates. Microb. Org. Acids Prod. Util. Waste Feedstocks 2024, 9, 125–147. [Google Scholar] [CrossRef]
  12. Zuo, X.; Yu, R.; Li, R.; Xu, M.; Liu, C.; Hao, K.; Zhou, Y.; Huang, A.; Wu, C.; Cao, Z.; et al. Itaconic Acid/Cellulose-Based Hydrogels with Fire-Resistant and Anti-Freezing Properties via Vat Photopolymerization 3D Printing. Int. J. Biol. Macromol. 2024, 283, 137911. [Google Scholar] [CrossRef]
  13. Pezzana, L.; Melilli, G.; Sangermano, M.; Sbirrazzuoli, N.; Guigo, N. Sustainable Approach for Coating Production: Room Temperature Curing of Diglycidyl Furfuryl Amine and Itaconic Acid with UV-Induced Thiol-Ene Surface Post-Functionalization. React. Funct. Polym. 2023, 182, 105486. [Google Scholar] [CrossRef]
  14. Perković, I.; Beus, M.; Schols, D.; Persoons, L.; Zorc, B. Itaconic Acid Hybrids as Potential Anticancer Agents. Mol. Divers. 2022, 26, 1–14. [Google Scholar] [CrossRef]
  15. Kandil, H.; Ekram, B.; Abo-Zeid, M.A.M.; Abd El-Hady, B.M.; Amin, A. Hydroxyapatite/Hyperbranched Polyitaconic Acid/Chitosan Composite Scaffold for Bone Tissue Engineering. Polym. Compos. 2023, 44, 5633–5646. [Google Scholar] [CrossRef]
  16. Corici, L.; Pellis, A.; Ferrario, V.; Ebert, C.; Cantone, S.; Gardossi, L. Understanding Potentials and Restrictions of Solvent-Free Enzymatic Polycondensation of Itaconic Acid: An Experimental and Computational Analysis. Adv. Synth. Catal. 2015, 357, 1763–1774. [Google Scholar] [CrossRef]
  17. Brännström, S.; Finnveden, M.; Johansson, M.; Martinelle, M.; Malmström, E. Itaconate Based Polyesters: Selectivity and Performance of Esterification Catalysts. Eur. Polym. J. 2018, 103, 370–377. [Google Scholar] [CrossRef]
  18. Maturi, M.; Pulignani, C.; Locatelli, E.; Vetri Buratti, V.; Tortorella, S.; Sambri, L.; Comes Franchini, M. Phosphorescent Bio-Based Resin for Digital Light Processing (DLP) 3D-Printing. Green Chem. 2020, 22, 6212–6224. [Google Scholar] [CrossRef]
  19. Schoon, I.; Kluge, M.; Eschig, S.; Robert, T. Catalyst Influence on Undesired Side Reactions in the Polycondensation of Fully Bio-Based Polyester Itaconates. Polymers 2017, 9, 693. [Google Scholar] [CrossRef]
  20. Robert, T.; Eschig, S.; Biemans, T.; Scheifler, F. Bio-Based Polyester Itaconates as Binder Resins for UV-Curing Offset Printing Inks. J. Coat. Technol. Res. 2019, 16, 689–697. [Google Scholar] [CrossRef]
  21. Robert, T.; Friebel, S. Itaconic Acid-a Versatile Building Block for Renewable Polyesters with Enhanced Functionality. Green Chem. 2016, 18, 2922–2934. [Google Scholar] [CrossRef]
  22. Chanda, S.; Ramakrishnan, S. Poly(Alkylene Itaconate)s-An Interesting Class of Polyesters with Periodically Located Exo-Chain Double Bonds Susceptible to Michael Addition. Polym. Chem. 2015, 6, 2108–2114. [Google Scholar] [CrossRef]
  23. Chen, C.W.; Hsu, T.S.; Huang, K.W.; Rwei, S.P. Effect of 1,2,4,5-Benzenetetracarboxylic Acid on Unsaturated Poly(Butylene Adipate-Co-Butylene Itaconate) Copolyesters: Synthesis, Non-Isothermal Crystallization Kinetics, Thermal and Mechanical Properties. Polymers 2020, 12, 1160. [Google Scholar] [CrossRef]
  24. Yang, J.; Webb, A.R.; Ameer, G.A. Novel Citric Acid-Based Biodegradable Elastomers for Tissue Engineering. Adv. Mater. 2004, 16, 511–516. [Google Scholar] [CrossRef]
  25. Dey, J.; Xu, H.; Shen, J.; Thevenot, P.; Gondi, S.R.; Nguyen, K.T.; Sumerlin, B.S.; Tang, L.; Yang, J. Biomaterials Development of Biodegradable Crosslinked Urethane-Doped Polyester Elastomers. Biomaterials 2008, 29, 4637–4649. [Google Scholar] [CrossRef] [PubMed]
  26. Jeong, C.G.; Hollister, S.J. Mechanical, Permeability, and Degradation Properties of 3D Designed Poly(1,8 Octanediol-Co-Citrate) Scaffolds for Soft Tissue Engineering. J. Biomed. Mater. Res. Part B Appl. Biomater. 2010, 93, 141–149. [Google Scholar] [CrossRef] [PubMed]
  27. Wan, Y.; Feng, G.; Shen, F.H.; Balian, G.; Laurencin, C.T.; Li, X. Novel Biodegradable Poly(1,8-Octanediol Malate) for Annulus Fibrosus Regeneration. Macromol. Biosci. 2007, 7, 1217–1224. [Google Scholar] [CrossRef]
  28. Bettinger, C.J. Biodegradable Elastomers for Tissue Engineering and Cell-Biomaterial Interactions. Macromol. Biosci. 2011, 11, 467–482. [Google Scholar] [CrossRef] [PubMed]
  29. Yeshwant, K.; Ghaffari, R. A Biodegradable Wireless Blood-Flow Sensor. Nat. Biomed. Eng. 2019, 3, 7–8. [Google Scholar] [CrossRef]
  30. Chu, X.; Han, B.; Kuang, M.; Hou, S.; Zhao, H.; Xu, Q.; Zhang, X. Synthesis and Characterization of Biodegradable Poly(1,8-Octanediol-Co-Citrate)/Poly(ε-Caprolactone) Copolyester with Robust Resilience and Low Hysteresis. Macromol. Chem. Phys. 2025, 226, 2500038. [Google Scholar] [CrossRef]
  31. Prabakaran, R.; Marie, J.M.; Xavier, A.J.M. Biobased Unsaturated Polyesters Containing Castor Oil-Derived Ricinoleic Acid and Itaconic Acid: Synthesis, In Vitro Antibacterial, and Cytocompatibility Studies. ACS Appl. Bio Mater. 2020, 3, 5708–5721. [Google Scholar] [CrossRef]
  32. Giacobazzi, G.; Gioia, C.; Colonna, M.; Celli, A. Thia-Michael Reaction for a Thermostable Itaconic-Based Monomer and the Synthesis of Functionalized Biopolyesters. ACS Sustain. Chem. Eng. 2019, 7, 5553–5559. [Google Scholar] [CrossRef]
  33. Song, H.; Lee, S.Y. Production of Succinic Acid by Bacterial Fermentation. Enzym. Microb. Technol. 2006, 39, 352–361. [Google Scholar] [CrossRef]
  34. Tatara, A.M.; Watson, E.; Satish, T.; Scott, D.W.; Kontoyiannis, D.P.; Engel, P.S.; Mikos, A.G. Synthesis and Characterization of Diol-Based Unsaturated Polyesters: Poly(Diol Fumarate) and Poly(Diol Fumarate-Co-Succinate). Biomacromolecules 2017, 18, 1724–1735. [Google Scholar] [CrossRef]
  35. Floriańczyk, Z.; Penczek, S. Chemia Polimerów Tom II; Wydawnictwo Oficyna Wydawnicza Politechniki Warszawskiej: Warschau, Poland, 1997. [Google Scholar]
  36. Mahapatro, A.; Kumar, A.; Kalra, B.; Gross, R.A. Solvent-Free Adipic Acid/1,8-Octanediol Condensation Polymerizations Catalyzed by Candida Antartica Lipase B. Macromolecules 2004, 37, 35–40. [Google Scholar] [CrossRef]
  37. dos Santos, W.N.L.; Ferreira, S.L.C.; Bruns, R.E.; Ferreira, H.S.; Matos, G.D.; David, J.M.; Brandao, G.C.; da Silva, E.G.P.; Portugal, L.A.; dos Reis, P.S.; et al. Box-Behnken design: An alternative for the optimization of analytical methods. Anal. Chim. Acta 2007, 597, 179–186. [Google Scholar] [CrossRef] [PubMed]
  38. Miętus, M.; Cegłowski, M.; Gołofit, T.; Gadomska-Gajadhur, A. Enhanced Synthesis of Poly(1,4-Butanediol Itaconate) via Box–Behnken Design Optimization. Polymers 2024, 16, 2708. [Google Scholar] [CrossRef]
  39. Guarneri, A.; Cutifani, V.; Cespugli, M.; Pellis, A.; Vassallo, R.; Asaro, F.; Ebert, C.; Gardossi, L. Functionalization of Enzymatically Synthesized Rigid Poly(Itaconate)s via Post-Polymerization Aza-Michael Addition of Primary Amines. Adv. Synth. Catal. 2019, 361, 2559–2573. [Google Scholar] [CrossRef]
  40. Farmer, T.J.; Comerford, J.W.; Pellis, A.; Robert, T. Post-Polymerization Modification of Bio-Based Polymers: Maximizing the High Functionality of Polymers Derived from Biomass. Polym. Int. 2018, 67, 775–789. [Google Scholar] [CrossRef]
  41. Barrett, D.G.; Merkel, T.J.; Luft, J.C.; Yousaf, M.N. One-Step Syntheses of Photocurable Polyesters Based on a Renewable Resource. Macromolecules 2010, 43, 9660–9667. [Google Scholar] [CrossRef]
  42. Papadopoulos, L.; Malitowski, N.M.; Bikiaris, D. Bio-Based Additive Manufacturing Materials: An in-Depth Structure-Property Relationship Study of UV-Curing Polyesters from Itaconic Acid. Eur. Polym. J. 2023, 186, 111872. [Google Scholar] [CrossRef]
  43. Melilli, G.; Guigo, N.; Robert, T.; Sbirrazzuoli, N. Radical Oxidation of Itaconic Acid-Derived Unsaturated Polyesters under Thermal Curing Conditions. Macromolecules 2022, 55, 9011–9021. [Google Scholar] [CrossRef]
  44. Maturi, M.; Spanu, C.; Maccaferri, E.; Locatelli, E.; Benelli, T.; Mazzocchetti, L.; Sambri, L.; Giorgini, L.; Franchini, M.C. (Meth)Acrylate-Free Three-Dimensional Printing of Bio-Derived Photocurable Resins with Terpene- and Itaconic Acid-Derived Poly(Ester-Thioether)S. ACS Sustain. Chem. Eng. 2023, 11, 17285–17298. [Google Scholar] [CrossRef] [PubMed]
  45. Wu, D.; Leng, Y.M.; Fan, C.J.; Xu, Z.Y.; Li, L.; Shi, L.Y.; Yang, K.K.; Wang, Y.Z. 4D Printing of a Fully Biobased Shape Memory Copolyester via a UV-Assisted FDM Strategy. ACS Sustain. Chem. Eng. 2022, 10, 6304–6312. [Google Scholar] [CrossRef]
  46. Wang, G.; Hao, X.; Dong, Y.; Zhang, L. Bio-Based Poly(Butylene Succinate-Co-Dodecylene Succinate) Derived from 1,12-Dodecanediol: Synthesis and Characterization. J. Polym. Environ. 2023, 31, 4990–5002. [Google Scholar] [CrossRef]
  47. Papadopoulos, L.; Pezzana, L.; Malitowski, N.; Sangermano, M.; Bikiaris, D.N.; Robert, T. Influence of Reactive Diluent Composition on Properties and Bio-Based Content of Itaconic Acid-Based Additive Manufacturing Materials. Discov. Appl. Sci. 2024, 6, 290. [Google Scholar] [CrossRef]
  48. Mahdavi, R.; Zahedi, P.; Goodarzi, V. Application of Poly(Glycerol Itaconic Acid) (PGIt) and Poly(ε-Caprolactone) Diol (PCL-Diol) as Macro Crosslinkers Containing Cloisite Na+ to Application in Tissue Engineering. J. Polym. Environ. 2024, 32, 3392–3406. [Google Scholar] [CrossRef]
  49. Barrioni, B.R.; De Carvalho, S.M.; Oréfice, R.L.; De Oliveira, A.A.R.; Pereira, M.D.M. Synthesis and Characterization of Biodegradable Polyurethane Films Based on HDI with Hydrolyzable Crosslinked Bonds and a Homogeneous Structure for Biomedical Applications. Mater. Sci. Eng. C 2015, 52, 22–30. [Google Scholar] [CrossRef]
  50. Ghaffari-Bohlouli, P.; Golbaten-Mofrad, H.; Najmoddin, N.; Goodarzi, V.; Shavandi, A.; Chen, W.H. Reinforced Conductive Polyester Based on Itaconic Acids, Glycerol and Polypyrrole with Potential for Electroconductive Tissue Restoration. Synth. Met. 2023, 293, 117238. [Google Scholar] [CrossRef]
  51. Golbaten-Mofrad, H.; Seyfi Sahzabi, A.; Seyfikar, S.; Salehi, M.H.; Goodarzi, V.; Wurm, F.R.; Jafari, S.H. Facile Template Preparation of Novel Electroactive Scaffold Composed of Polypyrrole-Coated Poly(Glycerol-Sebacate-Urethane) for Tissue Engineering Applications. Eur. Polym. J. 2021, 159, 110749. [Google Scholar] [CrossRef]
  52. Mondal, S.; Martin, D. Hydrolytic Degradation of Segmented Polyurethane Copolymers for Biomedical Applications. Polym. Degrad. Stab. 2012, 97, 1553–1561. [Google Scholar] [CrossRef]
  53. Anjum, S.; Li, T.; Arya, D.K.; Ali, D.; Alarifi, S.; Yulin, W.; Hengtong, Z.; Rajinikanth, P.S.; Ao, Q. Biomimetic Electrospun Nanofibrous Scaffold for Tissue Engineering: Preparation, Optimization by Design of Experiments (DOE), in-Vitro and in-Vivo Characterization. Front. Bioeng. Biotechnol. 2023, 11, 1288539. [Google Scholar] [CrossRef] [PubMed]
  54. Mehdipour-Ataei, S.; Tabatabaei-Yazdi, Z. Heat resistant polymers. In Encyclopedia of Polymer Science and Technology; John Wiley and Sons: Hoboken, NJ, USA, 2015; ISBN 0471440264. [Google Scholar]
  55. Sahu, P.; Sharma, L.; Dawsey, T.; Gupta, R.K. Insight into the Synthesis and Thermomechanical Properties of “Short-Long” Type Biobased Aliphatic Polyesters. J. Appl. Polym. Sci. 2024, 141, e54972. [Google Scholar] [CrossRef]
  56. Pérocheau Arnaud, S.; Malitowski, N.M.; Meza Casamayor, K.; Robert, T. Itaconic Acid-Based Reactive Diluents for Renewable and Acrylate-Free UV-Curing Additive Manufacturing Materials. ACS Sustain. Chem. Eng. 2021, 9, 17142–17151. [Google Scholar] [CrossRef]
  57. Ouyang, L.; Highley, C.B.; Rodell, C.B.; Sun, W.; Burdick, J.A. 3D Printing of Shear-Thinning Hyaluronic Acid Hydrogels with Secondary Cross-Linking. ACS Biomater. Sci. Eng. 2016, 2, 1743–1751. [Google Scholar] [CrossRef]
Figure 1. Structural similarity of acrylic and itaconic compounds, where orange refers to the C=C double bond, and green and red to the carboxyl groups.
Figure 1. Structural similarity of acrylic and itaconic compounds, where orange refers to the C=C double bond, and green and red to the carboxyl groups.
Polymers 17 02220 g001
Figure 2. Undesirable side reactions involving itaconic compounds.
Figure 2. Undesirable side reactions involving itaconic compounds.
Polymers 17 02220 g002
Figure 3. Poly(octamethylene itaconate-co-succinate) synthesis.
Figure 3. Poly(octamethylene itaconate-co-succinate) synthesis.
Polymers 17 02220 g003
Figure 4. “Black box” of the poly(octamethylene itaconate-co-succinate) synthesis.
Figure 4. “Black box” of the poly(octamethylene itaconate-co-succinate) synthesis.
Polymers 17 02220 g004
Figure 5. Dependence of the (A) percentage conversion of carboxyl groups -COOH of the POItcSc product on the reaction’s temperature (x3) and time (x2), x1 = 1; (B) percentage of unreacted C=C double bonds of the POItcSc product on the reaction’s time (x2) and temperature (x3), x1 = 1; (C) Viscosity-Visual-Utility analysis of the POItcSc product on the reaction’s time (x2) and IA molar fraction (x1); x3 = 1. The blue circles correspond to the conditions of the performed reactions.
Figure 5. Dependence of the (A) percentage conversion of carboxyl groups -COOH of the POItcSc product on the reaction’s temperature (x3) and time (x2), x1 = 1; (B) percentage of unreacted C=C double bonds of the POItcSc product on the reaction’s time (x2) and temperature (x3), x1 = 1; (C) Viscosity-Visual-Utility analysis of the POItcSc product on the reaction’s time (x2) and IA molar fraction (x1); x3 = 1. The blue circles correspond to the conditions of the performed reactions.
Polymers 17 02220 g005
Figure 6. FTIR spectra of poly(octamethylene itaconate-co-succinate) (black line), itaconic acid (red line), 1,8-octanediol (pink line), and succinic anhydride (green line).
Figure 6. FTIR spectra of poly(octamethylene itaconate-co-succinate) (black line), itaconic acid (red line), 1,8-octanediol (pink line), and succinic anhydride (green line).
Polymers 17 02220 g006
Figure 7. 1H NMR spectra of poly(octamethylene itaconate-co-succinate).
Figure 7. 1H NMR spectra of poly(octamethylene itaconate-co-succinate).
Polymers 17 02220 g007
Figure 8. DSC for POItcSc.
Figure 8. DSC for POItcSc.
Polymers 17 02220 g008
Figure 9. Thermogravimetric degradation and derivative thermogravimetric curve for poly(1.8-octanediol itaconate-co-succinate).
Figure 9. Thermogravimetric degradation and derivative thermogravimetric curve for poly(1.8-octanediol itaconate-co-succinate).
Polymers 17 02220 g009
Figure 10. Continuous flow tests of POItcSc at a shear rate equivalent to those experienced during 3D printing at 25 °C and 36.6 °C (A), and temperature dependence of the storage modulus, loss modulus, and loss factor for POItcSc (B).
Figure 10. Continuous flow tests of POItcSc at a shear rate equivalent to those experienced during 3D printing at 25 °C and 36.6 °C (A), and temperature dependence of the storage modulus, loss modulus, and loss factor for POItcSc (B).
Polymers 17 02220 g010
Figure 11. Amplitude sweep of POItcSc at 25 °C (A) and 36.6 °C (B) and frequency sweep of POItcSc at 25 °C (C) and 36.6 °C (D).
Figure 11. Amplitude sweep of POItcSc at 25 °C (A) and 36.6 °C (B) and frequency sweep of POItcSc at 25 °C (C) and 36.6 °C (D).
Polymers 17 02220 g011
Table 1. Viscosity-Visual-Utility evaluation for the obtained POItcSc products.
Table 1. Viscosity-Visual-Utility evaluation for the obtained POItcSc products.
StructureConsistencyTransparencyAbility to Spread the Sample on the TableViscosity [Pa∙s]
1Hard and brittle1Wax1None1Yes1η < 10 or η > 1000
2Incompressible and sticky2Wax/Resin2Partial2Partial2500 < η < 1000
3Compressible and sticky3Resin3Full3No3100 < η < 500
410 < η < 100
Table 2. Experimental matrix and the calculated results for the output variables.
Table 2. Experimental matrix and the calculated results for the output variables.
No.Coded Variable%convCOOH tit [%]%C=C 1H NMR [%]%vvu [%]
x1x2x3Exp. 1Calc. 2Diff. 3Exp.Calc.Diff.Exp.Calc.Diff.
1−1−1060.263.5−3.370.973.3−2.478.178.5−0.4
21−1073.370.52.773.773.70.084.483.21.2
3−11081.083.7−2.770.069.90.071.973.0−1.2
411075.972.63.374.972.52.490.690.20.4
5−10−170.868.91.970.565.35.365.664.51.2
610−162.666.7−4.174.371.42.978.178.5−0.4
7−10182.778.74.180.082.9−2.968.868.40.4
810174.976.8−1.974.579.8−5.375.076.2−1.2
90−1−165.964.51.372.775.5−2.978.178.9−0.8
1001−176.075.20.859.765.0−5.378.178.10.0
110−1173.374.1−0.885.780.35.378.178.10.0
1201184.385.6−1.389.286.32.981.380.50.8
1300072.473.5−1.177.979.0−1.281.380.21.0
1400074.873.51.477.179.0−2.078.180.2−2.1
1500073.273.5−0.382.279.03.281.380.21.0
1 Exp.—Experimental; 2 Calc.—Calculated; 3 Diff.—Difference.
Table 3. Calculated and experimental results of the output variables obtained for the product under optimal conditions.
Table 3. Calculated and experimental results of the output variables obtained for the product under optimal conditions.
Result%convCOOH tit [%]%C=C 1H NMR [%]%VVU [%]
Calculated85.686.380.5
Experimental83.388.787.5
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Miętus, M.; Gołofit, T.; Gadomska-Gajadhur, A. Towards Greener Polymers: Poly(octamethylene itaconate-co-succinate) Synthesis Parameters. Polymers 2025, 17, 2220. https://doi.org/10.3390/polym17162220

AMA Style

Miętus M, Gołofit T, Gadomska-Gajadhur A. Towards Greener Polymers: Poly(octamethylene itaconate-co-succinate) Synthesis Parameters. Polymers. 2025; 17(16):2220. https://doi.org/10.3390/polym17162220

Chicago/Turabian Style

Miętus, Magdalena, Tomasz Gołofit, and Agnieszka Gadomska-Gajadhur. 2025. "Towards Greener Polymers: Poly(octamethylene itaconate-co-succinate) Synthesis Parameters" Polymers 17, no. 16: 2220. https://doi.org/10.3390/polym17162220

APA Style

Miętus, M., Gołofit, T., & Gadomska-Gajadhur, A. (2025). Towards Greener Polymers: Poly(octamethylene itaconate-co-succinate) Synthesis Parameters. Polymers, 17(16), 2220. https://doi.org/10.3390/polym17162220

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop