Construction of functional coating with durable and broad-spectrum

Abstract. A novel surface coating with durable broad-spectrum antibacterial ability was prepared based on mussel-inspired dendritic polyglycerol (MI-d...
0 downloads 0 Views 4MB Size
Progress in Polymer Science 79 (2018) 26–39

Contents lists available at ScienceDirect

Progress in Polymer Science journal homepage: www.elsevier.com/locate/ppolysci

Review

Let spiropyran help polymers feel force! Meng Li a , Qi Zhang b , Yin-Ning Zhou a , Shiping Zhu a,∗ a b

Department of Chemical Engineering, McMaster University, Hamilton, Ontario, L8S 4L7, Canada College of Chemical Engineering, Zhejiang University of Technology, Hangzhou, 310014, China

a r t i c l e

i n f o

Article history: Received 13 July 2017 Received in revised form 1 October 2017 Accepted 6 November 2017 Available online 8 November 2017 Keywords: Color change Mechanophore Mechanoresponsive Spiropyran Stress sensing

a b s t r a c t Mechanoresponsive polymers have garnered significant attention in recent years, due to their great potential application in stress/strain sensing and damage warning. Several reviews of mechanochemistry have been published recently. In this review, we mainly focus on the most established mechanophore, spiropyran (SP), from an engineering perspective. We present a workflow for SP mechanochemistry, applications in various polymeric systems and impacting factors, as well as characterization techniques. Current limitations and future research directions are briefly highlighted in the end. This review aims to offer deep insight into polymer mechanochemistry and provide approaches to study other mechanophores using the example of SP mechanochemistry in polymers. © 2017 Elsevier B.V. All rights reserved.

Contents 1. 2.

3.

4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 Spiropyran mechanochemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 2.1. SP mechanophore . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 2.2. SP mechanophore in polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 2.2.1. Initiator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 2.2.2. Monomer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 2.2.3. Cross-linker . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 Applied SP mechanochemistry in polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 3.1. SP-Containing polymers of various states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 3.1.1. Solid polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 3.1.2. Dilute polymer solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 3.1.3. Polymer gels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 3.2. SP-containing polymers of various types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 3.2.1. Polyacrylates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 3.2.2. Polyesters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 3.2.3. Polyolefins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 3.2.4. Polyurethane (PU) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 3.2.5. Poly(dimethylsiloxane) (PDMS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 3.2.6. Polyamide. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .32 3.2.7. Polystyrene (PS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 Impacting factors of SP mechanoactivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 4.1. Impacting factors of mechanoactivation of SP-to-MC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 4.1.1. Operation mode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 4.1.2. Polymer structures and properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 4.2. Impacting factors of reversion of MC-to-SP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 4.2.1. Mobility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 4.2.2. Residual strain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

∗ Corresponding author. E-mail address: [email protected] (S. Zhu). https://doi.org/10.1016/j.progpolymsci.2017.11.001 0079-6700/© 2017 Elsevier B.V. All rights reserved.

M. Li et al. / Progress in Polymer Science 79 (2018) 26–39

5.

6.

27

4.2.3. SP type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 Characterization methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 5.1. RGB analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 5.2. Fluorescence imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 5.3. UV–vis spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 Conclusion and future prospect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

1. Introduction Recent decades have witnessed a growing interest in smart materials that are responsive to external stimuli, such as pH [1], temperature [2], light [3], ultrasound [4], redox agent [5], and voltage [6]. Among all fascinating responsiveness, mechanoresponsive polymers are of particular interest and importance due to their potential applications in stress/strain sensing and damage reporting. The most prevalent approach to design a mechanoresponsive polymer is to incorporate “mechanophore”, which contains mechanically labile bonds that are subjective to change under the influence of exogenous forces [7]. Based on different mechanoresponsiveness, mechanophores fall into several categories, such as mechanoluminescence [8], mechanochromism (optical changes under mechanical loadings) [9], mechanocatalysis [10], damage-healing [11], polymer remodeling [11], small molecule release [12]. So far there have been a variety of reviews covering various mechanochemistry, design principles of mechanoresponsive polymers and mechanoactivation approaches [13–17]. In this review, we mainly focus on spiropyran (SP) mechanochemistry, the most established and intensively studied among all mechanophores. When force is transmitted to the labile spiro C O bond, the colorless SP form is mechanically activated to a colorful merocyanine (MC) form. SP motif is a model mechanophore in fundamental studies of the mechanoactivation behaviors in polymers. Because it is not only easy to be incorporated into polymer backbones or networks, but also capable of changing color and fluorescence upon mechanical loading, which is easy and direct to be detected. Therefore, gaining a deep and comprehensive knowledge of SP mechanochemistry is of great significance in the exploration of mechanochemistry in general. Herein, this review is mainly focused on spiropyran mechanochemistry and its application in polymers. The remainder of this review is divided into five main sections: (1) to introduce spiropyran mechanochemistry; (2) to discuss applications of spiropyran mechanochemistry in polymer systems; (3) to systematically investigate impacting factors of spiropyran mechanoactivity (i.e. forward reaction of SP-to-MC and reversion of MC-to-SP); (4) to summarize characterization approaches; (5) to give a summary and a future perspective in terms of tackling the existing challenges of SP mechanochemistry. 2. Spiropyran mechanochemistry 2.1. SP mechanophore When the C O bond is cleaved by mechanical force, SP is activated to MC [18]. The closed SP is colorless/yellow and nonfluorescent, whereas the open MC form is blue/red/purple and fluorescent, as shown in Fig. 1a. Since the spiro C O is the determining point for SP mechanoactivation, the attachment points on SP molecule should be on the opposing sides of the spiro junction so that the force can be effectively transmitted across the labile

bond to rupture it [9]. Based on this designing principle of active SP mechanophores, three types of SP mechanophores, as shown in Fig. 1b, have been reported and intensively studied in literature so far. Being “super smart”, SP1 and SP2 are sensitive to both UV light and mechanical force. However, it is not always advantageous to be super smart, since either UV irradiation or mechanical activation can result in the ring-opening of SP for outdoor materials. Compared to SP1 and SP2, SP3 exhibit advantages in damage sensing in outdoor materials since it is sensitive to mechanical force but not to UV irradiation due to the absence of electron withdrawing nitro group [19,20]. Moreover, SP3 appears to be more thermally stable (no SP-to-MC transition at high-temperature processing) than SP1 and SP2, which is advantageous in polymer processing [21]. Last but not least, the reversion of MC-to-SP is faster in SP3 than in SP1 and SP2 [20]. Hence, SP1 and SP2 can be used to give longer visual color warning, whereas SP3 is applicable for real-time stress sensing. SP mechanophores are usually obtained by boiling salicylaldehyde (e.g. compound 2, 4, 5) and indole (compound 1, 3) in presence of piperidine or pyridine. Dihydroxyl group-containing SP mechanophores are easy to be prepared and thus used as precursor, which can be further functionalized into various groups, such as ␣-bromoisobutyryl group, methylacryloyl ester group, and bis-alkene group, to be covalently attached to different polymer systems. Fig. 2a–c shows the detailed synthetic route of three types of hydroxyl group-containing SP [9,19,22]. The most critical principle to design and prepare an effective HO-SP-OH mechanophore is to have the hydroxyl groups on the opposing sides of spiro junction, which requires both indole and salicylaldehyde contain hydroxyl group or hydroxyl group precursor.

2.2. SP mechanophore in polymers To incorporate SP mechanophore into various polymer systems via different chemistries and mechanisms, SP molecule has been functionalized with different functional groups, such as hydroxyl group, ␣-bromoisobutyryl group, methylacryloyl ester group, and bis-alkene group, as shown in Fig. 1b. Generally, SP moiety is covalently incorporated into polymers as initiator, cross-linker or monomer.

2.2.1. Initiator Hydroxyl-containing SP was used as initiator for ring opening polymerization (ROP) of caprolactone (CL) [22]. Functionalized with ␣-bromoisobutyryl group, SP was applied as atom-transfer radical-polymerization (ATRP) initiator for various polymer systems, including polyacrylates and block copolymers [9,23,24].

2.2.2. Monomer Diol SP can also be employed as a monomer in step growth polymerization to obtain polyurethane (PU). The step growth polymerization enables controlling mechanophore concentration independently of molecular weight or cross-linking density [25].

28

M. Li et al. / Progress in Polymer Science 79 (2018) 26–39

Fig. 1. (a) Mechanical force-induced ring-opening reaction of spiropyran (SP) to merocyanine (MC); (b) Three types of active SP mechanophores based on the location of attachment. The attachment can be functionalized by hydroxyl group, ␣-bromoisobutyryl group, methylacryloyl ester group, bis-alkene group.

Fig. 2. Synthetic route of (a) HO-SP1-OH [9]; (b) HO-SP2-OH [22]; (c) HO-SP3-OH [19].

M. Li et al. / Progress in Polymer Science 79 (2018) 26–39

29

Table 1 A summary of SP-incorporated polymer systems. Category

Polymers

Examples

References

Polymer States

Solid polymers

PMMA specimen PMA specimen PU specimen PDMS specimen EVA,EOC specimen PCL specimen PS coating P(BA-co-MMA-co-VTES-co-SP) coating PS-Pn BA-SP-Pn BA-PS specimen PMA-SP-PMA solution Pn BA10 -SP-Pn BA10 solution SP-(t-BA88 -b-NIPAM62 )2 microgel dispersion Solvent-swelling PMMA gel CO2 -breathing PDEA microgel Tough PAM hydrogel PMMA specimen and gel PMA solution and specimen PDEA microgel Pn BA solution PS-Pn BA-SP-Pn BA-PS specimen PEA, Pi BA, Pt BA solution P(BA-co-MMA-co-VTES-co-SP) latex coating PS coating PCL film 3D-printed PCL specimen EVA, EOC specimen Unsegmented PU specimen Segmented PU specimen Hydrogen bonding enhanced PU specimen Dual cross-linked PU specimen Healable PU specimen Waterborne PU specimen PDMS film PDMS balloon Electro-mechano-chemically responsive PDMS film PDMS composite film Mechanically strong PAM-co-MA/SP hydrogel

[9,26,28–34] [9,31,35,36] [25,37–42] [27,43–46] [21] [19,22] [47] [48] [23] [49,50] [23] [23] [51] [52] [53] [9,26,28–34,51] [9,20,31,35,36,49,50] [52] [23,50] [23] [50] [48] [47] [22] [19] [21] [25] [37] [38,39] [40] [41] [42] [27,43] [44] [45] [46] [53]

Dilute polymer solutions

Polymer gels

Polymer Types

Polyacrylates

Polystyrene Polyesters Polyolefins Polyurethanes

Poly(dimethylsiloxane)

Polyamide

2.2.3. Cross-linker As a cross-linker for polyacrylates, only a small amount of methylacryloyl ester-containing SP is needed as a stress/strain indicator, with ethylene glycol dimethacrylate (EGDMA) and poly (ethylene glycol) dimethacrylate (PEG) are usually added as primary cross-linkers. It is not only because SP mechanophore cross-linker is more complicated to synthesize than the commercially available EGDMA or PEG, but also because only a small amount is sufficient for obvious visual color changing [9,26]. Methylacryloyl ester-containing SP can also be used as co-cross-linker for polyolefins during curing [21]. SP functionalized with bis-alkene group can serve as cross-linker for poly(dimethylsiloxane) (PDMS) system [27]. 3. Applied SP mechanochemistry in polymers SP mechanophore has been incorporated into various polymer systems, shown in Table 1. Based on the states of polymer, SP mechanophore has been embedded into bulk polymers, polymer solutions, and polymer gels. Based on polymer chemistry, various types of polymers have been chosen for SP mechanophore incorporation, including polyacrylates, polyester, polyolefins, PU, PDMS, polyamide and polystyrene. 3.1. SP-Containing polymers of various states 3.1.1. Solid polymers Mechanoactivation of SP in solid polymeric materials was pioneered by Moore et al. in 2009. SP was covalently incorporated into the backbone of poly(methyl methacrylate) (PMMA) and

poly(methacrylate) (PMA) as initiator via ATRP and also into lightly cross-linked PMMA and PMA as a secondary cross-linker via free radical polymerization (FRP). Mechanoactivation of SP in PMA and PMMA bulk polymers was demonstrated via tensile tension and compression, as shown in Fig. 3a [9]. However, the mechanoactivation in the glassy PMMA under tensile tension was difficult to achieve. The mechanoactivation in glassy polymers requires large deformation, following the yield point. The glassy polymers usually fail at the low strain before mechanoactivation occurs. Alternative activation methods are called for to apply higher stress to the material for activation before failure. Moore’s group later demonstrated that SP-containing PMMA could be mechanically activated by shearing (monotonic torsion) (Fig. 3b) and that SPlinked polystyrene (PS) thin film could be activated by acoustic sound waves (Fig. 3c) [26,47].

3.1.2. Dilute polymer solutions Ultrasonication induced elongation flow can stretch polymer chains and transduce mechanical forces through acoustic cavitation, which describes the sonication induced nucleation, growth and collapse of bubbles in liquid. The collapse of bubbles leads to solvodynamic shear forces on the polymer chains, because the chain ends are pulled to the voids created by the imploding bubbles [54–57]. The forces developed on polymer chains can contribute to mechanoactivation of SP moiety in the backbone. Polymer chains need to be sufficient long for the mechanical force to be focused onto the SP moiety at center to be mechanically activated,[23,49,50] which will be introduced in details in Section 4. Recently, Du’s group revealed that mechanoactivation of SP was significantly enhanced in micelles compared to the dissolved

30

M. Li et al. / Progress in Polymer Science 79 (2018) 26–39

Fig. 3. Mechanical force-induced activation of SP in solid polymers. (a) Mechanoactivation of SP in bulk PMA specimens under stretching and bulk PMMA beads under compression. [9], Copyright 2009, reproduced with the permission from the Nature Publishing Group; (b) Mechanoactivation of SP in lightly cross-linked PMMA via torsion and characterized by fluorescence imaging. [26], Copyright 2011, reproduced with the permission from the Royal Society of Chemistry; (c) Mechanoactivation of SP-linked PS films subjected to laser-generated high amplitude acoustic pulses. [47], Copyright 2014, reproduced with the permission from the American Chemical Society.

Fig. 4. Mechanoactivation of SP-(t-BA88 -b-NIPAM62 )2 was enhanced within micelles in THF/H2 O mixed solvent compared to those dissolved in THF under sonication. [24], Copyright 2016, reproduced with the permission from the American Chemical Society.

counterparts. As the self-assembly of poly(tert-butyl acrylateb-N-isopropylacrylamide) with SP molecule at the center of (SP-(t-BA88 -b-NIPAM62 )2 ) in tetrahydrofuran (THF)/water mixture, the micelles exhibited about 5-fold of those dissolved in THF under sonication, shown in Fig. 4 [24]. It was attributed to a synergetic effect of the entangled and swollen polymer chains in the micellar cores, and an improved dielectric constant of the local environment of SP moiety in the micelles.

3.1.3. Polymer gels Recently, there has been a new trend in the research of SP mechanoactivation, which is to mechanically activate SP in polymer gels. Swelling force can be used to stimulate the SP mechanoactivation in polymer gels. As is shown in Fig. 5a, Moore’s group initially introduced the concept of solvent-swelling induced mechanoactivation of SP in lightly cross-linked PMMA and demonstrated that swelling force was sufficient to drive the ring-opening reaction of SP-to-MC by excluding the solvent

M. Li et al. / Progress in Polymer Science 79 (2018) 26–39

31

Fig. 5. SP mechanoactivation within polymer gels. (a) Solvent swelling induced mechanoactivation of SP-cross-linked PMMA gel. [51], Copyright 2014, reproduced with the permission from the American Chemical Society; (b) CO2 -breathing induced mechanoactivation of SP within microgel. [52], Copyright 2016, reproduced with the permission from John Wiley and Sons Inc; (c) Stretching/Compression induced activation of SP within stretchable hydrogel. [53], Copyright 2017, reproduced with the permission from John Wiley and Sons Inc.

effect [51]. Later, Zhu’s group prepared CO2 -switchable poly(2(diethylamino)ethyl-methacrylate) (PDEA) microgels with SP as cross-linker via soap-free emulsion polymerization and demonstrated CO2 -breathing induced SP mechanoactivation [52]. After treatment of CO2 , the microgels became swollen due to protonation of the tertiary amine group of PDEA (Fig. 5b). The swelling force further drove the ring-opening reaction of SP-to-MC in microgels. The emerging purple coloration and fluorescence of the microgel dispersion after CO2 treatment indicated the activation of MC. Very recently, Zheng’s group demonstrated that mechanoactivation of SP in polymer hydrogels could also be achieved directly via tension or compression [53]. A novel poly(acrylamideco-methyl acrylate/spiropyran) (P(AM-co-MA/SP)) hydrogel with

excellent mechanical properties was developed via micellarcopolymerization method. Mechanical force-induced color change was demonstrated under a high tensile stress level (tensile stress of 1.45 MPa at the strain of ca. 600%), as shown in Fig. 5c.

3.2. SP-containing polymers of various types 3.2.1. Polyacrylates SP mechanophore was initially incorporated into PMMA and PMA as cross-linker or initiator by Moore et al. [9] as shown in Fig. 3. Thereafter, intensive fundamental studies on SP mechanoactivation behaviors have been carried out based on SP-containing polyacrylates, including PMA, PMMA, poly(ethyl acrylate) (PEA),

32

M. Li et al. / Progress in Polymer Science 79 (2018) 26–39

Fig. 6. Mechanochemically active acrylic latex coating. [48], Copyright 2017, adapted with the permission from the American Chemical Society.

poly(n-butyl acrylate) (PnBA), poly(iso-butyl acrylate) (PiBA), poly(tert-butyl acrylate) (PtBA), and PDEA, which were prepared via FRP or ATRP [20,23,26,28–32,35,36,47,50,52]. The studies not only enriched fundamental knowledge of mechanochemistry, but also provided valuable information and experience for potentially broadening industrial applications of mechano-responsive polymers. To push one step closer to the real applications, Zhu et al. [48] incorporated SP mechanophore into acrylic latex coating and demonstrated mechanoactivation of latex thin films. Acrylic latexes were prepared using BA and MMA as monomer, vinyltriethoxysilane (VTES) as interparticle cross-linker and SP as intraparticle cross-linker via emulsion polymerization and then casted on to Teflon coated surface to give latex coatings (Fig. 6). The mechanochromic latex coatings provided good insight into a variety of acrylic latex-based applications of mechanochemistry in adhesives, protective coatings, and paints.

3.2.2. Polyesters As a ductile polyester, polycaprolactone (PCL) has been selected for incorporating SP mechanophore. As is shown in Fig. 7a, McElhanon et al. [22] synthesized SP-containing PCL with diolcontaining SP as initiator via ROP. The mechanoactivation of SP in PCL during straining was illustrated in Fig. 7b. Boydston’s group further advanced the method of preparing SP-containing PCL using 3D printing technology. The computer aided design (CAD) model and the 3D-printed mechanically active PCL are shown in Fig. 7c–e [19]. Exploiting 3D-printing technology provided the possibility of rapid production of mechanically responsive materials, which is of great importance for broadening mechanochemistry into potential commercial polymers.

3.2.3. Polyolefins As the most commonly and widely used polymer type, polyolefins have more than half of the total polymer consumption in the world. The past decades have witnessed the expanding application of polyolefins in medical, packaging, sports, cable, wire coatings, etc. Lately, Zhu’ group demonstrated the mechanoactivation of SP in polyolefins [21]. As is shown in Fig. 8, SP was covalently crosslinked into two most widely used commercial polymers, polar poly(ethylene-vinyl acetate) (EVA) and non-polar poly(ethylene octene) (EOC) via free radical mechanism during the curing process, which was carried out under hot press at 160 ◦ C for one hour. The example of color-changing polyolefins in response to mechanical force provided meaningful insights into expanding mechanochemistry into the commercial commodity polymers.

3.2.4. Polyurethane (PU) As a ubiquitous engineering polymer, PU exhibits excellent properties, including high abrasion, chemical resistance, and advanced mechanical properties. To help PU to feel the force through color changing, Moore et al. [25] synthesized SP-linked PU via step growth polymerization method and demonstrated mechanoactivation of SP-linked PU, as shown in Fig. 9a. The segmented PU is of particular interest since its elasticity and strength can be well tuned by changing the ratio of hard and soft segments. A typical PU elastomer consists of a polyol, a diisocyanate and a chain extender. The hard segment contains isocyanate and chain extender, and soft segment contains long chain polyols. The advanced engineering properties and adjustable mechanical properties inspired the researchers to incorporate SP mechanophore into segmented PU to study the mechanoactivation behavior in different phases. Moore et al. [37] reported a segmented polyurethane with SP mechanophore in either soft or hard segments (Fig. 9b). To ensure effective mechanoactivation of SP, polymer candidates should possess excellent load-bearing capacity, resilience and defect tolerance. Weng’s group have been dedicated to improving the mechanical properties of polymers for better mechanoactivation of SP and have developed several thermoplastic elastomers containing SP moiety. They developed a novel PU containing both covalent SP mechanophore and quadruple hydrogen bonding 2-ureido-4-pyrimidone (UPy) (Fig. 9c) [39]. The effective SP mechanoactivation was due to the superb mechanical properties of the polymer, which was attributed to the fragmentation of hard domains and dissociation of UPy stacks and dimers to dissipate energy during stretching. Later, they also found incorporating UPy at chain ends gave enhanced mechanoactivation of SP, compared to those without UPy [38]. The phenomenon was attributed to the supramolecular interactions, which could promote chain orientation and strain induced crystallization. Fig. 9d illustrates the mechanoactivation of SP in a PU elastomer doubly cross-linked by oligo-poly(lactic acid)-capped pentaerythritol (PTT-PLA) as covalent cross-links and UPy as physical cross-links [40]. Weng’s group also developed Zn2+ -containing polyurethane with SP mechanophore on the backbone, which exhibited damage sensing and self-healing properties (Fig. 9e) [41]. The development of polymers with strong, tough, elastic, self-healing and stress/strain sensing properties is of great importance for fundamental studies and potential applications. Moreover, developing environmentally friendly polymers integrated with damage sensing feature is also one of the most important directions of mechanochemistry. Lately, SP mechanophore was incorporated in waterborne PU, the mechanoactivation of which was investigated by Jia et al. [42] 3.2.5. Poly(dimethylsiloxane) (PDMS) The mechanoactivation of SP in elastomers requires large deformation, which is mostly not recoverable. Craig et al. [27] reported the covalent incorporation of bis-alkene functionalized SP into the network of PDMS as a cross-linker and characterized the mechanoactivation of SP in PDMS with full shape recovery and high reversibility (Fig. 10a). The PDMS with repeatable mechanoactivation and full shape recovery renders its possibility in various applications, such as writing panels, soft robots, fluorescent patterning and touch screens (Fig. 10b–c) [27,43–46]. 3.2.6. Polyamide The mechanoactivation of SP in polyacrylamide (PAM) is lately achieved by introducing SP-cross-linked PMA microspheres [53], shown in Fig. 5c. During gelation process, PAM was connected to SP/PMA microspheres via coupling termination and/or copolymerization of AM and MA. Due to the covalent bonding between PAM and PMA, the macroscopic force on the PAM

M. Li et al. / Progress in Polymer Science 79 (2018) 26–39

33

Fig. 7. Mechanochemically active PCL. (a) Chemical structure of SP-linked PCL. [22]; (b) Mechanoactivation of SP in PCL under stretching; [22], Copyright 2010, reproduced with the permission from the American Chemical Society; (c-e) CAD model and 3D-printed mechanochemically active PCL. [19], Copyright 2014, reproduced with the permission from the American Chemical Society.

Fig. 8. (a) Schematic illustration of processing mechanoresponsive polyolefin and mechanoactivation under the tensile test; (b) Chemical structure of cross-linked EVA and EOC with SP covalently incorporated. [21], Copyright 2017, reproduced with the permission from Elsevier Ltd.

hydrogel was transmitted into microspheres and eventually to SP molecule for mechanoactivation. Surprisingly, both this work and the mechanoactivation of SP in acrylic latex coating reported by Zhu’s group (introduced in 3.2.1) revealed a novel design principle, transmission of macroscopic force from outside to inside of SP-containing microgels, which was covalently connected by polymer chains, to mechanically ring open SP to MC. We believe that SP-containing soft microgels can be covalently introduced into a variety of polymers and potentially expand the applications of mechanochemistry.

under laser-generated acoustic pulses, shown in Fig. 3c [47]. The mechanoactivation of the bulk SP-linked PS could not be achieved under quasi-static loading because the specimen already failed at stresses of ca. 60 MPa before SP was mechanoactivated. However, the thin film of SP-linked PS could accommodate plastic strains before failure and thus be mechanically activated at the high strain rates (ca. 1 × 107 to 1 × 108 s−1 ) induced by the acoustic pulses. The mechanoactivation of mechanophore under high strain rates and high energy can be applied in some extreme situations. 4. Impacting factors of SP mechanoactivity

3.2.7. Polystyrene (PS) Moore et al. prepared mechanically active SP-linked PS via ATRP method and demonstrated the mechanoactivation of SP-linked PS

Intensive studies on SP mechanoactivity have been carried out based on SP-containing polymers. In general, most studies focused

34

M. Li et al. / Progress in Polymer Science 79 (2018) 26–39

Fig. 9. Mechanochemically active (a) unsegmented PU. [25], Copyright 2010, reproduced with the permission from the American Chemical Society. (b) segmented PU with SP embedded in either soft segments or hard segments. [37], Copyright 2013, reproduced with the permission from the American Chemical Society; (c) PU with quadruple hydrogen bonding UPy moieties. [39], Copyright 2013, reproduced with the permission from the American Chemical Society; (d) dual cross-linked PU elastomer (chemical and physical cross-links). [40], Copyright 2014, reproduced with the permission from the American Chemical Society; (e) Healable Zn2+ -containing PU. [41], Copyright 2013, reproduced with the permission from the American Chemical Society.

Fig. 10. Mechanochemically active PDMS. (a) Active PDMS network obtained by incorporating bis-alkene functionalized SP moiety through platinum-catalyzed hydrosilylation, and the mechanoactivation of SP in PDMS with full and repeatable shape recovery. [27], Copyright 2014, reproduced with the permission from the American Chemical Society; (b) Electro-mechano-chemically responsive PDMS for on-demand fluorescent patterning. [45], Copyright 2014, reproduced with the permission from the Nature Publishing Group; (c) Force-sensitive mechanochromic touch screen. [46], Copyright 2017, reproduced with the permission from the American Chemical Society.

M. Li et al. / Progress in Polymer Science 79 (2018) 26–39

on the forward reaction of SP-to-MC induced by mechanical loadings and discussed the effect of operating mode (stress/strain, deformation rate, stress relaxation) and polymer structures (chain length, chain mobility, the location of SP in blocks, linking architecture, chain orientation) on SP mechanoactivity. Moreover, the reversion process of MC-to-SP after mechanoactivation has also been investigated in a few literatures. The reversion of MC-to-SP is mainly influenced by chain mobility, residual strain, and SP type. 4.1. Impacting factors of mechanoactivation of SP-to-MC 4.1.1. Operation mode 4.1.1.1. Stress/strain. Moore et al. [9] predicted of the selective rupture of spiro C O bond in SP mechanophore using a modeling method called “Constrained Geometries simulate External Force” (COGEF) and found that applying force alters the potential energy surface through lowering activation energy barrier to rupture the C O bond and convert SP to MC. They further confirmed that applying mechanical stress not only lowers activation energy for the forward reaction (SP-to-MC), but also increases the barrier for the reverse reaction (MC-to-SP) [58]. Later, Craig et al. [59] studied the force-rate relationship of two SP isomers via single molecule force spectroscopy studies and estimated the threshold stress for SP mechanoactivation on a molecular scale. The threshold stress was approximately 240 pN for ring opening SP1 and 260 pN for SP2 on the time scale of tens of milliseconds, respectively. McElhanon et al. [22] also suggested C O rupture occurred at a molecular elongation of more than 39% via density functional theory (DFT) and time-dependent DFT calculations. However, it is difficult to monitor the force felt by SP molecules in the local environment or at a molecular level under the experimental conditions. Sottos et al. [30,35] simulated the local force felt by SP within glassy and rubber polymers via multiscale computational analysis and concluded that microscale force transmission could be interpreted by macroscale stress response. These modeling and theoretical studies laid significant foundations for studying SP mechanoactivation simply using macroscopic stress. On a macroscopic level of force, when the force is transmitted to C O bond through polymer chains, SP mechanoactivation starts to occur. There is a threshold stress of SP mechanoactivation, which is the minimum stress to mechanically activate SP in bulk polymers. In literature, the onset of detectable SP activation is usually determined as the moment, at which the activation is significant enough to counteract with the thinning effect [36]. Therefore, critical stress and critical strain for mechanoactivation are defined as the stress and strain when the onset of detectable SP occurs. Take mechanoactivation of SP-cross-linked PMMA at 90 ◦ C as an example, the SP-to-MC conversion is minimal in the elastic regime and significantly increases after ca. 10% of strain [34]. Therefore, the critical strain for SP mechanoactivation in this polymer system is estimated to be 10%. 4.1.1.2. Deformation rate. The deformation rate has an important effect on the mechanoactivation of SP in three aspects, including SPto-MC conversion, critical stress for mechanoactivation and time required for initiating activation. (1) SP-to-MC conversion: Higher strain rate leads to higher macroscopic stress, causing greater chain alignment and more force applied across mechanophore. Therefore, greater fluorescence intensity (higher SP-to-MC conversion) was observed for the SP-linked linear PMA under higher deformation rates [36]. (2) Critical stress: Monotonic torsion experiment of SP-cross-linked PMMA revealed that increasing strain rate increases critical stress for SP to be mechanoactivated by increasing yield stress [26]. The compression test of SP-cross-linked PMMA beads revealed that the threshold stress and yield stress increases linearly with natural log of the strain rate, while the

35

threshold strain remains almost constant [9]. (3) Time required for mechanoactivation: Torsional creep experiment of SP-cross-linked PMMA revealed that higher stress level led to higher creep rates, resulting in higher force applied through polymer chains and thus shorter time needed to mechanoactivate SP. In the torsional creep and monotonic torsion, the time required for mechanoactivation t versus the shearing rate r˙ is fitted into a power law relationship, t = 0.06 × r˙ −0.73 [32]. Moreover, when strain rate is increased to a very high level, SP can even be mechanically activated at low strains in glassy polymers. SP could not be mechanically activated in the uniaxial tension test of SP-linked polystyrene (PS) prior to failure (failure stress ca. 60 MPa), while SP mechanoactivation of PS thin film was achieved under laser-generated and high amplitude acoustic pulses, which had a high strain rate (ca. 1 × 107 to 1 × 108 s−1 ) [47]. The acoustic pulse-generated shockwave mechanically activated SP in PS thin film at a high level of stress but small strains (critical stress ca. 180 MPa). However, high strain rates can lead to an ineligible thermal effect in the glassy polymer. Hooper et al. [33] revealed that thermal effects have a large synergetic effect on the mechanoactivation of SP at high strain rates in solid PMMA. Therefore, the mechanoactivation of SP in glassy polymers under high strain rates cannot be completely distinguished from thermal effects. However, the time-dependent feature of SP-to-MC reaction is another factor that influences SP mechanoactivation. Silberstein et al. [34] reported a contradictory result with the previously stated evidence. Higher strain rates led to less activation of SP in the uniaxial tension test of SP-cross-linked PMMA at 90 ◦ C. The counterintuitive result was attributed to the time-dependent feature of SP-to-MC reaction feature. Even though lower strain rates result in lower stresses, slower strain rates render more time for SP to convert to MC. The similar trend was also observed in SP-cross-linked polyolefins and acrylic latex coatings [21,48]. 4.1.1.3. Stress relaxation. Stress relaxation was mainly affected by stress across mechanophore, time-dependent feature of the SP-toMC reaction and type of SP mechanophore. (1) Time-dependent feature: A stress relaxation experiment of SP-linked PU was performed by Moore et al. [25] They found that the sample turned increasingly purple when it was held under a constant strain and reached steady-state after 1–2 h. No reversion of MC-to-SP was observed for the SP-linked PU during stress relaxation. Visible light was also applied on the sample when it was held under constant strain. Due to the visible light, MC was reverted to SP, which could be recovered upon removal of the light. All the evidence indicated that SP reached mechanically biased equilibrium when held at constant strain, suggesting that SP-to-MC conversion is a timedependent process. (2) Stress-dependent feature: During stress relaxation experiment, however, the stress reduction inevitably affects the SP-to-MC conversion. When the stress is too low, the equilibrium of SP-to-MC conversion is no longer favorable. For the SP-linked PMA, the SP-toMC conversion continued to proceed until the stress was below ca. 10 MPa [36]. Therefore, the time-dependent feature only occurs, provided the stress remains high enough. Stress-relaxing experiment of SP-linked PU also confirmed the existence of threshold stress. In the stress relaxing of PU, the SP-to-MC transition rate constants increased linearly with the applied stress, and MC-to-SP reversion rate constants decreased linearly with the stress, when stresses were above the threshold level [58]. (3) SP type: The type of SP also makes a difference in the SPMC equilibrium during stress relaxation. Zhu et al. [21] reported fast reversion of SP3-cross-linked polyolefin in the stress relaxation experiment. The plot of the blue intensity vs. time almost overlapped the plot of stress vs. time. Upon held at constant strain, the stress dropped sharply below ca. 9 MPa, probably accounting

36

M. Li et al. / Progress in Polymer Science 79 (2018) 26–39

for the decay of blue coloration. It was also believed that fast reversion might also result from the unique molecular structure of SP3, lacking electron-withdrawing nitro group. Just as the stress-relaxation has an effect on the mechanical equilibrium of SP and MC, the ring-opening reaction of SP-to-MC also influences the stress relaxation of SP-containing polymer in turn. McElhanon et al. [22] found UV irradiation led to a decrease of 50 kPa in stress of SP-containing PCL, which previously reached a flat baseline of ∼5.2 MPa in stress under constant 1% of strain (not mechanically activated yet). The light-induced stress relaxation was attributed to lengthening polymer due to the cleavage of C O bond and ring-opening of SP-to-MC. 4.1.2. Polymer structures and properties 4.1.2.1. Chain length. Polymer chain length plays a crucial role in the ultrasound-induced SP mechanoactivation. Polymer chain length determines whether the ultrasound-induced SP mechanoactivation can occur. Similar to the evidence of other mechanophores activated by sonication reported by Moore, Bielawski and Craig et al. [49,60–62] Weng et al. [23] found there is a threshold molecular weight for effective mechanoactivation using sonication. SP-linked PnBA was mechanically activated under sonication when MW > 14 kD, and could not be activated when MW < 14 kD. However, which is a better descriptor for mechanical force transduction, molecular weight or polymerization degree? Moore et al. [50] measured the rate constant for the sonication induced SP-to-MC conversion in various polymers, including PMA, PEA, and PBA. For the sonication  induced  mechanoactivation of SP follows the equation, At = B 1 − e−kt , where At is the absorbance at time t, B is the amplitude and k is the rate constant. By plotting k as a function of polymerization degree, they revealed that PMA, PEA and PBA all collapsed into a single linear regression, whereas all different linear regressions when k was plotted against MW . The quantitative evidence suggested that polymerization degree is a descriptor of mechanochemical transduction and a determining factor for the mechanoactivity rate, instead of molecular weight. 4.1.2.2. Chain mobility. Sufficient chain mobility is required for effective mechanoactivation of SP. Activation of SP in glassy PMMA (Tg = 127 ◦ C) could be achieved at a temperature range of 90–105 ◦ C or with plasticizer added [28]. At higher temperature than 105 ◦ C, no SP activation was observed because the viscous flow-dominated deformation resulted in lower stress applied on SP moiety. At lower temperature than 90 ◦ C, no mechanoactivation occurred due to insufficient plastic flow. The results suggested that plastic flow is necessary for SP mechanoactivation. Rubber toughening is another alternative to increasing plastic deformation to afford SP mechanoactivation in PMMA [29]. Shearing test of cross-linked PMMA also showed that activation of SP occurred after yielding and large plastic flow, suggesting the significance of molecular motion for activation [26]. Torsional creep deformation of SP-cross-linked PMMA coupled with full field fluorescence further confirmed that SP mechanoactivation needs large-scale deformation [32]. It has been concluded the initiation of SP mechanoactivation in glassy polymer occurs after the yield peak even though the stress is much higher than the critical stress to mechanically activate in elastomers [34]. 4.1.2.3. Location of SP in blocks. Moore et al. [37] utilized SP mechanophore as a molecular probe of force and orientation by incorporating SP moiety into either soft (SPSS) or hard phase (SPHS) of segmented PU at a low (22 wt%) and a high (40 wt%) content of hard segment. For the PU containing 22 wt% hard segment, the hard and soft segments were miscible, and thus the load was evenly distributed across the hard and soft segments. The same orientation parameters of SP in SPHS and SPSS at 22 wt% hard segment were

attributed to the high miscibility of hard and soft segments. However, the phase separation was evident for the PU containing 40 wt% hard segment and SPSS showed higher orientation degree than SPHS due to the higher strain across the soft segments, suggesting that each phase possessed different ability to feel the force. 4.1.2.4. Linking architecture. The mechanoactivation behavior of SP also depends on the linking architecture. The mechanoactivation of linear PMMA with SP in the center occurred at much lower strains (strain 20%) [28]. Moreover, the relative length of primary cross-linker to SP cross-linker also has an effect on SP mechanoactivation behavior. Moore et al. [26] synthesized SP-cross-linked PMMA with three different primary cross-linkers of various length, which were PMMA-EGDMA, PMMA-PEG550, PMMA-PEG750. It was revealed that the critical stress decreased with longer primary cross-linkers than SP but no further decrease at further length via torsion test. The threshold strain remained almost constant with varying length of primary cross-linker at low shear rate (0.001 rad s−1 ), whereas increased with the increasing length of primary cross-linker at high shear rate (0.1 rad s−1 ). 4.1.2.5. Chain-orientation. The orientation of SP moiety is of great importance for mechanical activation. The degree of orientation of SP mechanophore-containing linear PMA and PMMA was characterized by an order parameter. When the pre-stretched ratio of SP-embedded PMA was increased, the UV-activated MC gave higher order parameter up to an order of parameter of ca. 0.5 [31]. The orientation degree of SP was also higher when activated by mechanical force than UV irradiation. The quantitative results indicated that mechanophores preferentially aligned along the tensile direction. The high orientation degree of SP was also observed in SP-linked PMMA, which was tested under 90 ◦ C and methanol-plasticized, respectively. It was found that the order parameter dropped below 0.1 near failure surface after failure, probably attributed to polymer recoil and relaxation. Therefore, polymer chain alignment plays a significant role in SP mechanoactivation. Using birefringence as an indication of polymer chain alignment, Moore et al. [36] discovered rapid mechanoactivation of SP started when polymer chains aligned along stretching direction. It was also proved that pre-stretching SP-containing polymer before tests improved chain alignment and thus increased mechanoactivation degree [29]. 4.2. Impacting factors of reversion of MC-to-SP After the removal of the load, the mechanically activated MC form can be finally reverted to the SP form. The reversion process is a “double-edged sword” for different engineering purposes. On the one hand, we expect SP-containing polymers to have slow reversion so that the persistent coloration will give us warning signals to prevent catastrophic failure from occurring. On the other, the fast recovery of MC-to-SP enables repeatable laboratory experiments within a short period of time and real-time stress sensing purpose. In general, the MC-to-SP reversion mainly depends on the mobility of the local environment, residual strain and SP types. 4.2.1. Mobility Moore et al. [9] revealed the reversion of MC-to-SP can be affected significantly by the mobility of local environment, reflected by Tg . The reversion of MC-to-SP in elastomeric PMA (Tg = 12 ◦ C) needed 6 h under fluorescent room light, whereas several weeks for glassy PMMA (Tg = 128 ◦ C). For the SP-linked PU (Tg ≈ −60 ◦ C) almost disappeared after 1 h in the dark [25]. SP reverts to MC slower in the dark than in the visible light. So it can be concluded

M. Li et al. / Progress in Polymer Science 79 (2018) 26–39

that the lower the polymer Tg , the higher mobility the polymer chains, and thus faster reversion of MC-to-SP. 4.2.2. Residual strain A comparison of UV-activated and mechanically activated SPlinked PU was carried out by Moore et al. [25] The plot of the fluorescence intensity as a function of time was fitted into a sing-exponential decay with a time constant of  ≈ 30 min for the UV-activated sample and  ≈ 38 min for the mechanically activated sample. The slower reversion of mechanoactivated SP in PU was probably attributed to the slow relaxing residual strain remained in the polymer. 4.2.3. SP type Boydston et al. [19] initially discovered the purple coloration of mechanoactivated SP3-linked PCL disappeared within approximately 30 s after the removal of the load. Then Zhu et al. [20] further reported that SP3 mechanophore showed a faster reversion than the other two types of SP mechanophore, probably related to the absence of nitro group on 6-position of benzopyran in SP3. The reversion needed ca. 12 min for the SP3-cross-linked PMA to revert to MC, whereas ca. 300 min for the SP2-cross-linked PMA. The plot of the change of green color intensity vs. time was fitted into a sing-exponential decay with a time constant of  ≈ 4.7 min for SP3-cross-linked PMA and  ≈ 104 min for SP2-cross-linked PMA. However, a deep understanding of reversion speed difference between various SP mechanophores must be further explored. Another noticeable difference on the behavior of color changing upon unloading was also related to SP type. Several studies reported a secondary color change upon relaxation for SP1- and SP2containing polymers, identified as a transient state [9,22,25,39,41]. The blue/purple transition upon relaxation was probably related to the isomerization about methine bridge [27]. However, no color transition was observed for SP3-containing polymer upon relaxation. SP3-cross-linked PMA turned purple during stretching until failure and persisted after failure [20]. The lack of color transition might also be related to the absence of electron-withdrawing NO2 group. 5. Characterization methods The SP is colorless/yellow and non-fluorescent, whereas the activated form, MC state, is purple/blue/red and fluorescent, which can be characterized by RGB analysis, fluorescence imaging and UV–vis spectroscopy. Each method is introduced as follows. 5.1. RGB analysis The emergence of color is recorded by taking optical images of the SP-containing polymer specimen during stretching or compression experiments, and the optical images are analyzed to obtain the average intensity in Red (R), Green (G) and Blue (B) channel at the region of interest. The RGB ratio can be calculated as rRC = R/(R + G + B), rGC = G/(R + G + B) and rBC = B/(R + G + B), based on Grassmann’s law [63]. Mechanoactivation process is directly illustrated via the plot of RGB ratio against strain or stress. As an easy and direct approach to characterize SP mechanoactivation, the RGB analysis does not require customized facilities and thus has been widely used in a lot of studies [9,20,21,48,52]. 5.2. Fluorescence imaging In situ fluorescence imaging is achieved by exciting a mechanically activated SP-containing polymer by a 532 nm laser. The resulting fluorescence intensity is recorded and later plotted as

37

a function of stress/strain for mechanoactivation analysis. Compared with RGB analysis, fluorescence imaging appears to be more accurate and utilized extensively in studies [25,28,32,47,58]. Furthermore, fluorescence intensity at various order parameters has been used to investigate the chain orientation feature of SP mechanoactivation [31,37].

5.3. UV–vis spectroscopy UV–vis spectroscopy has been used to study the mechanoactivation of SP-to-MC in bulk polymers via the plot of the intensity at the maximum absorption wavelength (∼570 nm) against stress/strain [25]. This method has also been employed in sonication induced mechanoactivation of SP in polymer solution. Real-time UV–vis spectroscopy with flow-cell design was used to synchronously quantify the SP-to-MC conversion during sonication of polymers containing SP mechanophore [50].

6. Conclusion and future prospect This review has given a comprehensive scope of SP mechanochemistry and its applications in polymers. On the one hand, the study of SP mechanochromism provides fundamental knowledge of mechanochemistry, which is of great significance in developing more mechanophores with the desired mechanoresponsiveness, such as stress induced self-strengthening [64–67], which converts destructive force into constructive functions. On the other hand, the investigation of SP mechanoactivation also gives deep insight and useful guidance in industrial applications in stress/strain sensing and damage warning using SP as a force probe. However, there are still some challenges to be tackled regarding applications. First, the conversion of SP-to-MC under mechanical loading is very limited and reported to be only 20–30% relative to UV-activated [25]. Strength limitation of the polymer itself and insufficient chain orientation might contribute to the low degree of activation. More studies on developing polymers with superior mechanical properties (i.e. strong, tough and elastic) and prestraining are called for to improve SP mechanoactivation yield. Second, effective SP mechanoactivation usually occurs at large deformation. This largely impedes its application in constructional components, which requires instant color-changing at very low strains. Improving chain mobility and pre-straining to align mechanophores can be employed to achieve activation at low strains. More related studies are needed to enable acute mechanoactivation at very low strains. Moreover, mechanoactivation of SP is still limited to one phase so far. It is desired to investigate mechanoactivation behavior of SP at interfaces of organic/inorganic and organic/organic. To be anchored at the interface, SP molecule should contain two different functional groups. There have been some studies on mechanoactivation of other mechanophores at interfaces [68,69]. Employing SP mechanophore at interfaces can be used to monitor force distribution of composite materials via visual color changing, which is a more direct and easier method for stress detection. Last but not least, SP mechanoactivation in polymer network shows a discrepancy in the results obtained from simulation and experiments, which indicates the inhomogeneity of the polymer network in the experiment [30,35]. It is worthwhile to study how homogeneity influences SP mechanoactivation, which can not only assists in the understanding of stress distribution in polymer networks, but also provides another approach to study the homogeneity of polymer networks.

38

M. Li et al. / Progress in Polymer Science 79 (2018) 26–39

Acknowledgements The authors sincerely acknowledge the Natural Science and Engineering Research Council (NSERC) (RGPIN-2015-05841) of Canada for supporting this fundamental research through Discovery Grant program. We also thank the Canada Foundation for Innovation (CFI) (200154) for the equipment and facilities. SZ thanks the Canada Research Chair (CRC) (950-229035) program for supporting his research.

References [1] Dai S, Ravi P, Tam KC. pH-Responsive polymers: synthesis, properties and applications. Soft Matter 2008;4:435–9. [2] Lutz J-F, Akdemir Ö, Hoth A. Point by point comparison of two thermosensitive polymers exhibiting a similar LCST: is the age of poly(NIPAM) over. J Am Chem Soc 2006;128:13046–7. [3] Zhao Y. Light-responsive block copolymer micelles. Macromolecules 2012;45:3647–57. [4] Chen W, Du J. Ultrasound and pH dually responsive polymer vesicles for anticancer drug delivery. Sci Rep 2013;3:2162–70. [5] Napoli A, Valentini M, Tirelli N, Müller M, Hubbell JA. Oxidation-responsive polymeric vesicles. Nat Mater 2004;3:183–9. [6] Peng L, Feng A, Zhang H, Wang H, Jian C, Liu B, et al. Voltage-responsive micelles based on the assembly of two biocompatible homopolymers. Polym Chem 2014;5:1751–9. [7] Caruso MM, Davis DA, Shen Q, Odom SA, Sottos NR, White SR, et al. Mechanically-induced chemical changes in polymeric materials. Chem Rev 2009;109:5755–98. [8] Chen Y, Spiering JH, Karthikeyan S, Peters GWM, Meijer EW, Sijbesma RP. Mechanically induced chemiluminescence from polymers incorporating a 1,2-dioxetane unit in the main chain. Nat Chem 2012;4:559–62. [9] Davis DA, Hamilton A, Yang J, Cremar LD, Van Gough D, Potisek SL, et al. Force-induced activation of covalent bonds in mechanoresponsive polymeric materials. Nature 2009;459:68–72. [10] Jakobs RTM, Ma S, Sijbesma RP. Mechanocatalytic polymerization and cross-linking in a polymeric matrix. ACS Macro Lett 2013;2:613–6. [11] Lenhardt JM, Black AL, Craig SL. Gem-dichlorocyclopropanes as abundant and efficient mechanophores in polybutadiene copolymers under mechanical stress. J Am Chem Soc 2009;131:10818–9. [12] Diesendruck CE, Steinberg BD, Sugai N, Silberstein MN, Sottos NR, White SR, et al. Proton-coupled mechanochemical transduction: a mechanogenerated acid. J Am Chem Soc 2012;134:12446–9. [13] Wiggins KM, Brantley JN, Bielawski CW. Methods for activating and characterizing mechanically responsive polymers. Chem Soc Rev 2013;42:7130–47. [14] Calvino C, Neumann L, Weder C, Schrettl S. Approaches to polymeric mechanochromic materials. J Polym Sci Part A Polym Chem 2017;55:640–52. [15] Li J, Nagamani C, Moore JS. Polymer mechanochemistry: from destructive to productive. Acc Chem Res 2015;48:2181–90. [16] Brantley JN, Wiggins KM, Bielawski CW. Polymer mechanochemistry: the design and study of mechanophores. Polym Int 2013;62:2–12. [17] Black AL, Lenhardt JM, Craig SL. From molecular mechanochemistry to stress-responsive materials. J Mater Chem 2011;21:1655–63. [18] Klajn R. Spiropyran-based dynamic materials. Chem Soc Rev 2014;43:148–84. [19] Peterson GI, Larsen MB, Ganter MA, Storti DW, Boydston AJ. 3D-printed mechanochromic materials. ACS Appl Mater Interfaces 2015;7:577–83. [20] Li M, Zhang Q, Zhu S. Photo-inactive divinyl spiropyran mechanophore cross-linker for real-time stress sensing. Polymer 2016;99:521–8. [21] Li M, Liu W, Zhu S. Smart polyolefins feeling the force: color changeable poly(ethylene-vinyl acetate) and poly(ethylene-octene) in response to mechanical force. Polymer 2017;112:219–27. [22] O’Bryan G, Wong BM, McElhanon JR. Stress sensing in polycaprolactone films via an embedded photochromic compound. ACS Appl Mater Interfaces 2010;2:1594–600. [23] Jiang S, Zhang L, Xie T, Lin Y, Zhang H, Xu Y, et al. Mechanoresponsive PS-PnBA-PS triblock copolymers via covalently embedding mechanophore. ACS Macro Lett 2013;2:705–9. [24] Wang L-J, Zhou X-J, Zhang X-H, Du B-Y. Enhanced mechanophore activation within micelles. Macromolecules 2016;49:98–104. [25] Lee CK, Davis DA, White SR, Moore JS, Sottos NR, Braun PV. Force-induced redistribution of a chemical equilibrium. J Am Chem Soc 2010;132:16107–11. [26] Kingsbury CM, May PA, Davis DA, White SR, Moore JS, Sottos NR. Shear activation of mechanophore-crosslinked polymers. J Mater Chem 2011;21:8381–8. [27] Gossweiler GR, Hewage GB, Soriano G, Wang Q, Welshofer GW, Zhao X, et al. Mechanochemical activation of covalent bonds in polymers with full and repeatable macroscopic shape recovery. ACS Macro Lett 2014;3:216–9. [28] Beiermann BA, Davis DA, Kramer SLB, Moore JS, Sottos NR, White SR. Environmental effects on mechanochemical activation of spiropyran in linear PMMA. J Mater Chem 2011;21:8443–7.

[29] Celestine A-DN, Beiermann BA, May PA, Moore JS, Sottos NR, White SR. Fracture-induced activation in mechanophore-linked, rubber toughened PMMA. Polymer 2014;55:4164–71. [30] Silberstein MN, Min K, Cremar LD, Degen CM, Martinez TJ, Aluru NR, et al. Modeling mechanophore activation within a crosslinked glassy matrix. J Appl Phys 2013;114, 023504/1-9. [31] Beiermann BA, Kramer SLB, Moore JS, White SR, Sottos NR. Role of mechanophore orientation in mechanochemical reactions. ACS Macro Lett 2012;1:163–6. [32] Degen CM, May PA, Moore JS, White SR, Sottos NR. Time-Dependent mechanochemical response of SP-cross-linked PMMA. Macromolecules 2013;46:8917–21. [33] Hemmer JR, Smith PD, van Horn M, Alnemrat S, Mason BP, de Alaniz JR, et al. High strain-rate response of spiropyran mechanophores in PMMA. J Polym Sci Part B Polym Phys 2014;52:1347–56. [34] Kim JW, Jung Y, Coates GW, Silberstein MN. Mechanoactivation of spiropyran covalently linked PMMA: effect of temperature, strain rate, and deformation mode. Macromolecules 2015;48:1335–42. [35] Silberstein MN, Cremar LD, Beiermann BA, Kramer SB, Martinez TJ, White SR, et al. Modeling mechanophore activation within a viscous rubbery network. J Mech Phys Solids 2014;63:141–53. [36] Beiermann BA, Kramer SLB, May PA, Moore JS, White SR, Sottos NR. The effect of polymer chain alignment and relaxation on force-Induced chemical reactions in an elastomer. Adv Funct Mater 2014;24:1529–37. [37] Lee CK, Beiermann BA, Silberstein MN, Wang J, Moore JS, Sottos NR, et al. Exploiting force sensitive spiropyrans as molecular level probes. Macromolecules 2013;46:3746–52. [38] Chen Y, Zhang H, Fang X, Lin Y, Xu Y, Weng W. Mechanical activation of mechanophore enhanced by strong hydrogen bonding interactions. ACS Macro Lett 2014;3:141–5. [39] Fang X, Zhang H, Chen Y, Lin Y, Xu Y, Weng W. Biomimetic modular polymer with tough and stress sensing properties. Macromolecules 2013;46:6566–74. [40] Zhang H, Chen Y, Lin Y, Fang X, Xu Y, Ruan Y, et al. Spiropyran as a mechanochromic probe in dual cross-linked elastomers. Macromolecules 2014;47:6783–90. [41] Hong G, Zhang H, Lin Y, Chen Y, Xu Y, Weng W, et al. Mechanoresponsive healable metallosupramolecular polymers. Macromolecules 2013;46:8649–56. [42] Zhang Q, Wang Y, Xing C, Cai Y, Xi K, Jia X. Light and force dual-responsive waterborne polyurethane in multiple states. RSC Adv 2017;7:12682–9. [43] Brown CL, Barbee MH, Ko JH, Maynard HD, Craig SL. Writing without ink: a mechanically and photochemically responsive PDMS polymer for science outreach. J Chem Educ 2017, acs.jchemed.6b00806 (4 pp). [44] Gossweiler GR, Brown CL, Hewage GB, Sapiro-Gheiler E, Trautman WJ, Welshofer GW, et al. Mechanochemically active soft robots. ACS Appl Mater Interfaces 2015;7:22431–5. [45] Wang Q, Gossweiler GR, Craig SL, Zhao X, Bao Z. Cephalopod-inspired design of electro-mechano-chemically responsive elastomers for on-demand fluorescent patterning. Nat Commun 2014;5:4899. [46] Cho S, Kang S, Pandya A, Shanker R, Khan Z, Lee Y, et al. Large-area cross-aligned silver nanowire electrodes for flexible, transparent, and force-sensitive mechanochromic touch screens. ACS Nano 2017;11:4346–57. [47] Grady ME, Beiermann BA, Moore JS, Sottos NR. Shockwave loading of mechanochemically active polymer coatings. ACS Appl Mater Interfaces 2014;6:5350–5. [48] Li M, Liu W, Zhang Q, Zhu S. Mechanical force sensitive acrylic latex coating. ACS Appl Mater Interfaces 2017;9:15156–63. [49] Potisek SL, Davis DA, Sottos NR, White SR, Moore JS. Mechanophore-linked addition polymers. J Am Chem Soc 2007;129:13808–9. [50] May PA, Munaretto NF, Hamoy MB, Robb MJ, Moore JS. Is molecular weight or degree of polymerization a better descriptor of ultrasound-induced mechanochemical transduction. ACS Macro Lett 2016;5:177–80. [51] Lee CK, Diesendruck CE, Lu E, Pickett AN, May PA, Moore JS, et al. Solvent swelling activation of a mechanophore in a polymer network. Macromolecules 2014;47:2690–4. [52] Li M, Lei L, Zhang Q, Zhu S. CO2—Breathing induced reversible activation of mechanophore within microgels. Macromol Rapid Commun 2016;37:957–62. [53] Chen H, Yang F, Chen Q, Zheng J. A novel design of multi-mechanoresponsive and mechanically strong hydrogels. Adv Mater 2017, 1606900/1-8. [54] Suslick KS. Sonochemistry. Science 1990;247:1439–45. [55] Suslick KS, Price GJ. Applications of ultrasound to materials chemistry. Annu Rev Mater Sci 1999;29:295–326. [56] Basedow AM, Ebert KH. Ultrasonic degradation of polymers in soltion. Adv Polym Sci 1977;22:83–148. [57] May PA, Moore JS. Polymer mechanochemistry: techniques to generate molecular force via elongational flows. Chem Soc Rev 2013;42:7497–506. [58] Kim TA, Beiermann BA, White SR, Sottos NR. Effect of mechanical stress on spiropyran-merocyanine reaction kinetics in a thermoplastic polymer. ACS Macro Lett 2016;5:1312–6. [59] Gossweiler GR, Kouznetsova TB, Craig SL. Force-rate characterization of two spiropyran-based molecular force probes. J Am Chem Soc 2015;137:6148–51. [60] Kean ZS, Black Ramirez AL, Yan Y, Craig SL. Bicyclo[3.2.0] heptane mechanophores for the non-scissile and photochemically reversible generation of reactive bis-enones. J Am Chem Soc 2012;134:12939–42. [61] Wiggins KM, Syrett JA, Haddleton DM, Bielawski CW. Mechanically facilitated retro [4 + 2] cycloadditions. J Am Chem Soc 2011;133:7180–9.

M. Li et al. / Progress in Polymer Science 79 (2018) 26–39 [62] Kryger MJ, Munaretto AM, Moore JS. Structure–mechanochemical activity relationships for cyclobutane mechanophores. J Am Chem Soc 2011;133:18992–8. [63] Schanda J, editor. Colorimetry: Understanding the CIE System. Hoboken, NJ, USA: John Wiley & Sons, Inc.; 2007, 467 pp. [64] Ramirez ALB, Kean ZS, Orlicki JA, Champhekar M, Elsakr SM, Krause WE, et al. Mechanochemical strengthening of a synthetic polymer in response to typically destructive shear forces. Nat Chem 2013;5:757–61. [65] Wang J, Piskun I, Craig SL. Mechanochemical strengthening of a multi-mechanophore benzocyclobutene polymer. ACS Macro Lett 2015;4:834–7. [66] Kean ZS, Niu Z, Hewage GB, Rheingold AL, Craig SL. Stress-responsive polymers containing cyclobutane core mechanophores: reactivity and mechanistic insights. J Am Chem Soc 2013;135:13598–604.

39

[67] Zhang H, Gao F, Cao X, Li Y, Xu Y, Weng W, et al. Mechanochromism and mechanical-force-triggered cross-linking from a single reactive moiety incorporated into polymer chains. Angew Chemie Int Ed 2016;55:3040–4. [68] Li J, Shiraki T, Hu B, Wright RAE, Zhao B, Moore JS. Mechanophore activation at heterointerfaces. J Am Chem Soc 2014;136:15925–8. [69] Woodcock JW, Beams R, Davis CS, Chen N, Stranick SJ, Shah DU, et al. Observation of interfacial damage in a silk-epoxy composite, using a simple mechanoresponsive fluorescent probe. Adv Mater Interfaces 2017;4, 1601018/1-7.