Jialiang
Lai†
a,
Xijin
Xing†
b,
Huanzhi
Feng
b,
Zhanhua
Wang
*a and
Hesheng
Xia
*a
aState Key Laboratory of Polymer Materials Engineering, Polymer Research Institute, Sichuan University, Chengdu 610065, China. E-mail: zhwangpoly@scu.edu.cn; xiahs@scu.edu.cn
bCNOOC Research Institute Co., Ltd, Beijing 100028, China
First published on 7th September 2023
Polymer materials with covalent adaptive networks (CANs) structures have attracted considerable attention in recent years due to their excellent repairable, reprocessable, reconfigurable, recyclable, and re-adhesive (5R) performance. Many types of CANs based on reversible dissociation or association reactions have been developed. Of these, CANs via dynamic isocyanate chemistry have made significant progress on the creation of smart polyurethane (PU) and polyurea (PUR) materials. Herein, we provide a comprehensive review on the recent development of CANs via dynamic isocyanate chemistry. First, we provide a brief introduction of dynamic isocyanate chemistry. Second, several categories of dynamic isocyanate chemistry (and the mechanism behind them) are discussed in detail. Third, we focus on the characterization of CANs via dynamic isocyanate chemistry by physical and chemical approaches. Fourth, we focus on novel types of “smart” polymer materials containing a CANs structure with 5R properties via dynamic isocyanate chemistry. The influence of different categories of dynamic isocyanate chemistry on the stress relaxation and 5R performance are summarized in detail in this part. The advantages and disadvantages of different types of dynamic isocyanate chemistry for 5R applications are also discussed. Finally, conclusions and the outlook on the development and challenges of CANs via dynamic isocyanate chemistry are provided.
Huanzhi Feng | Huanzhi Feng is a senior oilfield chemistry engineer at CNOOC Research Institute. His research interest mainly focuses on deep-water drilling and well integrity. |
Fig. 1 Dynamic process of dissociative and associative CANs with 5R performance formed by different categories of dynamic isocyanate chemistry where X denotes O, S, NH, or NR groups (schematic). Reproduced from ref. 166 with permission from the American Association for the Advancement of Science, copyright 2018. Reproduced from ref. 87 with permission from the American Chemical Society, copyright 2020. Reproduced from ref. 67 with permission from the Royal Society of Chemistry, copyright 2020. |
Dynamic isocyanate chemistry can involve dissociative or associative exchange during thermal processing (Fig. 1). In theory, the crosslinking density remains constant for associative CANs via dynamic isocyanate chemistry during the dynamic process if side reactions are absent. There is virtually no difference in the rate constant of the forward reaction and reversible reaction because the reactant and product are usually the same type of compound (Fig. 1). The dynamic process is a kinetic control step dependent only upon temperature. An obvious exchange reaction induces network rearrangement above the topological temperature, leading to a 5R performance.42–45 As for dissociative CANs via dynamic isocyanate chemistry, the rate constant of the forward reaction should be larger than the reversible reaction; and then stable polymer materials can be prepared (Fig. 1).13 The crosslinking density decreases obviously with increasing temperature, leading to the destruction of polymer networks and decrease in the system viscosity, and resulting in 5R performance.46
One review summarized recent developments on dynamic covalent polymers enabled by reversible isocyanate chemistry. Different preparation approaches of dynamic covalent polymers based on reversible isocyanate chemistry were described for applications in self-healing materials, recycling, shape-memory polymers, and three-dimensional (3D) printing.47 Different from that review article, herein we focus on recent developments on CANs via dynamic isocyanate chemistry and the derived “smart” polymer materials with 5R performance.
First, we provide a brief introduction of dynamic isocyanate chemistry. Second, three main categories of dynamic isocyanate chemistry and the dynamic mechanisms behind them are discussed. Third, we focus on characterization of CANs via dynamic isocyanate chemistry by physical and chemical approaches. Fourth, we focus on novel kinds of smart polymer materials with 5R properties obtained via dynamic isocyanate chemistry. The relationship between the stress-relaxation character and 5R performance is discussed in detail in this part. The influence of different parameters on the stress-relaxation characteristics of CANs via dynamic isocyanate chemistry is discussed in this part. In addition, the advantages and disadvantages of different types of dynamic isocyanate chemistry for 5R applications are discussed. Finally, conclusions and the outlook for the development of CANs via dynamic isocyanate chemistry are provided.
Fig. 2 Possible mechanism for the reversible reaction of oxime-carbamate (A), and the forward (B) and reversible (C) reaction between phenol and isocyanate with tertiary amine as a catalyst (schematic). Reproduced from ref. 56 with permission from the Royal Society of Chemistry, copyright 2016. |
Although phenols react more slowly with isocyanates than alcohols, the formed phenol-carbamate moieties dissociate at lower temperatures than aliphatic urethanes, which is consistent with the slower rate of the reverse reaction. Amine bases are effective catalysts for forward and reverse reactions (Fig. 2B and C).56 The dissociation temperature and rate depend on the electronic effects and acidity values of the nucleophilic agents. The phenolate anion is bonded loosely with the carbonyl group with a phenol compound possessing high acidity. The nucleophilic agent dissociates at lower temperatures with electron-withdrawing substituents such as chlorine, trifluoromethyl and ester groups, and at higher temperatures with electron-releasing substituents such as methyl and methoxy groups.56–58 Similar to phenol, urethane formed by pyridinol and isocyanate is also a type of dynamic covalent bond.59 The electron-withdrawing substituent on the ortho position reduces the electron density and nucleophilicity of the hydroxyl group. Thus, the urethane bond will be more labile and the dissociation temperature will be lower. In addition, high steric hindrance blocks the hydroxyl group to attack the NCO group, so the urethane bond becomes more labile.60
The dissociation transcarbamoylation reaction also occurs. The exchange rate is highly dependent on the catalyst category and content. The DBTDL catalyst used for the formation of carbamate bonds also promotes its dynamic exchange at high temperatures.61,62 Whether the carbamate exchange reaction undergoes an associative or dissociative process depends on the catalyst, temperature, and position of the free hydroxyl.34,63–65 Dissociative exchange dominates the carbamate bonds exchange reaction in the presence of stannous octanoate irrespective of whether a free hydroxyl group is left (Fig. 3).66 If a free hydroxyl group is involved in the transcarbamoylation reaction, the dissociative process is much slower due to the competitive coordination of the free hydroxyl group (Fig. 3A). In the absence of a free hydroxyl group, the urethane group can bind more freely to the catalyst (stannous octanoate) to induce a rapid dissociation reaction (Fig. 3B). p-Toluenesulfonic acid also catalyzes the transcarbamoylation reaction via associative exchange in the presence of a free hydroxyl group.67 If DBTDL is used as a catalyst, associative and dissociative exchange reactions occur for carbamate bonds.67 If the hydroxyl group is located at the β-position of the carbamate moiety, an obvious exchange reaction occurs at 140 °C.64 The transcarbamoylation reaction via an associative process can be activated by mechanical stress due to the twisting of nitrogen lone pairs out of conjugation with the π orbitals of the carbonyl group. Hence, whether carbamate bonds underdo associative or dissociative exchange reactions is a complicated process, and can be influenced strongly by the: nature of the isocyanate group; catalyst species and content; free hydroxyl content and position; temperature. This challenge is hindered further due to the challenging capture of isocyanate species during the dynamic process because the regenerated isocyanate species are transient at most temperatures.67
Fig. 3 Transcarbamoylation reaction with Sn(Oct)2 as a catalyst in the presence (A) or absence (B) of free hydroxyl groups (schematic). Reproduced from ref. 66 with permission from the American Chemical Society, copyright 2019. |
Fig. 4 Possible dynamic mechanism of urea bonds (schematic). A: Urea formed by a secondary amine with a bulky substitution group. Reproduced from ref. 18 with permission from the Springer Nature, copyright 2014. B: Urea formed by pyrazole. Reproduced from ref. 74 with permission from the Springer Nature, copyright 2019. C: Urea formed by a primary aliphatic amine with zinc acetate as a catalyst. Reproduced from ref. 19 with permission from the Royal Society of Chemistry, copyright 2019. |
The dissociation temperature of the urea bond formed by a secondary amine is much lower than that of the primary amine due to the steric effect of the substituent,48,73 and decreases with an increase in the substituent volume.18 The reversible dissociation and reformation process of urea bonds bearing a hindered group on the nitrogen atom is shown in Fig. 4A. Dissociation of the urea bond formed by pyrazole and isocyanate is promoted by undergoing an intermediate five-membered-ring state (Fig. 4B). Theoretical calculations indicate that the pathway with nucleophilic addition is a rate-limiting step that accords with first-order kinetics. The resonance energy in pyrazole–urea bonds is lower than that of normal urea bonds formed by a primary amine, but is higher than that of hindered urea bonds. Therefore, the stability and reversibility of a pyrazole–urea moiety originates from the aromatic character of pyrazole moderately weakening the resonance stabilization and the presence of an adjacent nitrogen atom facilitating intramolecular 1,4-hydrogen transfer.74 The dissociation rate of the pyrazole–urea moiety increases if the α-carbon is substituted by groups with high steric hindrance. The kd increases from 0.00654 s−1 to 0.0457 s−1 if the neighboring group of α-carbon is changed from hydrogen to the tert-butyl group.75
Fig. 5 Possible thiourethane exchange mechanisms (schematic). A: Exchange mechanisms occurring via nucleophilic (route A) and basic (route B) catalysis. B: Thiourethane exchange mechanism for a combined nucleophilic and basic catalyst, DBN. Reproduced from ref. 80 with permission from Elsevier, copyright 2020. |
Fig. 6 A: Stress-relaxation curves commonly used to monitor the network exchange of CANs. B: Arrhenius plot of ln(τ*) against the inverse of temperature. |
The relaxation active energy (Ea) can be calculated from the Arrhenius plot of ln(τ*) against the inverse of temperature (Fig. 6B). Theoretically, bond-exchange reactions should be the primary contribution to Ea.84 The Ea can be affected by several aspects of the network, including polymer identity, topology, and crosslinking density. As such, CANs with the same dynamic chemistry may demonstrate different values of Ea in different systems.83 The temperature range used for the stress-relaxation experiment varies. In principle, the start temperature should be higher than the glass transition temperature (Tg) because stress relaxation can occur due to segment movement. If side reactions are absent, this temperature range should be very wide for associative CANs because the crosslinking density is well maintained during the exchange process. With respect to dissociative CANs, this temperature range should be between Tg and the gel-to-sol transition temperature (Tgel). This is because a sufficient number of crosslinking points will be broken above Tgel, which leads to thermoplastic-like segment flow and the Arrhenius law is no longer obeyed.79 Therefore, to ascertain of CANs undergo dissociative or associative isocyanate chemistry, a wide temperature range for stress-relaxation tests should be used. In principle, a linear function between relaxation time and inverse of the temperature should be obtained in a wide temperature range for CANs via associative isocyanate chemistry due to the maintained crosslinking density. Conversely, this range is relatively narrow for dissociative CANs. A further increase in the temperature will result in a gel-to-sol transition, and finally leads to the recovery of monomers or the production of homo-polymers with low molecular weight.
Investigation of evolution of the storage modulus with temperature by DMA or a rheology experiment will show a rubber plateau for associative CANs in a wide range of temperatures due to the maintained crosslinking density. The plateau modulus (G0) varies very weakly with temperature. In comparison, an obvious rubber plateau or a relatively narrow rubber state is not observed for CANs via dissociative isocyanate chemistry. In most cases, the modulus deceases with an increase in temperature due to the equilibrium shifting towards the dissociation of the product. This is because the strong increase in temperature leads to disruption of the network, which results in a decrease in crosslinking density and, finally, Tgel is obtained from the DMA curve. This obvious G0 variation with a temperature increase and generation of Tgel can be used to determine the dynamic mechanism. For example, Rowan and coworker developed a series of CANs by hindering the urea bond. The rubber state became narrower with an increase in hindrance. When the urea bond was connected with the 2,2,6,6-tetramethyl-4-piperidinol group, almost no rubber state was observed, which confirmed the dissociation mechanism.85
Similar to stress relaxation, creep experiments can also determine if the dynamic crosslinked network structure is associative or dissociative. Creep experiments lead directly to obtaining the shear viscosity, which can be used to calculate the viscous flow energy. As expected from networks maintaining a constant crosslinking density, comparable values for viscous-flow activation energy (Ea) and kinetic Ea can be obtained for associative CANs via dynamic isocyanate chemistry. In theory, in the plateau temperature range, a linear relationship can be obtained between shear viscosity and inverse of the temperature. For dissociative systems, this situation varies differently due to the change in crosslinking density. Montarnal et al. proposed that dissociative CANs display four distinct viscoelastic regimes.86 The crosslinking density decreases dramatically between Tg and Tgel, and the shear viscosity exhibits a stronger temperature dependence. Although a linear function between the shear viscosity and inverse of the temperature has also been reported, the temperature range was very narrow. A comparable Ea from stress-relaxation and creep experiments has been reported for dissociative CANs, but the temperature range varied differently. Therefore, conducting a creep experiment in a wide temperature range is essential to judge if a CAN undergoes an associative or dissociative process. For example, a CAN containing reversible acylsemicarbazide moieties displays almost the same stress-relaxation Ea (100.0 kJ mol−1) and viscous-flow Ea (99.6 kJ mol−1), but it dissociates with the temperature increase, which has been verified by the in situ IR analysis.87
Another common approach for characterizing the dynamic mechanism is through spectroscopy. The most popular form spectroscopy is in situ FTIR. The obvious signal change at 2250–2270 cm−1 attributed to the isocyanate group in IR spectra can show whether CANs undergo a dissociative process. Obtaining the threshold temperature for the dissociation of urethane, urea, or thiourethane groups is difficult because a small part of the NCO group in the IR spectra cannot be captured readily.
In this part, we discuss how these different categories of dynamic isocyanate chemistry affect the 5R performance through varying the polymer identity, catalyst, free hydroxyl group, and relaxation temperature. Also, because stress relaxation is used to study the network arrangement of CANs, we also discuss the correlation between stress relaxation and the 5R performance.
Fig. 7 A and B: Schematic illustration and optical images showing of the self-healing mechanism for polyurea containing tert-butyl as a hindered group. B: Preparation and reprocessing of crosslinked polyurea containing tert-butyl as a hindered group (schematic). Reproduced from ref. 18 with permission from Springer Nature, copyright 2014. C: Schematic illustration of the dynamic mechanism of the polyurea formed by a primary amine and isocyanate. Optical images showing the damged (D1–D2) and repaired (D3–D4) polyurea formed by a primary amine and isocyanate. Reproduced from ref. 68 with permission from the Royal Society of Chemistry, copyright 2020. |
Dynamic reaction | Temperature/°C | E a/kJ mol−1 | Conditions | 5R performance | Ref. |
---|---|---|---|---|---|
110–150 | 114 | 130 °C 40 min | Reconfiguration | 61 | |
140–160 | 143–188 | 160 °C 8 MPa 12 min | Reprocessable | 88 | |
160–200 | 140 | 160 °C, 6 MPa 0.5 h | Reprocessable | 89 | |
120–160 | 110 | 150 °C 60 min | Reconfiguration | 90 | |
110–150 | 139–165 | — | — | 66 | |
170–190 | 109–196 | 160 °C 30 MPa 4 h | Reprocessable | 91 | |
160 °C 30 MPa 1 h | Repairable (with pressure) | ||||
110–140 | 143 | 160 °C 2.9 MPa 12 min | Reprocessable | 92 | |
110–140 | 146 | 160 °C 3.5 MPa 12 min | Reprocessable | 92 | |
110–140 | 146 | 160 °C 3.5 MPa 12 min | Reprocessable | 92 | |
140–170 | 184 | 160 °C 40 min | Reconfiguration | 93 | |
170–190 | 99–134 | 160 °C 4 MPa 4 min annealed 225–580 min | Reprocessable | 82 | |
140–170 | 121 | — | — | 94 | |
140–160 | 91–127 | 160 °C 4 MPa | Reprocessable | 95 | |
150–180 | 86–105 | 200 °C 10 MPa 3 h | Reprocessable and recyclable | 96 | |
140–170 | 85–131 | 140 °C 10 MPa 2 h | Reprocessable | 97 | |
150–165 | 82–135 | 150 °C MPa 3 h | Reprocessable and reconfiguration | 98 | |
150 °C 4 h | |||||
110–170 | 92 | 130 °C 10 MPa 10 min | Reprocessable | 99 | |
150–165 | 77 | 150 °C 10 MPa 6 h | Reprocessable | 100 | |
150 °C 5 h | Reconfiguration | ||||
60–120 | 55 | 100 °C, 130 °C 5–10 MPa 2 h, | Repairable, reprocessable and recyclable | 101 | |
160 | — | 160 °C 0.2 MPa | Reprocessable | 102 | |
160 °C 3 h | Reconfiguration | ||||
120–160 | 135 | 140 °C 11 MPa 2 h | Reprocessable | 103 | |
110–115 | 118 | 130 °C 1.5–2 Mt 30 min | Reprocessable | 104 | |
80–120 | 124 | 110 °C 10 MPa 30 min, 110 °C 30 min | Reprocessable and repairable (scratch healing) | 17 | |
80–130 | 98–110 | 120 °C 10 MPa 30 min, 120 °C 10 min | Reprocessable and repairable | 16 | |
120–160 | 145 | 130 °C 5 MPa 1 h, 80 °C 15 h | Reprocessable and repairable | 105 | |
80–90 | 105 | 120 °C 5 MPa 10 min | Reprocessable | 106 | |
140–150 | 183 | 140 °C 5 MPa 10 min | Reprocessable | 106 | |
130–170 | 110 | 100 °C 5 MPa 30 min, 80 °C 10 s | Reprocessable and reconfiguration | 107 | |
170–190 | 60–117 | 140 °C 4 MPa 40 min, 808 nm NIR laser | Reconfiguration and re-adhesive | 108 | |
40–80 | 93 | 100 °C 5 MPa 1 min, 80 °C 120 min | Reprocessable and repairable | 109 | |
180–240 | 66–107 | 130 °C 5 MPa 1 h, 140 °C 0.5 h | Reprocessable and repairable | 110 | |
70–120 | 100–120 | 130 °C 15 MPa 20 min | Reprocessable | 111 | |
60–100 | 60–128 | 100 °C 10 MPa 10 min, 100 °C, 2 h | Reprocessable and repairable | 112 | |
70–120 | 101 | 90 °C−140 °C 5 min 5 MPa | Reprocessable | 113 | |
Annealed 40 °C for 12 h | |||||
70–100 | 52 | 90 °C 24 h | Repairable | 19 | |
100–140 | 29 | NIR 5 min | Reprocessable and repairable | 68 | |
125–145 | 120 | 130 °C 1 h 20 MPa | Reprocessable and repairable (with solvent) | 114 | |
140–160 | 100 | 140 °C 15 MPa 1 h | Reprocessable | 87 | |
120 °C 1 h DMF | Repairable (with solvent) | ||||
90–130 | 109 | 130 °C 10 MPa 30 min | Reprocessable | 74 | |
90–130 | 81 | 120 °C 10 MPa 30 min | Reprocessable and repairable | 75 | |
110–140 | 80 | 120 °C 10 MPa 30 min | Reprocessable and repairable | 75 | |
120–160 | 57 | 120 °C 10 MPa 30 min | Reprocessable and repairable | 75 | |
130–155 | 158 | 150 °C 5 MPa 2 h | Reprocessable and repairable | 115 | |
90–110 | 182 | 150 °C 5 MPa 2 h | Reprocessable and repairable (partial scratch healing) | 115 | |
140–165 | 66 | 150 °C 5 MPa 2 h | Reprocessable and repairable (partial scratch healing) | 115 | |
90–110 | 85 | 150 °C 5 MPa 2 h | Reprocessable and repairable (partial scratch healing) | 115 | |
120–160 | 91–98 | 130 °C 0.5 MPa 5 min | Reprocessable and repairable | 116 | |
130 °C 0.5 MPa 10 min | |||||
120–150 | 101 | 150 °C 10 min | Reconfiguration | 117 | |
130–160 | 74 | 80 °C 16 MPa 1 h | Reprocessable | 118 | |
35–65 | 2.5–46 | RT 200 kPa, 35 °C 24 h | Reprocessable and repairable | 119 | |
70–100 | 119 | 80 °C 60 min | Repairable, reprocessable | 120 | |
60 °C 10 MPa 10 min | Re-adhesive | ||||
80 °C 60 min | |||||
90–130 | 92 | 110 °C 12 h–48 h, 110 °C 15 MPa 1 h | Repairable and reprocessable | 121 | |
70–120 | 52–69 | 80 °C 3 MPa 60 min, 40 °C 12 h | Reprocessable, repairable and reconfiguration | 122 | |
80–170 | 109–113 | 120 °C 4 tons 25 min | Reprocessable | 123 | |
110–180 | 59 | — | — | 85 | |
110–180 | 130 | — | — | 85 | |
60–100 | 132 | — | — | 85 | |
130–170 | 105–128 | 170 °C 20 MPa 60 min, 40 °C 12 h | Reprocessable, re-adhesive | 124 | |
130–170 | 104 | 180 °C 25 min | Reprocessable | 125 | |
130–170 | 115 | 180 °C 25–40 min | Reprocessable | 125 | |
130–170 | 123 | 180 °C 40 min | Reprocessable | 125 | |
110–150 | 85 | 120 °C 5 MPa 30 min, 130 °C 30 min | Reprocessable and reconfiguration | 126 | |
110–180 | 34 | 150 °C 300 kPa 2 h, 130 °C 30 min | Reprocessable, repairable | 127 | |
100–130 | 71–229 | 100 °C 12 MPa 120 min | Reprocessable | 72 | |
110–130 | 121 | 120 °C 10 MPa 30 min | Reprocessable, repairable | 128 | |
120 °C 60 min | Reconfiguration | ||||
120 °C 30 min | |||||
140–170 | 96–125 | 150 °C 10 MPa 40 min | Reprocessable | 129 | |
130–200 | 138 | 150 °C 100 bar 1 h | Reprocessable | 70 | |
140–200 | 122–147 | 160 °C 100 bar 1 h | Reprocessable | 71 | |
120–180 | 119 | 140 °C 100 bar 1 h | Reprocessable and repairable (scratch healing) | 78 | |
150–180 | 75 | 80 °C 3 MPa 30 min | Reprocessable | 130 | |
100 °C 3 MPa 30 min | Re-adhesive | ||||
100 °C 3 MPa 30 min | Repairable (with pressure) | ||||
160–185 | 72–102 | 180 °C 40 min | Reconfiguration | 131 | |
160–185 | 186 | 165 °C 8 MPa 2.5 h | Reprocessable | 80 | |
160–185 | 91–107 | 165 °C 8 MPa 2.5 h | Reprocessable | 80 | |
170–190 | 137–192 | — | — | 132 |
Dynamic motif | Mechanical properties | Healing conditions | Healing efficiency/% | Ref. | ||
---|---|---|---|---|---|---|
Tensile stress/MPa | Break strain/% | Modulus/MPa | ||||
HDI: hexamethylene diisocyanate, IPDI: isophorone diisocyanate, RT: room temperature, DMF: N,N-dimethylformamide, NIR: near infrared, DOU: dimethylglyoxime–urethane. | ||||||
Hindered urea bonds | 0.9 | 300 | 1 | 37 °C, 12 h | 87 | 18 |
Hindered urea bonds | HDI 8 | 145 | — | 60 °C 4 d | 89 | 140 |
IPDI 9 | 105 | — | 60 °C 4 d | 87 | ||
Normal urea bonds | HDI 0.36 | 330 | — | 90 °C 24 h | 99 | 19 |
IPDI 7 | 650 | — | 110 °C 16 h | 90 | ||
Hindered urea bonds | 3.8 | 570 | — | 60 °C 0.5 h | 92 | 145 |
Hindered urea bonds | 17 | — | 600 | 140 °C 6 d | 40 | 146 |
Hindered urea bonds | 14 | 140 | 21.7 | 75 °C 7 d | 82 | 147 |
Hindered urea bonds | 8 | 2200 | 27 | RT | 62 | 148 |
Acylsemicarbazide bonds | 22.8 | 285 | 77.8 | 75 °C 7 d | 57 | 65 |
Hindered urea bonds | 2.31 | 529 | 1.52 | 40 °C 12 h | 97 | 122 |
Normal urea bonds | 5 | 820 | — | NIR 5 min | 90 | 68 |
Hindered urea bonds | 24 | 650 | — | 100 °C 0.5 h | 95 | 149 |
Hindered urea bonds | 5.9 | 997 | — | 70 °C 5 h | 99 | 150 |
Hindered urea bonds | >90 | 3–4 | 1006 | 70 °C 1 h | >95 | 151 |
Hindered urea bonds | >7 | >2600 | — | 90 °C 10 h | 96.5 | 152 |
Hindered urea bonds | 5.3 | 500 | — | 80 °C 6 d | 100 | 153 |
Hindered urea bonds | — | — | — | 75 °C 24 h | 100 | 154 |
Hindered urea bonds | 5.4 | 200 | — | 120 °C 0.5 h | 94 | 155 |
Hindered urea bonds | 7.8 | 600 | — | 50 °C 24 h | 95 | 156 |
Hindered urea bonds | 17.3 | — | 340 | 100 °C 72 h | 50 | 139 |
Hindered urea bonds | 10 | 150 | — | 60 °C 96 h | 75–95 | 140 |
Hindered urea bonds | 39.5 | 3.2 | 1900 | 100 °C 0.5 h 30 kPa | 95 | 157 |
Phenol-carbamate bonds | 14 | 550 | — | 140 °C 0.5 h | 96.5 | 158 |
Phenol-carbamate bonds | 10.4 | — | 250 | 150 °C 48 h | 85 | 58 |
Pyrazole urea bonds | 3.5 | 185 | 3.1 | 120 °C 1 h | 99 | 75 |
Phenol-carbamate bonds | 2.8 | 240 | — | 80 °C 2 h | 98 | 109 |
Phenol-carbamate bonds | 52 | 12 | — | 110 °C 0.5 h | >95 | 159 |
Phenol-carbamate bonds | 11 | 600 | — | 150 °C 8 h | 68 | 160 |
Phenol-carbamate bonds | 5.2 | 510 | — | 120 °C 8 h | 86 | 143 |
Acylsemicarbazide bonds | 69 | 181 | 1740 | 140 °C DMF 1 h | 99.7 | 87 |
Acylsemicarbazide bonds | 100 | 5 | 2840 | 120 °C 1 h | 94.4 | 114 |
Thiourethane bonds | 62.7 | <0.1 | 2000 | 120 °C 3 MPa 1 h | 99 | 130 |
Phenol-carbamate bonds | 46.4 | 615 | — | 100 °C 2 h | 93 | 161 |
Oxime-carbamate bonds | 5.38 | 620 | — | 90 °C 1.5 h | 94 | 162 |
Oxime-carbamate bonds | 13.5 | 812 | 10.9 | 100 °C 2 h | 99 | 17 |
Oxime-carbamate bonds | 7.43 | 677 | 4.79 | RT 75 min | 94 | 141 |
Hydroxyl-urethane exchange | 25.6 | 513 | 13.2 | 110 °C 0.5 h | 92 | 163 |
Most of healable polymer materials can repair only micro/nano-scratches; achieving repair of macro-scratches has rarely been achieved. Wang and coworkers developed a type of dynamic crosslinked PUR–polydimethylsiloxane elastomer in which polydopamine nanoparticles were introduced to induce rapid photothermal transition. The obtained composite elastomers exhibited a rapid solid–liquid transition under NIR exposure because of the fast dissociation rate of the high-density dynamic urea bonds present in the polymer network, and so demonstrated excellent macro-damage repair. A macro-size mechanical scratch (width: 1 mm; depth: 0.25 mm) can be repaired under NIR exposure for 5 min (Fig. 7C and D).68 High energy or catalysis is required to induce decomposition of the primary aliphatic urea bond to achieve repair, but recovery of the mechanical property can be obtained in minutes due to the rapid reformation of the urea linkage from the cleaved amine and isocyanate compound. The temperature generated by photothermal transition during repair far surpasses the temperature range (90–120 °C) used to determine the stress-relaxation Ea, which results in very fast segment relaxation, and leads to flowing of segments to achieve macro-damage repair.
The first typical self-healing PUs via dynamic isocyanate chemistry were developed by Xu and coworkers. They developed a class of poly(oxime-urethanes) prepared from the addition of oxime compounds and diisocyanate at ambient temperature.16 The oxime-carbamate structures could be made after being heated through oxime-enabled transcarbamoylation via a thermally dissociative mechanism, enabling efficient repairable performance. Xia and coworkers extended this dynamic reaction to prepare a self-healing PU elastomer.17 The poly(oxime-urethane) with a crosslink density of 0.2 mmol cm−3 possessed a tensile strength of 13.5 MPa, a breaking strain of 812%, toughness up to 40 MJ m−3, an excellent elastic recovery of 90%, and an average optical transmittance of 86% in the visible range. This material could be healed completely at 110 °C in 0.5 h. The healing temperature used for determining the Ea was 80–120 °C, which demonstrated that the reversible dissociation of the oxime-urethane bond was mainly responsible for the repair process. In situ structural characterizations revealed that the repairable properties originated from the reversibility of oxime-carbamate bonds and hydrogen bonds. You and coworkers reduced the repair temperature to room temperature by introducing a copper catalyst. The latter helped physical crosslinking through metal coordination and weakening the bond strength of oxime-urethane moieties. The as-developed copper-based PU elastomer displayed a tensile strength and toughness of 14.8 MPa and 87.0 MJ m−3, respectively, and could self-heal at room temperature due to the accelerated dissociation rate of the oxime-carbamate bond from the coordination of copper.141 Silica nanoparticles can be incorporated into a poly(oxime-urethane) elastomer to develop damage-tolerant self-healing composites.142 The examples stated above demonstrate that the dissociation reaction of oxime-urethane moieties can be finely adjusted by changing the segment flexibility or introducing metal coordination. Also, the derived reparable polymers can be thermosets or elastomers with a repair temperature spanning from room temperature to >100 °C. Huang and coworkers reported a thermal self-healing PU thermoset based on a dynamic phenolic urethane reaction.143 The phenolic urethane partially decomposed into isocyanates and phenolic hydroxyls above 120 °C and could reform into the phenolic urethane group after cooling down. These properties contributed to the thermal self-repair of the thermosetting PU network. The repaired PU thermoset recovered ∼70% of its tensile strength and 86% of its elongation at break. Different from the urea bond or oxime-urethane bond, the rate of the addition reaction of the phenol group with an isocyanate compound was relatively slow, which resulted in a longer self-healing process and sometimes in sufficient recovery of the mechanical property.
The temperature of self-healing spans from ambient temperature to 150 °C depending on the different categories of dynamic chemistry. Most of the reported self-healing systems can achieve >90% healing with sufficient energy input. Efforts should focus on how to obtain efficient self-healing with minimal energy input (including at a low temperature for a short time). Combination of metal coordination with dynamic isocyanate chemistry has been proposed to weaken the bond strength, leading to a lower self-healing temperature. Among all self-healing materials reported so far, the highest mechanical strength and Young's modulus can reach 100 MPa and 2.8 GPa, respectively. Usually, extra pressure is required to assist the healing process in these systems. NIR or UV irradiation and solvent assistance also facilitates the repair process. Although dynamic isocyanate chemistry is responsible for the healing process, other chemical events that occur during the damage–repair cycle should also be considered. For example, the hydrogen bonding formed among urethane/urea/thiourethane groups contribute significantly to the self-healing process. In addition, polymer self-healing is a complicated process; many chemical and physical events occur apart from the recombination of dynamic covalent bonds. In crosslinked PU networks, damage-induced formation of primary amines can attack urethane bonds if zinc acetate is the catalyst, which would lead to self-healing.65 Damage to part of PU networks containing methyl-α-D-glucopyranoside can lead to reactions with atmospheric CO2 and H2O, thus reforming covalent linkages to repair the cleaved network segments and resulting in self-healing.144 For materials with a designed architecture, energy input and the healing rate should be balanced to achieve self-healing while maintaining the geometry. For example, Sun et al. realized the repair of a 3D-printed sole without damaging the porous structure (Fig. 11C and D).
Tables 1 and 2 suggest that only dissociative chemistry via carbamate or urea moieties (including phenol-carbamate bonds, oxime-carbamate bonds, hindered urea bonds, and metal-catalyzed normal aliphatic urea bonds) has been employed to prepare repairable materials. This is because the dissociation of the dynamic moiety accelerates movement of the segment, thereby facilitating repairable processes. Most of these systems exhibit a stress-relaxation Ea <120 kJ mol−1 or <100 kJ mol−1 (Table 1). The repairable efficiency and rate depends heavily on the equilibrium constant of isocyanate chemistry. An increase in temperature will shift the equilibrium to the decomposition of carbamate or urea moieties, which enables disruption of the network and rapid movement of the segment, thereby filling the damage to realize repair. Rapid reformation of cleaved carbamate or urea moieties should also be considered to fix the shape of the damaged area and recover the mechanical performance, otherwise fast flowing of the segment may result in the loss of materials. If dynamic chemistry via a primary aliphatic urea bond is utilized to develop repairable materials, then fast and highly efficient repair can be achieved. On the contrary, the phenol-carbamate bond dissociates at a lower temperature than the primary aliphatic urea bond, but the longer repair time is due to the its slower rate of combination between phenol and isocyanate. Therefore, the forward and reverse reaction rates of dynamic isocyanate chemistry should be considered when designing “ideal” repairable materials. Compared with dissociative dynamic isocyanate chemistry, the associative dynamic chemistry between hydroxyl and carbamate groups or exchange reaction between carbamate groups to develop repairable materials have rarely been reported. Usually, the stress-relaxation Ea of these two reactions is >100 kJ mol−1 (Table 1), which produces relatively slow movement of the segment, thereby blocking the repair process. This phenomenon explains why repairable performance has been realized via the dissociative primary aliphatic urea bond, whereas reparable materials have not been obtained via the primary aromatic urea bond because it undergoes an associative process.
Fig. 8 Optical images showing the reprocessing process of a polyurea containing tert-butyl as a hindered group (A) and acylsemicarbazide group. Reproduced from ref. 157 with permission from John Wiley and Sons, copyright 2016. (B) Optical (C) and mechanical (D) properties of a polyacylsemicarbazide after being reprocessed four times. Reproduced from ref. 87 with permission from the American Chemical Society, copyright 2020. |
Fortman et al. reported a class of polyhydroxyurethane (PHU) CANs derived from the attack of amines towards six-membered cyclic carbonates.63 PHU networks can be reprocessed at 160 °C at a pressure of 4 MPa for 8 h in the absence of an external catalyst, and recover 75% of their as-synthesized values after reprocessing. A stress relaxation occurs through an associative process which undergoes nucleophilic addition of free hydroxyl groups to carbamate groups. DFT calculations revealed that the transcarbamoylation could be attributed to the twisting of nitrogen lone pairs out of conjugation with the π orbitals of the carbonyl group. In a subsequent study, they systematically investigated the reprocessability of crosslinked PHUs synthesized from five- or six-membered cyclic carbonates.82 They concluded that the higher thermodynamic stability of five-membered cyclic carbonates led to reversion and subsequent decomposition of PHUs at high temperatures, but this decomposition was not observed in networks derived from six-membered cyclic carbonates. Torkelson and coworkers reported another type of reprocessable PHU CAN obtained from five-membered cyclic carbonates, which displayed full recovery of properties due to concurrent associative and dissociative dynamic chemistry via transcarbamoylation and reversible cyclic carbonate aminolysis.64 The cyclic carbonate contents in the network increased after each reprocessing confirmed the reverse reaction of aminolysis under reprocessing conditions, which indicated that the de-crosslinking caused by reverse cyclic carbonate aminolysis could be fully reversed. Reprocessing took 2 h at 140 °C to complete. The examples given above demonstrate that the reprocessing of PHUs via the exchange reaction of hydroxyl and urethane groups require a high temperature and long time to achieve recovery of mechanical performance. This harsh processing condition may cause side reactions and limits the application of PHUs in daily life.
Thermal pressing is the widely used method for the characterization of CANs reprocessing. Injection and extrusion are popular processing techniques for making polymer parts, and deserve to be studied for CANs reprocessing. For example, Sheppard et al. realized the reprocessing of PU foam into bulk materials by introducing a catalyst (DBTDL) through twin-screw extrusion via a carbamate exchange reaction.88 The relaxation time of the reprocessed PU films was 28 s at 160 °C, which demonstrated very fast exchange reactions among carbamate groups to be responsible for this reprocessing. Whether the amount of metal catalyst needed accords with real applications warrants investigation.
Table 1 suggests that dissociative and associative dynamic isocyanate chemistry can be employed to develop reprocessable materials. External pressure can boost the melt fusion and bonding together via dynamic isocyanate chemistry. One must balance the reprocessing temperature and degree of side reactions. Increasing the temperature can accelerate the rate of bond exchange whether the dynamic isocyanate chemistry undergoes a dissociative or associative process, thereby facilitating reprocessing. The bond exchange reaction for associative isocyanate chemistry is a kinetic process. The reprocessing temperature should be higher than the topology freezing transition temperature for associative isocyanate chemistry. For slower-exchange isocyanate chemistry, a catalyst is needed to achieve excellent reprocessing. For example, DBTDL and Fe(acac)3 have been used as catalysts to realize the reprocessing of normal PU through the bond exchange reaction between carbamate groups. Dissociative isocyanate chemistry can induce gel-to-sol transition through breaking the network into monomers or oligomers, strongly reducing the melt viscosity. Therefore, reprocessing can be realized without shear force. For example, shear-free reprocessing can be obtained for PUR formed by primary aliphatic amine and aliphatic isocyanate compounds.
Stress relaxation has a major influence on reprocessing. For example, thermoset PU elastomers with 1 wt% DBTDL as a catalyst displaying an Ea of 173 kJ mol−1 can be hot-pressed at 160 °C and 30 MPa for 4 h to achieve a good reprocessing performance.91 Similarly, another type of PU exhibiting an Ea of 140 kJ mol−1 (ref. 89) and 146 kJ mol−1 (ref. 92) needs to be reprocessed at 160 °C and 6 MPa for 0.5 h and at 160 °C and 5–10 MPa for 12 min, respectively. Increasing the content of soft segments72 or the chain length between the crosslinking points,115 or the dynamic moiety content110 can accelerate the relaxation process via the fast bond exchange reaction, which reduces Ea and facilitates reprocessing.
Xie and coworkers found that the stress relaxation induced by the DBTDL-catalyzed carbamate exchange reaction enabled the intrinsic plasticity of classical thermoset shape-memory PU.61 The Ea determined from 110–150 °C increased from 113.6 to 130.5 kJ mol−1 by varying the amount of aliphatic isocyanate to aromatic isocyanate, indicating slower stress relaxation for aromatic PUs. Incorporating two types of dynamic moieties which display an obvious exchange reaction at different temperatures can tune the stress-relaxation behavior. Without DBTDL, the exchange chemistry for carbonates is very slow, which results in slow stress relaxation and a high Ea (184 kJ mol−1).93 Efficient reconfiguration also needs high energy input to be realized (170 °C). For example, a poly(urethane–urea) copolymer was synthesized by reacting HDI with a mixture of a diol, a hindered amine, and a triol crosslinker, with DBTDL as the catalyst.165 Urethane bond exchange and hindered urea bond exchange occurred at different temperature ranges. Varying their bond ratio could lead to great flexibility in tuning the topological rearrangement kinetics of the network to allow reprocessing and plasticity to be achieved in a wide temperature range. Increase in the ratio between hindered urea and urethane bonds resulted in faster stress relaxation. The dynamic networks exhibited excellent self-healing, reprocessability, and solid-state plasticity with an optimized ratio between hindered urea and urethane bonds.
Jin et al. designed a dynamic polymer network that responded to light and heat.166 Thermally induced transesterification and transcarbamoylation triggered the topological rearrangement of the network. The resulting solid-state plasticity allowed permanent shape reconfiguration by manual folding (e.g., from 2D sheet to 3D airplane) followed by stress relaxation at 140 °C. As shown in Fig. 9, flowers, airplanes, and elephants were heated (80 °C) and cooled to achieve a reversible shape change of a specific part. For photo-reversibility, at the same temperature (80 °C) and with the pre-stretching force imposed, subsequent exposure to light of 312 nm resulted in cinnamate dimerization, which partially fixed the alignment even after the force had been removed. Because of this partial alignment (or network anisotropy), the sample exhibited actuation behavior via a reversible shape-memory mechanism.
Fig. 9 Reconfigured polymers based on dynamic covalent bonding. (A) Thermally induced transesterification, transcarbamoylation, hindered urea bond exchange, and photo-reversible dimerization of nitro-cinnamate. (B) Thermally triggered reversible deformation behavior of 3D flowers. (C) Photograph and thermally triggered reversible shape change of crane and elephant. Reproduced from ref. 166 with permission from the American Association for the Advancement of Science, copyright 2018. (D) Reversible construction of complex shapes with thermal drive. Reproduced from ref. 61 with permission from John Wiley and Sons, copyright 2016. |
In principle, dissociative and associative isocyanate chemistry can be employed to construct reconfigurable polymer materials. Obvious bond exchange should occur during the reconfiguration process. For CANs via dissociative isocyanate chemistry, determination of the threshold temperature for reconfiguration is difficult. Normally, the temperature range (Table 1) used for determining the stress-relaxation energy could be a good choice to reshape materials. Within this range, a higher temperature requires a shorter time to fully relax the stress and finish the reshaping process. Cheng and coworkers developed a reconfigurable poly(urea–urethane) thermoset based on hindered urea bonds.117 An Ea of 101.4 kJ mol−1 was obtained at 120–150 °C (Table 1). Full reconfiguration was achieved at 150 °C for 5 min to fully relax the stress. The plasticity temperature should be lower than the Tgel for dissociative CANs, otherwise flowing hampers shape fixing. For CANs via dissociative isocyanate chemistry, reconfiguration occurs above the Tv. Introducing a catalyst into CANs enables tuning of the kinetics of the bond exchange reaction, which can govern the plasticity process. A plasticity time longer than the full relaxation time is required to realize complete reconfiguration. Otherwise partial reshaping or the shape-memory effect will occur. Hence, all the systems listed in Table 1 can be reconfigured at the temperature range used for calculating the Ea. The difference lies in the reshaping time and deformation degree, which can be adjusted by changing the crosslinking density, the ratio between soft and rigid segments, and dynamic moiety content.
Fig. 10 A: Preparation of PU–PUA materials with PFPE as a soft segment (schematic). B and C: Schematic illustration and optical photograph showing the degradation of PU–PUA materials in diethylene glycol. Reproduced from ref. 167 with permission from the Royal Society of Chemistry, copyright 2019. |
Due to the water sensitivity of isocyanate groups, another alternative route to achieve the degradation of PU/PUR is to hydrolyze these materials in hot water. If the urethane and urea bonds can undergo a dissociation reaction, then the generated isocyanate groups can react with water molecules and generate amine compounds to realize recycling. For example, PUR containing a hindered urea bond formed by t-butyl and isocyanate groups can be fully degraded in PBS within 100 h, whereas no degradation is observed by replacing t-butyl with a n-butyl group.168 Although no heating is needed to degrade hindered urea-based polymers, their long-term stability at ambient temperature under high humidity is a challenge. Heating the dynamic polymer based on other types of dissociative isocyanate chemistry via phenol-carbamate bonds, oxime-carbamate bonds, and secondary aliphatic urea bonds could be alternative approach to develop recycling polymers. Balancing the service stability with a fast degradation rate should be the key feature.
Another aspect closely related to re-adhesive performance is 3D printing of CANs. 3D printing of polymer materials is a layer-by-layer process. The interlayer strength is highly dependent upon the adhesive properties between neighboring layers, which determines the mechanical property in the printing direction. Incorporating dynamic covalent bonds into a polymer network could be a solution. Thermal treatment of the printed parts can improve the interlayer bonding strength due to the re-adhesive performance arising from the dynamic covalent bonding linkage. Sun et al. verified this idea by developing PU-based CANs containing dynamic halogenated bisphenol carbamate bonds, which could be processed by selective laser sintering for 3D printing. A dramatic increase in the Z-axis tensile strength of printed parts was achieved through reversible adhesion between neighboring layers via the reversible dissociation and reformation of bisphenol carbamate bonds.175 The relatively slow addition reaction between phenol and isocyanate groups as well as the corresponding dissociation reaction resulted in complete bonding between neighboring layers. Post-treatment by heating the printed parts further strengthened the adhesion between adjacent layers, imparting the printed parts with excellent mechanical performance at different printing orientations. In their further study, PDMS CANs with hindered pyrazole-urea dynamic bonds were developed for selective laser sintering for 3D printing (Fig. 11A).75 The strength of the printed material in the three directions (X, Y, and Z) was essentially identical (Fig. 11B), indicating that the printed parts showed an isotropic mechanical performance in different printing directions. The PDMS CANs exhibited obvious stress relaxation at 90–130 °C with an Ea of 19.4 kcal mol−1, and the relaxation modulus decayed to nearly zero at 120 °C within 3 min. An optimized mild temperature of 80 °C was selected for the thermal treatment of printed parts, which increased the interlayer bonding strength due to reversible dissociation–reformation of the hindered pyrazole urea dynamic bonds, but also maintained the size of the printed parts.
Fig. 11 Preparatory structure of a PDMS covalent adaptive network (A). Tensile strength of printed splines in three directions (X, Y, and Z) was compared (B). Shoe soles for printing preparation (C), photograph of a self-healing 3D printed shoe sole (D). Reproduced from ref. 75 with permission from Elsevier, copyright 2021. |
Fig. 12 Relationship between the activation energy and 5R performance via different categories of dynamic isocyanate chemistry. |
First, thermodynamics and kinetics are the most important parameters to determine the reversibility of isocyanate chemistry. They can be adjusted based on electronegativity and steric hindrance. If a chemical group possesses electronegativity and steric hindrance, the dynamic mechanism become very complex. Second, whether dynamic isocyanate chemistry undergoes dissociative or associative process is quite complicated. In most cases, a dissociative process predominates the dynamic process. If free hydroxyl or amine groups dangle in the network, these two processes may occur together, but one of them will predominate. Models studied by NMR spectroscopy, FTIR spectroscopy, and high-performance liquid chromatography can provide evidence of whether the dynamic process is dissociative or associative. DFT can be used to calculate which process has lower energy, further confirming the dynamic reaction. Third, different catalysts, such as acids/bases, have a big influence on the dynamic process of isocyanate chemistry because protonation has a very important role in the mechanism. Studies on acid/base-catalyzed dynamic isocyanate chemistry are warranted. Fourth, although a clear dynamic mechanism form isocyanate chemistry can be verified by model study of small compounds, if this dynamic isocyanate chemistry is incorporated into a polymer matrix to develop CANs materials, the dynamic process remains quite complicated.
Thanks to dynamic isocyanate chemistry, different types of smart polymer materials with 5R performance have been developed. With regard to developing stimuli-responsive polymer materials based on dynamic isocyanate chemistry, six main aspects need to be focused on in future studies.
First, urethane or urea bonds are a hydrogen-bonding motif that can contribute to the mechanical performance of polymers. In some cases, the hydrogen bond will accelerate the dissociation process, which means that the hydrogen bond and dynamic covalent bond work collaboratively. Optimizing the concentration and position of the urea or urethane bonds in the polymer network are crucial for tuning the dynamic properties of the material. Second, the dissociation reaction of some urea or urethane bond occurs at >110 °C, which leads to side reactions of the formed isocyanate group. This phenomenon will influence the thermal stability of the polymer material or result in its degradation. At high humidity, the newly generated isocyanate group will react with water, leading to the degradation of PU or PUR materials. Strategies for protecting the degradation of CANs via dynamic isocyanate chemistry should be considered during the design process. Third, some isocyanate-based chemistry can be made dynamic at room temperature. This is an advantage for preparing smart polymer materials with variable responsive properties at room temperature, such as automatic self-healing and solid reconfiguration. However, the drawback is instability at room temperature, which should be avoided. Fourth, reports about PU or PUR degradation based on dynamic isocyanate chemistry are lacking. The isocyanate group formed during urethane or urea dissociation can be captured by an extra hydrogen-based nucleophilic agent, leading to the degradation of the polymer material. Fifth, in some cases, free hydroxyl/amine groups suppress the reversion of urethane links and minimize the side reactions associated with liberated isocyanate groups under reprocessing conditions to maintain network integrity in the presence of few side reactions.176 Sixth, long-term service stability needs to be improved. For example, CANs via an oxime-urethane moiety can be degraded under UV irradiation.177
Footnote |
† These authors contributed equally to this paper. |
This journal is © The Royal Society of Chemistry 2023 |