You are on page 1of 12

Research Article

Cite This: ACS Sustainable Chem. Eng. 2020, 8, 914−925 pubs.acs.org/journal/ascecg

Tailoring the Performance of Vegetable Oil-Based Waterborne


Polyurethanes through Incorporation of Rigid Cyclic Rings into Soft
Polymer Networks
Haiyan Liang,† Yanchun Li,† Siying Huang,† Kaixi Huang,† Xueyi Zeng,† Qianwen Dong,†
Chengguo Liu,‡ Pengju Feng,*,§ and Chaoqun Zhang*,†

College of Materials and Energy, South China Agricultural University, Guangzhou 510642, P. R. China

Institute of Chemical Industry of Forest Products, Chinese Academy of Forestry (CAF), National Engineering Laboratory for
Biomass Chemical Utilization, Key Laboratory on Forest Chemical Engineering, National Forestry and Grassland Administration,
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Key Laboratory of Biomass Energy and Material, Jiangsu Province, 16 Suojin Wucun, Nanjing 210042, P. R. China
§
Department of Chemistry, Jinan University, 601 Huang-Pu Avenue West, Guangzhou 510632, P. R. China
Downloaded via UNIV OF SAO PAULO on December 16, 2023 at 11:58:40 (UTC).

ABSTRACT: Vegetable oil-based polymeric materials always


suffer from relatively poor performance, such as lower tensile
strength and glass transition temperatures, than petroleum-
based polymeric materials, which greatly limit their practical
applications. In this study, octahydro-2,5-pentalenediol
(OPD) was synthesized from naturally occurring citric acid
and used together with castor oil as the polyol blends for the
production of bio-based waterborne polyurethane (PU)
dispersions (PUDs). The effects of the OPD contents on
the particle size and ζ potential of the resulting polyurethane
dispersion, and the thermal stability, mechanical properties,
and hydrophilicity of the resulting polyurethane films were
systematically investigated. Taking advantage of the rigid
cyclic structures of the OPD and the flexible fatty acid chain of castor oil, waterborne polyurethane (WPU) films with tailorable
mechanical performance ranging from elastomeric polymers to rigid plastics were successfully prepared and characterized. The
tensile strength of the samples increases from 9.5 to 22.3 MPa with the ratio between OPD and castor oil increasing from 0:10
to 5:5, whereas their elongation at break decreases from 192 to 12%. A significant increase in the glass transition temperature,
transparency, and anticorrosive properties was observed for the resulting polyurethane films with increasing the OPD content.
However, the thermal stability of the PU films exhibited a slight decrease as the OPD content increased. Moreover, the water
contact angle exhibited a slight increase for the polyurethane films prepared from polyol blends compared to the WPU film
prepared from pure castor oil. This work provides a novel route to tailor the performance of vegetable oil-based waterborne
polyurethanes through the incorporation of rigid cyclic rings into soft polymer networks.
KEYWORDS: polyols, waterborne polyurethane, octahydro-2,5-pentalenediol, castor oil

■ INTRODUCTION
For decades, polyurethanes (PUs) have been considered to be
Recently, the raw materials for the preparation of WPU,
including polyols, isocyanates, catalysts, additives, and so on,
one of the most versatile polymers and applied in various are still mainly derived from petroleum resources.5 With the
industrial fields, such as ink, adhesives, foams, sealants, and so increasing concern on limited crude oil reserves, research
on.1,2 However, during the production and application, studies on utilizing bio-based feedstock, including cellulose,
conventional solvent-borne PUs always used plenty of volatile lignin, and vegetable oils as substitutes for petroleum-based
organic compounds (VOCs) and hazardous air pollutants counterparts, are new promising domains.6 Among them,
(HAPs), which are serious environmental and health hazards.3 vegetable oils are one of the most promising options due to
Therefore, waterborne polyurethanes (WPU) have been their low cost, renewability, and ready availability.7 There are
considered as the replacement for traditional solvent-borne three esters and 0−7 carbon−carbon double bonds in the
PUs due to the intensive attention toward safety and molecular of vegetable oils, which provides the reactive sites
environmental issues on a global scale. WPUs exhibit many for chemical modification.8 Castor oil is the only natural
outstanding performances, including good applicability,
excellent abrasion resistance, low viscosity at high temper- Received: September 15, 2019
atures, etc., which undoubtedly expanded their applications in Revised: December 5, 2019
medical dressing, drug loading, and other biomedical fields.4 Published: December 11, 2019

© 2019 American Chemical Society 914 DOI: 10.1021/acssuschemeng.9b05477


ACS Sustainable Chem. Eng. 2020, 8, 914−925
ACS Sustainable Chemistry & Engineering Research Article

vegetable oil that contains hydroxyl groups that can be directly


employed as the polyol in the preparation of WPU without
■ EXPERIMENTAL SECTION
Materials. Glyoxal (40% aqueous solution) and dimethyl-1,3-
further modification, whereas other vegetable oils have to be acetonedicarboxylate were purchased from Sigma-Aldrich. Sodium
modified to introduce hydroxyl groups. However, the flexible hydroxide, sodium bicarbonate, and anhydrous MgSO4 were
long chain of fatty acids and the relatively low OH number purchased from Tianjin Damao Chemical Reagent Factory. NaBH4
(mostly 70−300 mg KOH/g)9 of vegetable oil-based polyols was provided by Energy Chemical Co. Ltd. Anhydrous methanol and
led to their resulting WPU films suffering from certain limits of aqueous hydrochloric acid were purchased from Tianjin Chemical
mechanical and thermal properties, such as relatively low Reagent Co. Ltd. Dichloromethane (DCM) was purchased from
Guangzhou Chemical Reagents Factory. Castor oil (OH number: 164
tensile strength and glass transition temperatures (Tg).10
mg KOH/g) was supplied by Fuyu Chemical Co. Ltd. Isophorone
Much effort has been dedicated to improve the tensile diisocyanate was provided by Wengjiang Chemical Reagent Co. Ltd.
strength and Tg of the vegetable oil-based WPU films, such as Dimethylolbutanoic acid was purchased from Adamas Reagent Co.,
nanoenhancement, fiber reinforcement, increasing cross-link- Ltd. Methyl ethyl ketone (MEK), glacial acetic acid, and triethylamine
ing density, and so on. Among them, incorporating rigid cyclic (TEA) were purchased from Aladdin reagent. Dibutyltin dilaurate
structures in the backbone of polymer chains stands out at the (DBTDL) was purchased from the Fuchen Chemical Reagent
privileged position for excellent mechanical strength and high Factory. All materials were used as received without further
Tg.5 For example, Wang et al. reported that employing purification.
isophorone diamine, containing cyclic structures, as the chain Synthesis of OPD. Under a N2 atmosphere, sodium hydroxide
(1.40 g, 35.0 mmol) was dissolved in anhydrous methanol (25 mL,
extender can significantly increase the water resistance and 1.46 M) at 0 °C to form a milky-white solution. Dimethyl-1,3-
tensile strength of the resulting WPU films, compared with the acetonedicarboxylate (6.0 g, 34.4 mmol) was added dropwise over a
films whose chain extended by diethylene triamine and period 15 min. A yellow precipitate was formed, and the resulting
ethylenediamine.11 Delpierre et al. have successfully tuned mixture was heated to reflux for 3 h at 65 °C under vigorous stirring.
the tensile strength of the PU films from 0.3 to 11 MPa via An aqueous solution of 40% glyoxal (1.14 g, 19.7 mmol) was added
varying the boroxine ring content from 25 to 75% molar, and dropwise to the stirred hot solution. After addition, the reaction was
the Tg of the films was improved from −11 to 41 °C.12 Xia et moved to room temperature and stirred overnight. The reaction
al. incorporated isosorbide into the WPU chains ranging from mixture was filtered and washed with methanol (5.0 mL) to obtain a
light yellow disodium salt that was directly dissolved in a mixture of
0 to 20% of the total diols, and thus obtained significant aqueous hydrochloric acid (1.0 M, 5.0 mL) and glacial acetic acid
improvement in mechanical properties (from 0.69 to 8.15 (10.0 mL). After refluxing for 5 h, the reaction mixture was cooled
MPa) and Tg. down to room temperature and extracted with DCM (3 × 15 mL).
In this work, octahydro-2,5-pentalenediol (OPD) was The combined organic phase was washed with aqueous sodium
synthesized from dimethyl-1,3-acetonedicarboxylate (derived bicarbonate (3 × 20 mL), dried over anhydrous MgSO4, filtered, and
from naturally occurring citric acid) and glyoxal through concentrated under vacuum to give tetrahydropentalene-2,5(1H,3H)-
Weiss−Cook condensation, decarboxylation under acidic dione as a brown solid (3.95 g, 28.6 mmol, 83%). 1H NMR (300
conditions and reduction with NaBH4. OPD, with two fused MHz, CDCl3) δ 2.93−2.90 (m, 2H), 2.46−2.37 (m, 4H), 2.21−1.95
(m, 4H); 13C NMR (75 MHz, CDCl3) 218.0 (CO), 43.4 (CH2),
cyclopentane rings, exhibits excellent rigidity and thermal 36.2 (CH).
stability, which envisioned to greatly restrict the motion of the A solution of tetrahydropentalene-2,5(1H,3H)-dione (4.26 g, 30.0
polymer chain if incorporated in the macromolecular mmol) in methanol (60.0 mL) was stirred under N2 at 0 °C. NaBH4
structures, leading to the high performance of polymers (2.27 g, 60.0 mmol) was added portion-wise to the above solution,
(high Tg and tensile strength).13−15 along with the generation of lots of bubbles. After stirring at room
Therefore, OPD was used as a partial replacement of castor temperature overnight, the reaction mixture was poured into ice and
oil as the polyol for the preparation of vegetable oil-based diluted with hydrochloric acid (80 mL, 2.0 mol/L), and extracted
WPUs. The OH group from OPD, castor oil, and emulsifier with DCM (3 × 150 mL). The combined organic phase was washed
with aqueous sodium bicarbonate (3 × 20 mL), dried over anhydrous
[dimethylolbutanoic acid (DMBA)] competitively reacts with
MgSO4, filtered, and concentrated under vacuum to give diaster-
the NCO of isophorone diisocyanate (IPDI) to form the eoisomers of ocatahydro-2,5-pentalenediol (syn/anti = 1/3) as a
polymer. To investigate the influence of the OPD content on white solid (3.80 g, 26.7 mmol, 89%). 1H NMR (300 MHz,
the resulting WPU, a series of polyurethane dispersions (CD3)2SO) δ (major isomer) 4.58 (d, J = 4.3 Hz, 2H), 4.01−3.92 (m,
(PUDs) are synthesized with different proportions between 2H), 2.27−2.15 (m, 2H), 1.98−1.85 (m, 4H), 1.44−1.33 (m, 4H); δ
OPD and castor oil. The structures of the intermediate (minor isomer) 4.50 (d, J = 4.3 Hz, 1H), 4.33 (d, J = 4.3 Hz, 1H),
[tetrahydropentalene-2,5(1H,3H)-dione] and OPD were con- 4.24−4.10 (m, 1H), 4.01−3.92 (m, 1H), 2.45−2.30 (m, 2H), 1.98−
firmed by 1H NMR, 13C NMR, and Fourier transform infrared 1.85 (m, 2H), 1.66−1.46 (m, 4H), 1.14−1.08 (m, 2H); 13C NMR (75
(FTIR) spectroscopy. The particle size and ζ potential of the MHz, CDCl3), 13C NMR (75 MHz, (CD3)2SO): δ (major isomer)
74.0 (CH), 42.4 (CH2), 38.3 (CH); δ (minor isomer) 73.4 (CH),
PUDs were characterized by a zetasizer. Dynamic mechanical 72.5 (CH), 42.31 (CH2), 42.26 (CH2), 38.2 (CH).
analysis (DMA), thermogravimetric analysis (TGA), and Preparation of WPUs from OPD and Castor Oil-Based
differential scanning calorimetry (DSC) were performed to Polyols. First, IPDI (3.89 g) and DBTDL (5 μL) were added into an
investigate the thermal properties of the resulting PU films, and oven-dried double-necked flask and stirred (240−260 rpm) at 78 °C
the mechanical properties were characterized by tensile testing. for 2 h. Then, the OPD/castor oil/DMBA mixture was added
Moreover, the effect of the OPD amount on the pencil dropwise to the flask for 30 min, while MEK was added to reduce the
hardness, crosshatch adhesion, and water contact angle of the viscosity. The mole ratio among the NCO groups of the IPDI, the
OH groups of the OPD and castor oil, the OH groups of the DMBA
PU films are also explored. Taking advantage of the rigid cyclic was 1.7:1.0:0.69. The mole ratio between OPD and castor oil in the
structures of the OPD and the flexible fatty acid chain of castor OPD/castor oil mixture was varied from 0:10 to 1:9, 2:8, 3:7, 4:6, 5:5,
oil, it was anticipated to obtain WPU films with tailorable and 6:4, which were coded as PU-OPD0, PU-OPD10, PU-OPD20,
mechanical performance ranging from elastomeric polymers to PU-OPD30, PU-OPD40, PU-OPD50, and PU-OPD60, respectively.
rigid plastics. After another 1 h at 78 °C, the reactants were cooled down to room

915 DOI: 10.1021/acssuschemeng.9b05477


ACS Sustainable Chem. Eng. 2020, 8, 914−925
ACS Sustainable Chemistry & Engineering Research Article

Scheme 1. Synthesis Routine of Octahydro-2,5-pentalenediol

Figure 1. (a) 1H NMR and (b) 13C NMR spectrum of tetrahydropentalene-2,5(1H,3H)-dione. (c) 1H NMR spectrum, (d) 13C NMR spectrum,
and (e) FTIR spectra of octahydro-2,5-pentalenediol.

temperature, and the mixture was neutralized with TEA at stirring for The particle size distribution and ζ potential of the PUDs were
30 min. Finally, deionized water (55−60 mL) was added for determined by a Zetasizer Nano ZSE (Malvern Instruments). The
emulsification, followed by vigorous stirring (350 rpm) for 2 h. After PUDs were diluted with distilled water to around 0.01 wt % before
the removal of MEK by rotary evaporator, the waterborne PUDs with conducting the tests.
a solid content of 15 wt % were obtained. The corresponding PU The transparency of the films was evaluated using an ultraviolet−
films were first obtained by casting 30 mL PUDs in a silicon mold (6 visible spectrophotometer (Evolution201) in the range of 200−800
× 6 cm2), drying at room temperature, and then the films were dried nm. The tensile testing was carried out on an electronic universal
in a vacuum at 80 °C for 24 h to remove the residual water. The testing machine (UTM-4202) with a crosshead speed of 100 mm/
representative synthesis routine of OPD/castor oil-based waterborne min. Rectangle sample films with a length of 25 mm and a width of 10
polyurethane dispersions is shown in Scheme 2. mm were dried at 60 °C for 12 h before use. At least three replicates
Characterization. A Bruker AVANCE III NMR spectrometer was of each resulting PU film were taken for the tests.
used to record the 1 H NMR and 13 C NMR spectra of The pencil hardness and crosshatch adhesion of the sample were
tetrahydropentalene-2,5(1H,3H)-dione and OPD separately operating measured according to ASTM D 3363 and ASTM D 3359 at room
at 300 and 100.6 MHz, using dimethyl sulfoxide (DMSO)-d6 as the temperature, respectively. The thickness of the coatings was measured
solvent at room temperature. using a Yuwen Ec-770 coating thickness gauge at an average of above
An FTIR spectrum for OPD powder was recorded using a Thermo- five measurements. The coatings used in the tests were obtained by
Nicolet Nexus 670 FTIR spectrometer over the wave numbers range casting 1 mL PUDs onto tinplate with a coated area of 5 × 10 cm2
of 400−4000 cm−1 using a KBr pelleting technique. The infrared and first drying at room temperature for 24 h, followed by drying in a
spectra of the dried WPU films were recorded on a VERTEX 70 vacuum at 60 °C for 24 h.
(Germany BRUKER) FTIR spectrometer ranging from 400 to 4000 Dynamic mechanical analysis (DMA) of the PU films was
cm−1. performed on a Netzsch DMA 242C dynamic mechanical analyzer

916 DOI: 10.1021/acssuschemeng.9b05477


ACS Sustainable Chem. Eng. 2020, 8, 914−925
ACS Sustainable Chemistry & Engineering Research Article

Scheme 2. Synthesis of OPD/Castor Oil-Based Waterborne Polyurethane Dispersions

in the tensile mode at 1 Hz, using rectangular specimens of 6 mm × The water contact angle of the resulting PU films was measured on
25 mm (length × width). The specimens were heated from −60 to a contact angle goniometer (Powereach JC2000C1) with the sessile-
120 °C at a rate of 5 °C min−1. The glass transition temperatures (Tg) drop method at room temperature. All of the samples were previously
were determined by the peaks of the loss factor curves. dried at 60 °C for over 48 h. The scanning electron microscopy
The thermogravimetric analysis of the PU samples was conducted (SEM) imaging of the PU films was performed on a JEOL JSM-
on a Netzsch-STA 449C thermal analyzer. All of the samples were 6380LA Quanta 200 microscope.
weighed between 8 and 14 mg and heated from 30 to 700 °C in a Tafel analysis of the samples was performed on an electrochemical
nitrogen atmosphere with a heating rate of 10 °C/min. working station (Shanghai Chenhua CHI 600E, China) at room
Differential scanning calorimetry (DSC) was carried out on a temperature using 3.5 wt % NaCl aqueous solution as the electrolyte.
thermal analyzer (PE DSC8000). All of the samples were weighed Potentiodynamic polarization curves were obtained using a three-
between 8 and 14 mg. They were first heated from 25 to 100 °C to electrode system [saturated Ag/AgCl electrode served as the reference
erase thermal history, then cooled to −70 °C at a rate of 10 °C/min, electrode, platinum (Pt) electrode as the counter electrode, and
and later heated to 150 °C at a rate of 5 °C/min. The midpoint uncoated or coated tin as the working electrode, respectively] at a
temperature in heat capacity change of the second DSC scan is scan rate of 10 mV/s. The samples used in the test possess a coated or
considered to be Tg. uncoated area of 1 × 1 cm2 with a coating thickness of 60 ± 5 μm.

917 DOI: 10.1021/acssuschemeng.9b05477


ACS Sustainable Chem. Eng. 2020, 8, 914−925
ACS Sustainable Chemistry & Engineering Research Article

■ RESULTS AND DISCUSSION


Synthesis and Characterization of Octahydro-2,5-
from 2827 to 2986 cm−1 gradually decrease as the OPD ratio
increased, resulting from the decreasing content of castor oil.
Furthermore, the small peak at 1741 cm−1 corresponding to
pentalenediol. As shown in Scheme 1, OPD was prepared
free urethane gradually decreases with the increase of the OPD
from the commercially available dimethyl-1,3-acetonedicarbox-
ratio, indicating the increasing hydrogen bond strength of the
ylate and glyoxal via Weiss−Cook condensation, decarbox-
films.18,19
ylation, and reduction reactions. The molecular structure of the
The PUDs prepared with different OPD contents and their
intermediates [tetrahydropentalene-2,5(1H,3H)-dione] and
particle size distribution are presented in Figure 2. With the
the desired product (OPD) were confirmed by 1H NMR,
13
C NMR, and FTIR, and the results are shown in Figure 1. As
seen in Figure 1a, only three signals were observed in the 1H
NMR spectrum of tetrahydropentalene-2,5(1H,3H)-dione, and
the peaks at δ 2.9, δ 2.4, and δ 2.0 ppm can be attributed to the
protons 1, 2, and 2′, respectively, indicating that the
intermediates have been successfully isolated and purified.
Furthermore, all of the characteristic peaks in the 13C NMR
spectrum of intermediates (Figure 1b) can be correctly
assigned to the various carbon atoms present in the
tetrahydropentalene-2,5(1H,3H)-dione chains, further con-
firming the structure of the intermediates. As shown in Figure
1c,d, the resonances assigned to the tetrahydropentalene-
2,5(1H,3H)-dione cannot be found after reduction reactions,
demonstrating that no intermediates remained in the desired Figure 2. (a) Appearance of PUDs with different OPD ratios in the
product (OPD). Moreover, the desired product is a mixture of polyol blends: (1) 0%, (2) 10%, (3) 20%, (4) 30%, (5) 40%, (6) 50%,
two stereoisomers, endo/endo isomer (identified as A), and and (7) 60%. (b) Particle size distribution of PUDs with different
OPD contents.
endo/exo isomer (identified as A′). The signals at around δ
4.5, δ 4.33, and δ 3.96 ppm can be assigned to the protons,
increase of the OPD contents, the transparency of the resulting
numbered as 3, 7, and 10, on carbons that were linked to
PUDs gradually increases, from milky of PU-OPD0 to visible
hydroxyl groups in isomer A′ and A, respectively, while the
transparency with the yellow light of PU-OPD60, which is
protons (numbered as 1 and 5) on bridged carbons exhibited associated with their particle sizes. As shown in Figure 2 and
overlapped multiplets at around 2.39 and 2.21 ppm, Table 1, the average particle size of the samples gradually
respectively.13 The integral area ratio of proton peaks 3 to 7,
3 to 10, and 1 to 5 are 6.1:1, 5.6:1, and 3.2:1, respectively,
Table 1. Properties of the Resulting PUDs
implying that the molar ratio of A to A′ is around 3:1.14 In the
FTIR spectra of OPD, an identical strong absorption band at hydrophilic Z-
3295 cm−1 can be observed, corresponding to the typical −OH DMBA groups average ζ
contents contents size potential
stretching vibration on the bicycles. In addition, the bands at samples (wt %) (wt %) appearance (nm) (mV)
1109 and 2948 cm−1 were assigned to the stretching of −C−O PU- 8.2 2.5 milky white 615 −50.3
and −CH2, respectively, which further confirmed the structure OPD0
of OPD. PU- 8.6 2.6 milky white 342 −58.4
Preparation and Properties of the PUDs with Differ- OPD10
ent OPD/Castor Oil Ratios. OPD was used together with PU- 9.0 2.7 milky white 255 −60.5
OPD20
castor oil as the polyol blends for the production of bio-based PU- 9.5 2.9 milky white 220 −60.8
waterborne polyurethane dispersions (PUDs). Taking advant- OPD30
age of the rigid cyclic structures of the OPD and the flexible PU- 10.0 3.0 translucent 43.8 −59.8
fatty acid chain of castor oil, WPU with tailorable performance OPD40
should be prepared. The effect of the OPD contents on the PU- 10.6 3.2 transparent 24.4 −45.2
OPD50 with yellow
particle size and ζ potential of the resulting polyurethane light
dispersion, and the thermal stability, mechanical properties, PU- 11.2 3.4 transparent 22.8 −34.8
and hydrophilicity of the resulting polyurethane films were OPD60 with yellow
systematically investigated. light
As shown in Scheme 2, the OH group from OPD, castor oil,
and emulsifier (dimethylolbutanoic acid) competitively reacts decreases as the OPD content increases. The average particle
with the NCO of isophorone diisocyanate (IPDI) to form sizes of PU-OPD0 (OPD ratio of the total diol is 0%) and PU-
carbamate groups. The molecular structures of the WPU films OPD60 (OPD ratio of the total diol is 60%) dispersions are
are confirmed by FTIR spectral analysis, and the FTIR spectra around 615 and 23 nm, respectively. Generally, two factors
are normalized with respect to the CO peak (1700 cm−1). mainly dominate the particle size and its distribution of the
As shown in Figure 3c, the characteristic peaks assigned to the resulting PUDs: hydroxyl numbers of the polyols and the
N−H (3322 cm−1), CH2 (2942 and 2856 cm−1), and CO hydrophilic contents (carboxylate ions).20,21
(1700 cm−1) stretching vibrations are both observed and the It is well known that a higher OH number of the polyol
peak assigned to the OH group (3295 cm−1) of OPD results in a higher cross-linking density of the resulting WPU,
disappeared for the PU films, which suggests the generation of leading to an increasing particle size of the resulting PUD.9,18
WPU.16,17 The peaks attributed to CH2 stretching vibrations Therefore, with the increase of the OPD content, the particle
918 DOI: 10.1021/acssuschemeng.9b05477
ACS Sustainable Chem. Eng. 2020, 8, 914−925
ACS Sustainable Chemistry & Engineering Research Article

Figure 3. (a) Appearance, (b) UV spectrum, (c) FTIR spectra, and (d) stress−strain curves for the PU films.

Table 2. Mechanical Properties of the Resulting PU Films


samples hard segment Content (wt %) tensile strength (MPa) elongation at break (%) toughness (J/mm3) Young’s modulus (MPa)
PU-OPD0 44.4 9.5 ± 2.7 191.7 ± 37.6 15.8 ± 3.1 32.1 ± 6.8
PU-OPD10 47.7 12.7 ± 1.5 168.8 ± 23.1 16.7 ± 7.0 119.9 ± 27.8
PU-OPD20 51.2 16.3 ± 2.2 140.9 ± 92.5 31.3 ± 12.7 172.5 ± 76.6
PU-OPD30 55.2 18.8 ± 1.8 28.6 ± 6.3 41.9 ± 18.5 183.7 ± 12.1
PU-OPD40 59.5 19.6 ± 1.0 14.6 ± 3.0 3.5 ± 0.5 192.2 ± 70.4
PU-OPD50 64.4 22.3 ± 1.1 12.3 ± 1.2 7.8 ± 0.4 382.8 ± 57.7

size of the PUDs theoretically increases, as a result of the precipitate or stratification were observed, further confirming
partial replacement of castor oil with high OH number OPD. their excellent storage stability.24,25
However, the use of OPD also leads to an increase of the Properties of the PU Films. As shown in Figure 3a,b, all
DMBA content in the backbone of polymer chains when the of the PU films with ∼0.8 mm thickness exhibited good
hydroxyl molar ratios between the OPD/castor oil and DMBA transparency. Notably, the incorporation of OPD increases the
remained constant (see Table 1). The carboxylate ions transparency of the samples. In detail, with the OPD ratio
provided by DMBA could keep the PUDs electrostatically increasing from 0 to 50%, the light transmittance value of the
stabilized via developing a stable electrical double layer around samples increases from 71 to 80%. This may mainly relate to
the PUD particles. 22 The increasing DMBA content the light absorption of castor oil,26,27 the ratio of which
contributes to the increase in the electric charges, which lead gradually decreases with the increase of the OPD ratio. As is
to the enhancement of the thickness of the electrical double known to all, the transparency of films is also influenced by the
thickness, cross-linking density, film-forming conditions, etc.,
layer and the electrostatic repulsive forces among the PUD
of the samples.28,29
particles. The aggregation of PUD particles is therefore
Figure 3d illustrates the tensile stress−strain curves for all
hindered, which results in the decreasing particle size and PU films, and tensile strength, elongation at break, toughness,
increasing transparency of the PUDs. With the smallest particle and the Young’s modulus are summarized in Table 2. The hard
size and the most transparent appearance among the samples, segment (HS) content is calculated based on the fraction of
PU-OPD60 possesses the highest DMBA content of 11.2 wt % the total mass of OPD, IPDI, DMBA, and TEA over the total
and the highest hydrophilic group content of 3.4 wt %. mass of castor oil, OPD, IPDI, DMBA, and TEA9
However, the particle size of the PUDs is also influenced by
other factors such as the ionic group position, the viscosity of hard segment content
the PUDs, process variables, etc.22,23
The absolute value of ζ potential of the PUDs shows a = mass(OPD + IPDI + DMBA + TEA)
similar trend with the particle size of the PUDs, and all of the /mass(castor oil + OPD + IPDI + DMBA + TEA)
samples exhibit the absolute value of ζ potential above 35 mV,
× 100% (1)
indicating the excellent stability of the samples (see Table 1).
For example, the ζ potential of PU-OPD0 is −50.3 mV, while The resulting PU film with 60% OPD content was found
that of PU-OPD50 is −45.2 mV. Moreover, all of the PUDs broken and could not be tensile tested due to its brittle nature
were centrifuged at 3000 rpm for 30 min, and no obvious when a high OPD content is incorporated into the backbone of
919 DOI: 10.1021/acssuschemeng.9b05477
ACS Sustainable Chem. Eng. 2020, 8, 914−925
ACS Sustainable Chemistry & Engineering Research Article

Table 3. DMA Data of the Resulting PU Films


Tg obtained from Tg obtained from E′ at 25 °C ve pencil hardness crosshatch adhesion film thickness
samples DSC (°C) DMA (°C) (MPa) (mol/m3) (6B-HB-6H, 6H = best) (5B = best) (μm)
PU- −7.4 35.2 78.3 208.1 2H 2B 32 ± 7
OPD0
PU- 0.05 48.1 315.7 340.3 2H 3B 30 ± 11
OPD10
PU- 4.8 48.0 374.8 304.7 2H 3B 29 ± 3
OPD20
PU- 6.5 68.0 939.4 232.0 5H 4B 27 ± 9
OPD30
PU- 15.4 71.4 1137.8 122.7 6H 4B 30 ± 8
OPD40
PU- 48.8 75.00 1619.7 123.8 6H 4B 28 ± 4
OPD50
PU- 52.6
OPD60

Figure 4. (a) DSC thermograms of the resulting PU films. (b) Dependence of Tg for the PU films. (c) DMA curves for the PU films with different
OPD contents.

the polymer network. Replacement of castor oil by OPD The coating properties of the resulting PU films, including
improves the mechanical properties of the PU films pencil hardness, crosshatch adhesion, and film thickness, were
significantly. In detail, as the OPD amount of total polyol also studied (see Table 3). The results show that PU films
increases from 0 to 50%, the Young’s modulus and tensile obtained from higher OPD contents exhibit improved pencil
strength of the resulting PU films increase from 32.1 to 382.8 hardness and crosshatch adhesion. The pencil hardness of the
and 9.5 to 22.3 MPa, respectively, while the elongation at break samples ranges from 2H to 6H. The lowest pencil hardness
of them decreases significantly from 192 to 12%, as expected. was observed in PU-OPD0, and the greatest pencil hardness
Toughness is considered as the characteristics of a material’s was observed in PU-OPD50. All samples show excellent
adhesion, which increases from 2B to 4B as the OPD content
resistance to fracture, which is obtained from the area below
increases. The increase of the OPD content results in a higher
the stress−strain curves. The toughness of the films increases
hard segment content of the samples and thus leads to the
from 15.8 to 41.9 J/mm3 with the increase of the OPD amount improvement in hardness and adhesion of the coatings.31,32
from 0 to 30% of total diol but sharply drop to around 8 J/ The thermal behaviors of the PU films were determined by
mm3 as the OPD amount further shifts above 40%. The DSC, and their data are shown in Table 3. As shown in Figure
increasing hard segment content of the PU films (see Table 2) 4a,b, only one Tg and no melting or crystallization transition
induced by the replacement of the castor oil by OPD and the are observed in all DSC curves, indicating the amorphous
increasing rigid bicyclic structure from the OPD lead to the nature of the OPD-based PU films. As the content of OPD
films with tailorable mechanical performance ranging from increases from 0 to 60%, the Tg of the PU films increases
ductile plastics to rigid plastics.30 steadily from −7.4 to 52.6 °C. On the one hand, the castor oil
920 DOI: 10.1021/acssuschemeng.9b05477
ACS Sustainable Chem. Eng. 2020, 8, 914−925
ACS Sustainable Chemistry & Engineering Research Article

Figure 5. (a) Network of the PU films from castor oil. (b) Network of the PU films from castor oil and OPD. (c) Chemical structure of monomers
for building the WPU network.

with high flexible long chains is gradually replaced by OPD, films reduces steadily upon further increase in the OPD molar
resulting in the increase of the hard segment (HS) content of fraction. This may result from the compensation of the
the PUs from 44 to 64 wt %. On the other hand, incorporation increased linear structure of OPD for the decreased distance
of OPD in the PU chains would decrease the free volume and between the cross-linking points. However, PU-OPD30 also
prevent close packing of polymer chains, leading to the shows a higher ve of 232 mol/m3 than PU-OPD0. Higher ve of
enhancement in Tg of the samples.13 the films will lower their chain mobility.37 Due to the fact that
Figure 4c depicts the storage modulus (E′) and tan δ curves higher temperature is required for rigid materials to achieve the
of the PU films prepared with different OPD contents. The E′ same degree of mobility, the higher the ve is, the lower the
of the resulting PU films over the whole temperature range chain mobility is, and Tg will be higher accordingly.
increases as the content of OPD increases. All of the PU films The Tgs, determined by both DMA curves and DSC curves,
are glassy below 0 °C, and the E′ decreases slightly as the show an upward trend (Figure 4b) as the OPD content
temperature increases. Then, a noticeable decline is observed increases. However, the Tgs observed from the DMA curves
in the E′ for all of the samples from 0 to 80 °C, originating are higher than the corresponding ones from DSC curves,
from the primary relaxation process of the films. The energy which could be explained by the different natures of the two
dissipation during the relaxation process is shown in the tan δ methods.20 The DSC method is based on the heat capacity
curves, where the maximum of the curve was taken as Tg. Only change when the samples are heated from frozen to unfrozen,
one Tg could be observed in all of the tan δ curves, which is which could be highly influenced by the cooling rate, whereas
consistent with the DSC results, corresponding to the the DMA method determines Tg by measuring the mechanical
homogenous nature of the PU films. The E′ at 25 °C, Tg, response change of the polymer chains.
and ve of the PU films are summarized in Table 3. In this work, TGA curves for PU films prepared from different OPD
the cross-linking density (ve) is calculated from the following contents are shown in Figure 6, and the corresponding data are
equation33 summarized in Table 4. Three stages of thermal degradation
E′ = 3veRT could be observed for all of the CO/OPD films. The
(2)
dissociation of labile urethane bonds is responsible for the
R is the universal gas constant, and T = Tg + 30 °C. An weight loss in the first stage between 200 and 320 °C, which
increase of Tg from 35 to 75 °C and a significant enhancement leads to the formation of isocyanates, alcohol, primary amines,
of E′ at 25 °C from 78.3 to 1619.7 MPa are observed for the secondary amines, and carbon dioxide. The temperature of 5
films when the OPD content increases from 0 to 50%. This can wt % weight loss (T5) is usually considered as the onset
be explained by the increase of hard segment contents in the decomposition temperature. As described in Table 4, there is a
samples induced by the replacement of castor oil by OPD with great improvement in T5 when OPD was introduced in the
the rigid bicyclic skeleton.14,34 Additionally, the increase of the backbone of the polymer network. Specifically, the T5 of PU
rigid bicyclic diol (OPD) reduces the free volume in the films without OPD is 209.9 °C, and the T5 of PU-OPD10 is
samples as mentioned above (shown in Figure 5), leading to 227.7 °C. This could be explained by the increased cross-
the impediment of molecular chain movements, then linking density of the sample from 208 to 340 mol/m3 when
improving the stiffness of the films and resulting in increased the OPD contents in the polyol blends increase from 0 to 10%,
E′ and Tg.35,36 Moreover, the ve also plays a vital role in which improve the thermal stability of the PU films.38
influencing the thermodynamic properties of the films. The ve However, a downward trend is observed when the OPD
of the PU film without OPD is 208.1 mol/m3, and it reaches contents continue to increase above 10% because the influence
340.3 mol/m3 when 10% of OPD was incorporated. This of the increasing urethane bonds takes the dominant position.
increase could be a result of the short molecular chains of As the content of OPD further increases to 50%, the T5 of the
OPD, which leads to a rather short distance between the cross- resulting films gradually decreases to 206.6 °C. The
linking points in the PU chains. However, the ve of the PU introduction of OPD leads to an increase of the hard segment
921 DOI: 10.1021/acssuschemeng.9b05477
ACS Sustainable Chem. Eng. 2020, 8, 914−925
ACS Sustainable Chemistry & Engineering Research Article

shows a declining trend as the OPD contents increase, which


manifests that the thermal stabilities decrease with the
replacement of castor oil by OPD. Generally, the polymers
with a lower molecular weight show poorer thermal stability.39
OPD has a much lower molecular weight than castor oil does,
and the increase of the OPD content would decrease the total
percentage of polyols by weight, thus leading to the decreased
thermal stabilities of the resulting PU films. The decreased
thermal stability, which results from the increasing urethane
content, compensates for the enhanced thermal stability from
the increased OPD content, which leads to the general
reduction of Tmax2 (the maximum degradation temperatures of
the OPD/castor oil chain). Additionally, all of the samples
exhibit similar Tmax1, which represents the maximum
degradation of the urethane/urea groups, whereas a slight
decrease is observed for the Tmax2 of the films with a higher
OPD content, which is attributed to the maximum degradation
of the OPD/castor oil fatty acid chains, originating from the
more thermally unstable OPD than the castor oil. The third
stage is mainly assigned to further thermo-oxidation of the
resulting PU films.
Figure 7a shows the water contact angle curves of PU films
with different OPD contents. It is found that the incorporation
of rigid OPD improves the water contact angles of the PU
Figure 6. (a) TGA curves and (b) their derivative curves of PU films
prepared from different OPD contents. films, indicating the enhancement of surface hydrophobicity.
There are two factors influencing the water contact angles of
Table 4. TGA and Water Contact Angle Data for the the films: the hydrophilic DMBA content and cross-linking
Resulting PU Films density. An obvious upward trend could be noted for the PU
films when the OPD content increases from 0 to 10%. In
water contact detail, the water contact angle of the film without OPD is 66°,
samples T5 T10 T50 Tmax1 Tmax2 angle (deg)
whereas the water contact angle of the film increases to 83°
PU- 209.9 253.7 330.4 305.2 343.4 66.3 ± 1.8 when 10% OPD was incorporated into the backbone of
OPD0
polyurethane chains. As mentioned above, with the introduc-
PU- 227.7 254.8 319.4 303.8 337.1 82.8 ± 3.3
OPD10 tion of 10% OPD, the cross-linking density increases drastically
PU- 218.7 246.6 318 301 340.2 74.5 ± 4.1 and therefore accompanies with a sharp increase of hydro-
OPD20 phobicity. Although the hydrophilic group (provided by
PU- 214 241.7 311.6 299.4 337.1 75.8 ± 1.2 DMBA) also increases with the increasing OPD content, the
OPD30 increase of hydrophobicity dominates the variation of the
70.3 ± 1.2
PU-
OPD40
219.2 250.9 306 301.2 326
water contact angle of the films.40,41 However, the water
PU- 206.6 242.5 304.3 301.8 326.2 72.0 ± 2.2 contact angles show a slight decline as the OPD content
OPD50 continues to increase, but they are still higher than PU-OPD0.
The water contact angle of the PU film with 50% OPD is 72°.
as discussed earlier (see Table 2), resulting in the enhance- This decrease in hydrophobicity can be explained by the
ment of labile urethane bonds in the PUDs, and therefore combination of the decreased cross-linking density and the
exhibits a decrease in thermal stability of the PU films. increased hydrophilic DMBA content of the samples with
The second stage in the range of 320−380 °C is mainly increasing OPD content.17 Besides, the surface roughness of
attributed to the chain scission of the polyols, namely, the the films is also an important factor in influencing the surface
OPD/castor oil fatty acid chains. In this degradation stage, T50 hydrophobicity of the samples. As shown in Figure 7b, the

Figure 7. (a) Water contact angle curves and (b) SEM images for the PU films prepared with different OPD contents.

922 DOI: 10.1021/acssuschemeng.9b05477


ACS Sustainable Chem. Eng. 2020, 8, 914−925
ACS Sustainable Chemistry & Engineering Research Article

Table 5. Electrochemical Corrosion Parameters for the PU Films


electrodes bare tin PU-OPD0 PU-OPD10 PU-OPD20 PU-OPD30 PU-OPD40 PU-OPD50
Ecorr (V) −1.014 −0.891 −0.869 −0.850 −0.801 −0.813 −0.778
Icorr (A) 3.62 × 10−4 4.76 × 10−5 3.89 × 10−5 2.427 × 10−5 2.21 × 10−5 2.071 × 10−5 1.92 × 10−5
IE (%) 86.9 89.3 93.3 93.9 94.3 94.7

surfaces of the samples are evenly distributed without an complex polymer network than the pure WPU network, which
increase in the large surface or recessed areas and gradually restricts the penetration of corrosive species more.48


become neat and smooth with the increase of the OPD ratio,
resulting from the relative increase in the cross-linking CONCLUSIONS
densities.42−44 This may be another reason for the slight
decrease in the water contact angle of the sample as the OPD In this work, OPD, with two parallel rigid cyclopentane rings,
ratio increased from 10 to 50%. was applied as the partial replacement of castor oil to prepare
The effect of the OPD ratio on the corrosion protection of bio-based waterborne polyurethane for enhanced mechanical
the samples is evaluated by the potentiodynamic polarization properties and Tg. The obtained PUDs exhibit improving
method.45 The electrochemical corrosion parameters for the transparency as the OPD content increases, and both the
PU films, including the corrosion potential (Ecorr), corrosion particle sizes and ζ potential of the PUDs decrease with the
current (Icorr), and inhibition efficiency (IE), are summarized increase of OPD. Experimental results demonstrate a
in Table 5. The Icorr is obtained by extrapolating the cathodic significant improvement in the mechanical properties of the
and anodic polarization curves of Tafel polarization for the PU films via the partial replacement of castor oil by OPD. An
samples, and the y coordinate of their intersection point is the increase of the OPD content from 0 to 50% leads to an
Icorr value. IE is calculated by the following equation46 increase of the hard segment in the PUDs, resulting in an
increase of Tg from 35 to 75 °C, Young’s moduli from 32 to
Icorr(0) − Icorr(i) 383 MPa, tensile strength from 9.5 to 22.3 MPa, and the water
IE(i)(%) = × 100%
Icorr(0) (3) contact angle from 66 to 72° of the samples. Moreover,
obvious upward trends in pencil hardness, crosshatch adhesion,
where Icorr(0) is the Icorr of bare tin and Icorr(i) is the Icorr of and anticorrosive properties are also observed for their films as
sample i. It is well known that a lower Icorr or a higher Ecorr the OPD content increases, whereas their thermal stability
corresponds to the improvement in the anticorrosive proper- decreases slightly. This work provides a new way for the
ties for coating.47 As shown in Figure 8 and Table 5, the Ecorr synthesis of vegetable oil-based waterborne polyurethane with
enhanced performance, which could find wide applications in
ink, adhesives, sealants, and anticorrosion coatings. However,
further optimizations are still needed, such as improving the
toughness while retaining the high tensile strength of the films.

■ AUTHOR INFORMATION
Corresponding Authors
*E-mail: pfeng@jnu.edu.cn (P.F.).
*E-mail: zhangcq@scau.edu.cn, nwpuzcq@gmail.com (C.Z.).
ORCID
Chengguo Liu: 0000-0001-5768-5184
Pengju Feng: 0000-0002-5470-0403
Chaoqun Zhang: 0000-0001-5754-8729
Figure 8. Tafel polarization curves of the PU films prepared with
Notes
different OPD contents after immersion in 3.5 wt % aqueous NaCl
solution for 2 h. The authors declare no competing financial interest.

of the coating obviously increases from −0.891 to −0.778 V


with an increase of the OPD ratio from 0 to 50% and both
■ ACKNOWLEDGMENTS
This work was sponsored by the Fundamental Research Funds
have more positive values than that of bare tin. Besides, the Icorr of CAF (CAFYBB2017QB006), the National Natural Science
of the samples decreases from 4.76 × 10−5 to 1.92 × 10−5 A as Foundation (51703068 and 21602078), the Guangdong
the OPD ratio increased from 0 to 50%, while the bare tin Province Science & Technology Program (2017A010103015
exhibits the largest Icorr of 3.62 × 10−4 A. Both these results and 2018B030306016), the Program of the Science,
demonstrate that the WPU coating provides effective Technology Department of Guangzhou, China
anticorrosion protection for the bare tin with an IE above (201803020039), Key Projects of Basic Research and Applied
86.9%, resulting from the barrier effect of coating.45 Moreover, Basic Research of the Higher Education Institutions of
the corrosion protection performance of the coating increases Guangdong Province (2018KZDXM014), Guangzhou Munic-
with the increase of the OPD content. In detail, the IE of the ipal Key Laboratory of Woody Biomass Functional New
samples increases from 86.9 to 94.7% as the OPD ratio Materials (201905010005), and the Open Project of Jiangsu
increased from 0 to 50%. This can be explained by the fact that Key Laboratory for Biomass Energy and Materials
the incorporation of OPD results in the formation of a more (JSBEM201803).
923 DOI: 10.1021/acssuschemeng.9b05477
ACS Sustainable Chem. Eng. 2020, 8, 914−925
ACS Sustainable Chemistry & Engineering Research Article

■ REFERENCES
(1) Zhu, Y.; Romain, C.; Williams, C. K. Sustainable polymers from
(20) Chen, R.; Zhang, C.; Kessler, M. R. Anionic waterborne
polyurethane dispersion from a bio-based ionic segment. RSC Adv.
2014, 4, 35476−35483.
renewable resources. Nature 2016, 540, 354−362. (21) Lu, Y.; Xia, Y.; Larock, R. C. Surfactant-free core−shell hybrid
(2) Ji, S.; Cao, W.; Yu, Y.; Xu, H. Visible-Light-Induced Self-Healing
latexes from soybean oil-based waterborne polyurethanes and
Diselenide-Containing Polyurethane Elastomer. Adv. Mater. 2015, 27,
poly(styrene-butyl acrylate). Prog. Org. Coat. 2011, 71, 336−342.
7740−7745.
(22) Gaddam, S. K.; Palanisamy, A. Anionic waterborne polyur-
(3) Hazarika, D.; Karak, N. Biodegradable tough waterborne
ethane-imide dispersions from cottonseed oil based ionic polyol. Ind.
hyperbranched polyester/carbon dot nanocomposite:an approach
Crops Prod. 2017, 96, 132−139.
towards eco-friendly material. Green Chem. 2016, 18, 5200−5211.
(23) Gaddam, S. K.; Palanisamy, A. Anionic waterborne polyur-
(4) Kang, S.-Y.; Ji, Z.; Tseng, L.-F.; Turner, S. A.; Villanueva, D. A.;
ethane dispersions from maleated cotton seed oil polyol carrying
Johnson, R.; Albano, A.; Langer, R. Langer,Robert. Design and
ionisable groups. Colloid Polym. Sci. 2016, 294, 347−355.
Synthesis of Waterborne Polyurethanes. Adv. Mater. 2018, 30,
(24) Bullermann, J.; Friebel, S.; Salthammer, T.; Spohnholz, R.
No. 1706237.
Novel polyurethane dispersions based on renewable raw materials
(5) Chattopadhyay, D. K.; Raju, K. V. S. N. Structural engineering of
Stability studies by variations of DMPA content and degree of
polyurethane coatings for high performance applications. Prog. Polym.
Sci. 2007, 32, 352−418. neutralisation. Prog. Org. Coat. 2013, 76, 609−615.
(6) Zhang, C.; Garrison, T. F.; Madbouly, S. A.; Kessler, M. R. (25) Taviot-Guého, C.; Prévot, V.; Forano, C.; Renaudin, G.;
Recent advances in vegetable oil-based polymers and their Mousty, C.; Leroux, F. Tailoring Hybrid Layered Double Hydroxides
composites. Prog. Polym. Sci. 2017, 71, 91−143. for the Development of Innovative Applications. Adv. Funct. Mater.
(7) Zafar, F.; Ghosal, A.; Sharmin, E.; Chaturvedi, R.; Nishat, N. A 2018, 28, No. 1703868.
review on cleaner production of polymeric and nanocomposite (26) Klöcking, R.; Felber, Y.; Guhr, M.; Meyer, G.; Schubert, R.;
coatings based on waterborne polyurethane dispersions from seed Schoenherr, J. I. Development of an innovative peat lipstick based on
oils. Prog. Org. Coat. 2019, 131, 259−275. the UV-B protective effect of humic substances. Mires Peat 2013, 11,
(8) Xia, Y.; Larock, R. C. Vegetable oil-based polymeric materials: No. 03.
synthesis, properties, and applications. Green Chem. 2010, 12, 1893− (27) Montenegro, L.; Santagati, L. M. Use of Vegetable Oils to
1909. Improve the Sun Protection Factor of Sunscreen Formulations.
(9) Liang, H.; Feng, Y.; Lu, J.; Liu, L.; Yang, Z.; Luo, Y.; Zhang, Y.; Cosmetics 2019, 6, No. 25.
Zhang, C. Bio-based cationic waterborne polyurethanes dispersions (28) Dang, C.; Wang, M.; Yu, J.; Chen, Y.; Zhou, S.; Feng, X.; Liu,
prepared from different vegetable oils. Ind. Crops Prod. 2018, 122, D.; Qi, H. Transparent, Highly Stretchable, Rehealable, Sensing, and
448−455. Fully Recyclable Ionic Conductors Fabricated by One-Step Polymer-
(10) Fu, C.; Zheng, Z.; Yang, Z.; Chen, Y.; Shen, L. A fully bio-based ization Based on a Small Biological Molecule. Adv. Funct. Mater. 2019,
waterborne polyurethane dispersion from vegetableoils: From syn- 29, No. 1902467.
thesis of precursors by thiol-ene reaction to study of final material. (29) Alagi, P.; Choi, Y. J.; Seog, J.; Hong, S. C. Efficient and
Prog. Org. Coat. 2014, 77, 53−60. quantitative chemical transformation of vegetable oils to polyols
(11) Wang, H.; Fan, J.; Fei, G.; Lan, J.; Zhao, Z. Preparation and through a thiol-ene reaction for thermoplastic polyurethanes. Ind.
property of waterborne UV-curable chain-extended polyurethane Crops Prod. 2016, 87, 78−88.
surface sizing agent: Strengthening and waterproofing mechanism for (30) Wang, X.; Zhang, Y.; Liang, H.; Zhou, X.; Fang, C.; Zhang, C.;
cellulose fiber paper. J. Appl. Polym. Sci. 2015, 132, No. 42354. Luo, Y. Synthesis and properties of castor oil-based waterborne
(12) Delpierre, S.; Willocq, B.; Manini, G.; Lemaur, V.; Goole, J.; polyurethane/sodium alginate composites with tunable properties.
Gerbaux, P.; Cornil, J.; Dubois, P.; Raquez, J.-M. Simple Approach for Carbohydr. Polym. 2019, 208, 391−397.
a Self-Healable and Stiff Polymer Network from Iminoboronate-Based (31) Resetco, C.; Dikić, T.; Verbrugge, T.; Du Prez, F. E. UV-cured
Boroxine Chemistry. Chem. Mater. 2019, 31, 3736−3744. multifunctional coating resins prepared from renewable thiolactone
(13) Liu, X.; Pang, C.; Ma, J.; Gao, H. Random Copolycarbonates derivatives. Prog. Org. Coat. 2017, 107, 75−82.
Based on a Renewable Bicyclic Diol Derived from Citric Acid. (32) Ma, S.; Kovash, C. S.; Webster, D. C. Effect of solvents on the
Macromolecules 2017, 50, 7949−7958. curing and properties of fully bio-based thermosets for coatings. J.
(14) Pang, C.; Jiang, X.; Yu, Y.; Liu, X.; Lian, J.; Ma, J.; Gao, H. Coat. Technol. Res. 2017, 14, 367−375.
Sustainable Polycarbonates from a Citric Acid-Based Rigid Diol and (33) Andjelkovic, D. D.; Valverde, M.; Henna, P.; Li, F.; Larock, R.
Recycled BPA-PC: From Synthesis to Properties. ACS Sustainable C. Novel thermosets prepared by cationic copolymerization of various
Chem. Eng. 2018, 6, 17059−17067. vegetable oilssynthesis and their structure−property relationships.
(15) Yu, Y.; Pang, C.; Jiang, X.; Yang, Z.; Ma, J.; Gao, H. Polymer 2005, 46, 9674−9685.
Copolycarbonates Based on a Bicyclic Diol Derived from Citric Acid (34) Zhang, Y.; Li, Y.; Li, J.; Gao, Y.; Tan, H.; Wang, K.; Li, J.; Fu,
and Flexible 1,4-Cyclohexanedimethanol: From Synthesis to Proper- Q. Synthesis and antibacterial characterization of waterborne
ties. ACS Macro Lett. 2019, 8, 454−459. polyurethanes with gemini quaternary ammonium salt. Sci. Bull.
(16) Zhou, X.; Fang, C.; Lei, W.; Du, J.; Huang, T.; Li, Y.; Cheng, Y. 2015, 60, 1114−1121.
Various nanoparticle morphologies and surface properties of water- (35) Chandra, S.; Karak, N. Environmentally Friendly Polyurethane
borne polyurethane controlled by water. Sci. Rep. 2016, 6, No. 34574. Dispersion Derived from Dimer Acid and Citric Acid. ACS Sustainable
(17) Liu, L.; Lu, J.; Zhang, Y.; Liang, H.; Liang, D.; Jiang, J.; Lu, Q.; Chem. Eng. 2018, 6, 16412−16423.
Quirino, R. L.; Zhang, C. Thermosetting polyurethanes prepared with (36) Fu, C.; Hu, X.; Yang, Z.; Shen, L.; Zheng, Z. Preparation and
the aid of a fully bio-based emulsifier with high bio-content, high solid properties of waterborne bio-based polyurethane/siloxane cross-
content, and superior mechanical properties. Green Chem. 2019, 21, linked films by an in situ sol−gel process. Prog. Org. Coat. 2015,
526−537. 84, 18−27.
(18) Lu, Y.; Larock, R. C. Soybean-Oil-Based Waterborne (37) Gurunathan, T.; Chung, J. S. Physicochemical Properties of
Polyurethane Dispersions: Effects of Polyol Functionality and Hard Amino−Silane-Terminated Vegetable Oil-Based Waterborne Polyur-
Segment Content on Properties. Biomacromolecules 2008, 9, 3332− ethane Nanocomposites. ACS Sustainable Chem. Eng. 2016, 4, 4645−
3340. 4653.
(19) Gurunathan, T.; Mohanty, S.; Nayak, S. K. Isocyanate (38) Xia, Y.; Larock, R. C. Castor-Oil-Based Waterborne Polyur-
terminated castor oil-based polyurethane prepolymer: Synthesis and ethane Dispersions Cured with an Aziridine-Based Crosslinker.
characterization. Prog. Org. Coat. 2015, 80, 39−48. Macromol. Mater. Eng. 2011, 296, 703−709.

924 DOI: 10.1021/acssuschemeng.9b05477


ACS Sustainable Chem. Eng. 2020, 8, 914−925
ACS Sustainable Chemistry & Engineering Research Article

(39) Lavilla, C.; Alla, A.; de Ilarduya, A. M.; Muñoz-Guerra, S. High


T(g) bio-based aliphatic polyesters from bicyclic D-mannitol.
Biomacromolecules 2013, 14, 781−793.
(40) Rabnawaz, M.; Liu, G. Graft-copolymer-based approach to
clear, durable, and anti-smudge polyurethane coatings. Angew. Chem.,
Int. Ed. 2015, 54, 6516−6520.
(41) Fu, C. Q.; Xu, M.; Yang, Z.; Lv, Q. G.; Shen, L. Emulsifier-Free
Core-Shell Hybrid Latexes from Castor Oil-Based Waterborne
Polyurethane and Polyacrylate Containing Fluorine and Silicon.
Adv. Mater. Res. 2013, 690−693, 1620−1623.
(42) Yong, Q.; Liao, B.; Huang, J.; Guo, Y.; Liang, C.; Pang, H.
Preparation and characterization of a novel low gloss waterborne
polyurethane resin. Surf. Coat. Technol. 2018, 341, 78−85.
(43) Yu, F.; Cao, L.; Meng, Z.; Lin, N.; Liu, X. Y. Crosslinked
waterborne polyurethane with high waterproof performance. Polym.
Chem. 2016, 7, 3913−3922.
(44) Zhong, Z.; Luo, S.; Yang, K.; Wu, X.; Ren, T. High-
performance anionic waterborne polyurethane/Ag nanocomposites
with excellent antibacterial property via in situ synthesis of Ag
nanoparticles. RSC Adv. 2017, 7, 42296−42304.
(45) Wei, H.; Ding, D.; Wei, S.; Guo, Z. Anticorrosive conductive
polyurethane multiwalled carbon nanotube nanocomposites. J. Mater.
Chem. A 2013, 1, 10805−10813.
(46) Mohammadi, A.; Barikani, M.; Doctorsafaei, A. H.; Isfahani, A.
P.; Shams, E.; Ghalei, B. Aqueous dispersion of polyurethane
nanocomposites based on calix[4]arenes modified graphene oxide
nanosheets: Preparation, characterization, and anti-corrosion proper-
ties. Chem. Eng. J. 2018, 349, 466−480.
(47) Li, J.; Cui, M.; Tian, H.; Wu, Y.; Zha, F.; Feng, H.; Tang, X.
Facile fabrication of anti-corrosive superhydrophobic diatomite
coatings for removal oil from harsh environments. Sep. Purif. Technol.
2017, 189, 335−340.
(48) Gaddam, S. K.; Kutcherlapati, S. N. R.; Palanisamy, A. Self-
Cross-Linkable Anionic Waterborne Polyurethane−Silanol Disper-
sions from Cottonseed-Oil-Based Phosphorylated Polyol as Ionic Soft
Segment. ACS Sustainable Chem. Eng. 2017, 5, 6447−6455.

925 DOI: 10.1021/acssuschemeng.9b05477


ACS Sustainable Chem. Eng. 2020, 8, 914−925

You might also like