Recent Developments in Polydiacetylene-Based Sensors - Chemistry

Jan 23, 2019 - US government shutdown woes hit chemical industry. Impacts include delays in chemical reviews, company acquisitions ...
258 downloads 0 Views 5MB Size
Subscriber access provided by EKU Libraries

Review

Recent Developments in Polydiacetylene-Based Sensors Xiaomin Qian, and Brigitte Städler Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.8b05185 • Publication Date (Web): 23 Jan 2019 Downloaded from http://pubs.acs.org on January 24, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 26 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Recent Developments in Polydiacetylene-Based Sensors Xiaomin Qian, Brigitte Städler* Interdisciplinary Nanoscience Center (iNANO), Aarhus University, Gustav Wieds Vej 14, 8000 Aarhus, Denmark ABSTRACT: Conjugated polymers are intriguing materials that offer potential practical applications in diverse interdisciplinary subjects. Among them, polydiacetylenes (PDAs) have been extensively studied due to their interesting structural, spectral and optical features. In particular, the unique colorimetric and fluorescent transition of PDAs in response to different external stimuli make them a novel class of sensing materials, and numerous applications of PDAs as bio- or chemosensors have been explored in the past few decades. In this review, we summarize the latest developments with regard to the applications of PDAs as a class of sensing materials presented in the literature since 2014. This review is sorted into categories based on the structural differences of diacetylene monomers, from which PDAs are generated. In addition, different forms of PDAs and various methods to improve the sensing performances of PDAs are also emphasized.

INTRODUCTION

that the conformational change of the polymer backbone from planar to non-planar upon stimulation contributes to the blue-shift in the spectra.14-15 Besides, the blue phase PDAs are non-fluorescent, while in the red phase, a red fluorescence with negligible beaching is usually observed. This is suggested to be the result of an energy shift in the lowest excited state from the blue phase to the red phase. Therefore, a PDA-based system has dual-signal outputs when being used as a sensing platform. Ever since the first report involving a sialic acid modified PDA as a selective sensor for influenza virus presented by Charych et al. in 1993,16 PDA-based sensors have attracted the attention of scientists yielding in a variety of reports using PDAs as chemo/bio-sensing materials based on their colorimetric and fluorescent transition. To date, the stimuli that are able to induce the optical transition of PDAs come in numerous forms, such as solvents (solvatochromism),17 heat (thermochromism),18 mechanical stress (mechanochromism),19 light (photochromism),20 pH,21-23 metal ions,24-27 anions,28 surfactants,29-32 microorganisms,33-39 and biomolecules,40-53 which have greatly broadened the applications of PDAs as versatile sensing materials in various fields.

The development of fluorescent and colorimetric sensors for the efficient detection of chemically, biologically, and environmentally important molecules has gained continuous attention during the last decades.1-4 As compared to small molecule-5-7 and nanoparticlebased8-9 sensors, conjugated polymer-based sensors 10-11 have several pronounced advantages, such as enhanced binding efficiency, amplified signal output, improved stability, and easy fabrication into devices, etc. Polydiacetylenes (PDAs), a unique class of conjugated polymeric material that combines highly ordered backbones with customizable side chains, have been extensively studied ever since the first preparation by Wegner in 1969.12 PDAs are usually prepared from the 1, 4-addition of diacetylene monomers facilitated by ultraviolet (UV) or γ light irradiation, which generates the alternating CC double bonds and triple bonds (ene-yne) polymeric backbone. The precondition for the successful topochemical polymerization of diacetylenes is that they must be self-assembled to meet specific geometrical parameters. An optimal packing orientation of the diacetylene units is required to promote propagation of the liner chain through the ordered phase. Polymerization of diacetylenes stabilizes the physical structure, enhances thermal stability and reinforces mechanical stability of the system.13 Unlike most of the covalent polymers that are synthesized via chemical reactions with high temperatures, chemical initiators, catalysts, or heating are not required for the polymerization process of diacetylenes, thus simplifies the preparative process and ensures the purity of the resulting polymer. The most meritorious aspect of PDAs is their unique optical feature, originating from the existence of extensively delocalized π-electron networks and conformational restrictions along the main backbone. It has been manifested that in most cases, PDAs show an absorption peak at ca. 640 nm due to electron delocalization within the conjugated backbone, which appear visually as an intense blue color. Upon interaction with external stimuli, the main absorption peak shifts hypsochromically to ca. 540 nm exhibiting a red color, which can be easily detected by naked eye. The detailed mechanism for the color transition remains to be fully elucidated. It is usually presumed

Even though the exact mechanism for the optical transition of PDAs induced by external stimuli remains debatable, it is widely accepted that the pendant side chains of PDAs play a significant role in the optical transition. Specifically, the interactions between the side chains themselves, as well as the interaction between functional groups (head groups) on the side chains and the stimuli prominently influence the overall conformation of the polymer chain, thus leading to the optical change. The interaction between the side chains can be manipulated by adjusting the integral topological structure of the side chains including the length of the alkyl chain, the position of the butadiene group, and the type of head group, which is crucial for sensing stimuli from ambient environment, such as temperature and mechanical stress instead of a specific analyte. This can be considered as “intramolecular” modification. The interaction between the head groups and the stimuli are much more versatile. Using a “lock and key” strategy, it is plausible to modify the head groups of PDAs that can exclusively interact with the target analytes, depending on the property of the analytes, whether the interaction is a metal-ligand 1

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 26

coordination, a bio-recognition, or a chemical reaction. This case is considered as “intermolecular” modification. The functions of a material are not only determined by the compositions but also by the way the components are arranged, namely, in which form the material is presented. PDA materials have been structured in various forms, such as Langmuir-Blodgett (LB)54-57 and Langmuir-Schaefer (LS) films,58-60 nanoparticles in aqueous solution,61-62 and solid matrices,63-67 depending on the chemo-physical features of the diacetylene monomers and in which environment the PDAs are intended to be used. The ultrathin LB and LS films are intriguing because the molecular orientation is controllable and they can be potentially used as colorimetric sensing surfaces in micro- and nano-devices, as well as for fundamental studies using surface-sensitive probes including scanning probe microscopy. The most widely investigated form of PDA materials, however, is presented as self-assembled nanoparticles in aqueous solution, namely micelles, liposomes, and vesicles. The majority of PDAs reported as a sensing material were prepared as such nanoparticles due to its easy preparation and bio-compatibility. However, such particles constantly show a relatively poor sensitivity and stability. To improve the sensing performances, PDAs can be prepared or incorporated into a host matrix, such as paper, hydrogels, microbeads, membranes and so forth, to form composites. The ambient matrix has a significant influence on the optical and sensing properties of PDAs. Further, the embedding could also give access to PDA sensors in solid state, which shows several advantages over their solution-based counterparts, such as easy handling, enhanced stability and good portability.

Scheme 1. Schematic illustration of the three categories of diacetylenes discussed in this review. i) linear diacetylene with a polar head group and a hydrophobic alkyl tail; ii) linear diacetylene with polar head groups (left) or hydrophobic alkyl tails (right) at both ends; and iii) other unique circular (left) and dendritic (right) diacetylenes.

on the sensing performance are also discussed in between. Among the three categories listed above, the first one contributes the biggest part partially due to its facile synthesis and the starting materials can be obtained in a large quantity from commercial suppliers.

LINEAR DIACETYLENES: SINGLE POLAR HEAD GROUP AND HYDROPHOBIC TAIL

To date, the applications of PDAs as a sensing material have gone beyond sensing a certain analyte in an academic environment. The fusion of multiple disciplines, such as chemistry, biology and engineering, has opened new paths for PDAs. PDAs also demonstrated potential applications in numerous fields besides sensing, such as carriers for catalysts,68-71 drugs,72-77 siRNA,78 cells,79 and genes,80-81 as components in cell imaging,82-83 tumor targeting,84-85 solar cells,86-88 gas sorption,89-90 organic field-effect transistors,91-92 actuator,93 supercapacitor,94 and polymer stabilizer,95 owing to their unique structural and physical features. In this review, only the sensing applications of PDAs are discussed.

Synthetic Strategy PDAs are formed via the 1,4-photopolymerization of diacetylenes. The majority of diacetylenes reported consist of two segments: a hydrophilic polar head group and a hydrophobic non-polar tail linked by the butadiyne moiety. From a synthetic point of view, diacetylenes in this group are usually prepared with the classical CadiotChodkiewicz coupling reaction, starting with an acetylene and a halo-acetylene, catalyzed by Cu(I) in the presence of an amine as the base.101 Various diacetylenes with different lengths of alkyl chains were synthesized using this protocol. However, an easier and more common way to obtain desired diacetylenes is to chemically modify the head groups of commercial diacetylenes. The selectivity of PDAs towards a certain analyte can be accomplished by modifying the head groups (in most cases, a carboxylic group) of diacetylenes with target-recognizable receptors. 10,12-Pentacosadiynoic acid (PCDA) and 10,12-tricosadiynoic acid (TCDA) are the two most common commercially available diacetylenes with a carboxylic acid head group. The high reactivity of the carboxylic acid group makes their modification straightforward, allowing to greatly enrich their complexities, as well as to afford more robust and sensitive PDA systems. Typically, the formation of amide, using either N-(3-dimethylaminopropyl)-N′-ethylcarbodiimide hydrochloride (EDC)/Nhydroxysuccinimide (NHS) coupling or acyl chloride, is employed. The EDC/NHS coupling reaction is especially useful for bio-conjugation with biomolecules because of the ubiquitously present amine groups on biomolecules. Further, the hydrogen bonds among the amide groups are also useful for the ordered self-assembly of diacetylenes, which is crucial for the successful topochemical polymerization. The esterification reaction of carboxylic acid group with hydroxyl group are also regularly used. To functionalize diacetylenes with peptide segments, a standard Fmoc (9-fluorenylmethoxycarbonyl) solid-phase peptide synthesis is frequently exploited.102-105

A considerable amount of comprehensive reviews on the topic of PDAs from different perspectives were published in the past decade. Lebègue et al. primarily discussed PDA vesicles for bio-sensing microorganisms.96 Lee et al. put emphasis on the varieties of the target analytes that were detectable by PDAs.97 Diegelmann et al. focused on one dimensional (1D) nanomaterials self-assembled by peptidefunctionalized PDAs.98 Chen et al. summarized the PDA sensors according to the approaches of inducing acceptors into a polymer matrix.99 Yarimaga et al. reviewed PDA materials mainly from a microscopic dimensional point of view.100 Complementary, in this review, we highlight the most recent developments of PDAs and their sensing applications from the past 5 years by setting focus on their structural features. Specifically, we organized the review based on the different topological structures of diacetylenes used to form PDAs (Scheme 1), which include i) linear diacetylene with a polar head group and a hydrophobic alkyl tail. ii) linear diacetylene with polar head groups or hydrophobic alkyl tails at both ends; and iii) other unique diacetylenes such as circular and dendritic diacetylenes. The classification is proposed based on the fact that the topological structures of diacetylenes significantly affect their self-assembly, which in a way determines the final optical properties of PDAs. Further, the different forms of PDA-based sensors and their influences 2

ACS Paragon Plus Environment

Page 3 of 26 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 1. PDA sensor arrays for the differentiation of organic solvents based on solvatochromism. (a) i) Chemical structures of the four diacetylenes used for inkjet printing. ii) Schematic illustration of depositing the four PDAs onto paper using inkjet printing. iii) Color patterns of the PDA-embedded paper after exposure to different organic solvents. THF = tetrahydrofuran, EA = ethyl acetate, IPA = isopropyl alcohol, ether = diethyl ether, AcN = acetonitrile. (b) i) Chemical structures of the three diacetylenes used to incorporate into silica aerogel. ii) Schematic presentation of the preparation of PDA/aerogel composite. iii) Color patterns of the three composites after exposure to different organic solvents. (c) The colorimetric responses of the PDA embedded matrix polymers upon exposure to different organic solvent vapors at 0.4% (v/v) for 5 min. Panel a adapted with permission from ref 117. Copyright 2018 American Chemical Society. Panel b adapted with permission from ref 118. Copyright 2017 American Chemical Society. Panel c reprinted with permission from ref 119. Copyright 2017 The Royal Society of Chemistry.

teracting directly with certain solvents, which could again be a challenge and hinders the wide application. Alternatively, the use of sensor arrays consisting of a group of solvatochromic molecules to build a “fingerprint” of each solvent is considered. PDAs are widely exploited in this context due to their relatively simple chemical modification, giving access to a large structural diversity of diacetylenes as the precusors.113-116

Applications It is widely accepted that the alkyl chain length, the position of butadiyne moiety in the molecule, and the polar head have a pronounced impact on self-assembly behaviors of diacetylenes,106-107 as well as the optical properties of PDAs, i.e., the blue-to-red colorimetric transition along with a fluorescence enhancement upon interaction with various stimuli.108-109 The applications of PDAs are in a way determined by the molecular structure of the diacetylene monomers, but it is also greatly affected by the matrix that they are embedded in. In this section, we sorted the sensing applications of PDAs based on their different sensing mechanisms. Further, in the “affinochromism” section, sensors were again subdivided based on the forms of PDAs sensors for a better clarity.

A smartphone-based application of PDAs for the differentiation of common organic solvents was presented by Park et al.117 Starting with four well-designed diacetylenes (Figure 1ai), the authors deposited the four monomers onto conventional paper using an inkjet printer, followed by photo-polymerization (Figure 1aii). Exposing this PDA sensor array to 11 different organic solvents generated a color pattern for each solvent as a result of the solvatochromism of PDAs, as shown in Figure 1aiii. A database of color changes (red and hue values) of each PDA upon interaction with a specific solvent, i.e., a “fingerprint” of each solvent, was established. More impressively, the authors developed a smartphone application, which allowed for the identification of an unknown solvent by simply taking a photograph of the paper after interaction with the solvent and compare the color patterns with the database. This successful effort opens up endless opportunities for the practical applications of PDA sensor arrays.

Solvatochromism PDAs undergo colorimetric and fluorescent changes that are dependent on the polarity of solvents, which is usually non-specific and irreversible. The optical transition is intimately related to a corresponding conformational switch of PDAs upon interaction with different solvents. As an alternative to conventional methods like gas chromatography-mass spectrometry (GC-MS), the solvatochromism of PDAs makes them potential sensors to differentiate solvents visually while circumventing shortcomings such as the need for welltrained experts and expensive equipment as well as the lack of immediate response. Undoubtedly, the colorimetric differentiation of solvents with similar polarities can be extremely challenging using a single-component sensor since the sensor inevitably exhibits broad overlaps in the absorption and emission bands in different solvents. As far as we know, only a few reported PDAs showed a specific color transition to a specific solvent.110-111 To overcome the non-specificity, Lee et al. developed a protective layer method that enabled the visual differentiation of dichloromethane from chloroform.112 However, in this strategy, specific polymers are needed to prevent PDAs from in-

Another example of constructing a PDA-based sensor array for the differentiation of organic solvents is presented by Susmita and co-workers.118 Inspired by the fact that silica aerogels have very low density and exhibit porous structures with high surface areas, which are promising for the rapid adsorption of small guest molecules, three PDA/aerogel composites with different PDAs were prepared using a facile drop-casting method (Figure 1bi, 1bii). The aerogels acted as the adsorbents for the solvent molecules while the PDAs were the signal transduction elements. The PDA/aerogel composites exhibited rapid color change (Figure 1biii) and fluorescence 3

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 26

enhancement towards different solvents even at a concentration as low as 100 to 200 ppm, allowing for the differentiation of the volatile solvents. The key to construct a fingerprint for each solvent is to introduce a reasonable number of variables. Here the variable can be different diacetylenes, as illustrated in the two examples above, but it can also be, for example, different matrices that PDAs are embedded into. Tu et al. incorporated PCDA into the matrices of four different polymers, polyvinylpyrrolidone (PVP), polyethylene glycol (PEG), polyacrylic acid (PAA), and poly-4-vinylpyridine (P4VP).119 It was found that the polymer matrix has a significant influence on the color transitions of PDAs. By carefully optimizing the weight ratio between diacetylene and polymer, a fingerprint for each solvent was obtained (Figure 1c), which allowed for the colorimetric differentiation of the solvents. Thermochromism When the ambient temperature varies in a certain range, PDAs with thermochromism undergo colorimetric transitions, usually from blue to red as the temperature increases gradually. Thermochromism is likely the most exhaustively investigated chromatic effect of PDAs ever since the first report in 1976.120 These studies not only facilitated the fundamental understanding of the color transition but also made the development of temperature sensitive materials possible. Numerous studies have revealed that the molecular structures of diacetylenes,108, 121-129 together with the matrices that PDAs are embedded in,130-131 have a significant influence on the thermochromic behaviors of PDAs. PDAs with different color transition temperatures could be obtained by carefully manipulating these two parameters. The most intriguing characteristic of PDAs thermochromism is the reversibility within a certain temperature range. This is in stark contrast to other stimuli-triggered chromism of PDAs, most of which are typically irreversible. Studies have revealed that the reversible thermochromic behaviors of PDAs are largely dependent on the intramolecular non-covalent interactions among the side chains, i.e. hydrogen bonds, π-π stacking, Van der Waal’s force and so forth.59, 124, 132-140 Generally speaking, the stronger the interactions are, the better thermochromic properties PDAs show, i.e., reversibility in a wider temperature range. Thermochromic reversibility of PDAs have been achieved by strengthening the intramolecular interactions among the side chains. For example, Guo et al. prepared a peptide-decorated PDA, which has multiple hydrogen bonding sites, that has shown a reversible thermochromism at temperature up to 200 °C.141 The transition took place even when the heating rate was up to 5000 K s-1. This remarkable thermochromism was attributed to the hierarchically assembled structure promoted by the multiple hydrogen bonds among the peptide segments. As alternatives for obtaining reversible thermochromism from a molecular level, intercalation with polymers,142-144 organic amines,145-146 layered double hydroxide nanosheets,64 and metal ions147-154 to form layered nanocomposites or co-assembly with small molecule melamine155 were successfully employed.

Figure 2. PDA materials with reversible thermochromism. (a) i) Photograph of the purple-colored MMT-polyPCDA-APTES-I artificial nacre. ii) Scanning electron microscopy (SEM) images of the front fracture morphology of the artificial nacre. iii) Photographs showing the reversible thermochromism of the artificial nacre at different temperatures. iv) A schematic representation of the spraying coating process. v) The thermochromic behavior of a leaf pattern on paper (scale bar = 1 cm). (b) i) Chemical structure of the peptide-modified diacetylene. ii) Schematic illustration of the preparation of the PDA-embedded fiber. iii) Photographs of the fiber upon power on/off. iv) and (v) Photographs of the fiber showing excellent flexibility and stretchability while retaining thermochromism. Panel a adapted with permission from ref 156. Copyright 2017 American Chemical Society. Panel b adapted from ref 157. Copyright 2016 The Royal Society of Chemistry.

colored nanocomposite MMT-polyPCDA-APTES-I (Figure 2ai) was fabricated with montmorillonite (MMT), PCDA and (3ami)triethoxysilane (APTES) following a facile procedure. The crosslinking of MMT with APTES through a condensation reaction imparted the nanocomposite with excellent tensile strength. Microscopic studies revealed that the nanocomposite showed a nacre-like layered morphology (Figure 2aii). More impressively, the nanocomposite showed reversible thermochromism between 20 to 70 °C due to the encapsulation of PDAs (Figure 2aiii). The fabrication of such an artificial nacre could easily be scaled up e.g., by using spray coating technique, and thermo-responsive patterns on different substrates could be easily obtained (Figure 2aiv, 2av). Instead of direct heating, the thermal energy required to fulfill thermochromism could be generated by applying an electrical current on a substrate, which is quite useful in situations where direct heating is not feasible. Lu et al. prepared a flexible and stretchable fiber with excellent reversible thermochromism.157 A diacetylene monomer modified with a peptide segment (Figure 2bi) was synthesized and dip-coated onto an elastic conductive fiber, followed by photopolymerization in situ and coating with a layer of silicon for safety considerations (Figure 2bii). It was found that when an electric current passed through the blue fiber, it became red, which is the typical color of PDAs under perturbation. The response time was relevant to the strength of the current, and the original color was recovered in 20 s after electricity was cut off (Figure 2biii). The redpurple color switch could be repeated for 1000 cycles without significant loss, exhibiting excellent reversibility and stability.

Based on the thermochromism of PDAs, various novel temperature-sensitive materials were fabricated. Inspired by natural nacres, which have outstanding mechanical properties derived from their layered structure and synergistic interfacial interactions, Peng and co-workers fabricated a PDA-embedded artificial nacre, which integrated mechanical robustness with thermochromism.156 A purple

4

ACS Paragon Plus Environment

Page 5 of 26 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 3. Mechanical stress induced chromism of PDA sensors. (a) i) Photographs of the blue PDA/PDMS film and schematic illustration of PDAs in the polymer matrix. ii) Photographs of the PDA/PDMS films after immersion into pentane at different time points. iii) Plots of red intensity and degree of swelling of the PDA/PDMS film in alkanes with different carbon lengths after 10 min of incubation. iv) Red intensity of the PDA/PDMS film after 10 min of exposure to different kerosene/diesel oil mixtures. Horizontal axis is the volumetric percentage of kerosene in the mixtures. Inset shows the colors of corresponding PDA/PDMS films. (b) Photographs of the PDA/PDMA microbeads upon exposure to different SAHCs: pentane (top left), heptane (top right), nonane (bottom left), and undecane (bottom right). The time interval is 10 s. (c) Photograph of the dehydrated PDA/alginate hydrogel after exposure to water. Panel a adapted with permission from ref 161. Copyright 2014 John Wiley and Sons. Panel b reprinted with permission from ref 162. Copyright 2015 American Chemical Society. Panel c reprinted with permission from ref 163. Copyright 2015 American Chemical Society.

workers reported an example of incorporating PDAs into polydimethylsiloxane (PDMS) to generate a blue PDA/PDMS film (Figure 3ai) with a straightforward mixing-irradiation-curing method.161 Such PDA/PDMS films expressed excellent colorimetric response from blue-to-red towards SAHCs arising from the hydrocarbon-induced swelling of PDMS matrix (Figure 3aii). Results showed that the alkyl chain length of the SAHCs greatly affected the rate of swelling and the color change of the PDA/PDMS film (Figure 3aiii). As can be seen in Figure 3aiv, the sensor could distinguish kerosene from diesel oil with a clear different colorimetric response, even though they have very similar chemical compositions. In a follow up effort, PDMS microbeads with embedded PDAs were prepared, which could visually differentiate SAHCs with different lengths based on the same swelling-induced color change principle (Figure 3b).162 Similarly, 1D PDA nanofibers were integrated into the three dimensional (3D) matrix of a hygroscopic alginate hydrogel via ionic interaction with the aim to detect water.163 Swelling of the dry hydrogel matrix caused volumetric expansion upon interaction with water, which imposed mechanical stress on the PDA nanofibers that interrupted the conformation of the PDA backbone, leading to a distinct blue-to-red color change (Figure 3c).

The color transition was primarily triggered by an electrothermal effect. In other words, when the electric current passed through the fiber, the outermost PDAs were rapidly heated, which exceeded the thermochromic transition temperature of PDAs, thus leading to the color switch. The multiple hydrogen bonds of the peptide segment ensured the reversibility of the color transition. Furthermore, the thermochromism was not affected when the composite fiber was twisted, bended and winded, showing superior flexibility and stretchability (Figure 2biv, 2bv). Such an electrothermal effect-induced thermochromism of PDAs could also be used for flexible image display, as illustrated by Shin et al.125 Mechanochromism An alternative sensing mechanism considers applying mechanical stress to induce the color transition of PDAs.158-160 When a proper amount of mechanical stretching energy is delivered to the PDA backbone, the disruption of the π–orbitals takes place to induce a colorimetric transition. Such a mechanical stress-induced phenomenon is rather useful for sensing. Specifically, it is possible to detect a certain analyte, which does not interact directly with PDAs but induces a mechanical change in the matrices that PDAs are encapsulated in. Once the matrices are disrupted by the analyte mechanically, the mechanical stretching energy is delivered to the PDA backbone, which leads to the color change. This sensing mechanism is particularly useful for analytes that are chemically or biologically inert and difficult to be detected, for example, saturated aliphatic hydrocarbons (SAHCs). The colorimetric differentiation of SAHCs remains challenging because of the nonpolar nature and deficiency of functional groups that can interact with the sensor system. Park and co-

Similar to thermochromism, the mechanochromism of PDAs could also be manipulated by an intercalation approach. The group of Yuya Oaki prepared organic layered composites by the intercalation of a series of different organic amines into diacetylene crystals, followed by photopolymerization.164-165 Results showed that the color transition temperature of such composites were dependent on structures of the amines intercalated. Moreover, the accumulated

5

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 26

Figure 4. Photochromic PDAs. (a) i) Chemical structures of PCDA-NP and PCDA-DMN. ii) Photograph of PCDA-NP casted on a glass slide after polymerization with patterning in the center using 365 nm UV light via a photomask (left, scale bar = 5 mm) and fluorescence image of a patterned polymerized PCDA-NP casted on a glass slide (right, scale bar = 100 μm). (b) i) Chemical structure of HDDA-Azo. ii) Optical microscopy image of the polymerized HDDA-Azo crystal before and after irradiation with UV light. The white dashed circle indicates where the crystal tearing started. (c) i) Chemical structures of SPDA and CODA. ii) Illustration of the reversible multi-stimuli-responsive property of PDA vesicles. Panel a adapted with permission from ref 170. Copyright 2018 American Chemical Society. Panel b adapted with permission from ref 171. Copyright 2016 The Royal Society of Chemistry. Panel c adapted with permission from ref 172. Copyright 2013 John Wiley and Sons.

mechanical stress applied could be visualized by the color, which was determined by the structural stability of the composites. Namely, the thermochromic and mechanochromic properties of PDAs could be fine-tuned by the intercalation of different amines, which can be applied to a variety of sensing materials.

of NP and DMN upon irradiation (l = 365 nm). Therefore, when PCDA-NP and PCDA-DMN were polymerized into PDAs, the intense blue color shifted towards red upon irradiation, accompanied with a fluorescence enhancement due to the light induced cleavage of NP and DMN and resulting PDA backbone disturbance, which allowed for convenient photo patterning using a photo mask (Figure 4aii).

Photochromism Diacetylene monomers per se are photo-responsive once they selfassembled into well-aligned patterns that can go through topochemical polymerization to form PDAs because of the light-sensitive butadiyne moiety. Generally, PDAs can be tailor-made to be light-sensitive by grafting photo-responsive moieties into the head groups of diacetylenes20, 135, 166-167 or by doping photo-sensitive moieties that could interact with the head groups of PDAs into the system.168 However, Yan et al. demonstrated that PDA vesicles without functionalization of the head group could also be made light-sensitive via a combination method of thin-film hydration and supercritical CO2 fluid treatment.169 Importantly, the photo wavelength for the stimulation of PDAs should be different from the 254 nm UV light used for the polymerization of diacetylenes in order to ensure control over the process. Upon irradiation with light, the conformational or structural change of the light-sensitive moieties could lead to the interruption of the PDA backbone, which may result in the color transition. Such light-driven color transition could be used for e.g., encrypting information and anti-counterfeiting, where rapid and naked-eye detection is needed. Lee et al. synthesized two photo-responsive PDAs and explored their potential applications as anticounterfeiting materials.170 Two diacetylene monomers PCDA-NP and PCDA-DMN with 6-nitropiperonyl (NP) and 4,5-dimethoxy2-nitrobenzyl alcohol (DMN) as the head groups were synthesized (Figure 4ai) for the purpose of the well-known photolysis process

Baek et al. modified the head group of the diacetylene 4,6-heptadecadiynoic acid with azobenzene, which is one of the most extensively studied photo-responsive moiety that undergoes photo-isomerization of trans and cis isomers.171 Interestingly, microscopic images showed that PDA crystals derived from monomer HDDA-Azo (Figure 4bi) showed a UV-induced (l = 330-385 nm) blue-to-red color transition, accompanied with crystal tearing. The recovery of the pristine shape and blue color of HDDA-Azo crystal was observed when it was placed in the dark (Figure 4bii). Studies of a series of HDDA-Azo analogues suggested that the trans-cis conformational transition of the azobenzene moiety and the amide-provided hydrogen bonds in the polymer matrix were the reason for this novel reversible color switch and crystal tearing phenomenon. A novel multi-stimuli-responsive fluorescence probe is presented by Xia et al.172 It is well known that spiropyran units undergo reversible isomerization between the colorless closed (SP) form and the purple open merocyanine (MC) form with the assistance of visible light and UV light. Also the purple MC form could be protonated in acidic condition to form a colorless protonated MC (PMC) form, which could be deprotonated again in basic environment.173 Therefore, by incorporating this light and pH-sensitive spiropyran group into diacetylene (SPDA), then co-assembled with a coumarin-sub6

ACS Paragon Plus Environment

Page 7 of 26 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials finitive behavior between them is the primary way to allow the detection of a certain analyte. The affinity can be a non-covalent interaction or biomolecular recognition. Affinochromism has been one of the most widely explored properties of PDAs because it is customizable for a specific target analyte. Depending on the chemical or biological activity of the analyte, the head groups of PDAs can be tailor-made to fulfill the detection of the specific analyte with high selectivity. Most of the PDA-based sensors can be included into this group. However, it is worth noting that sometimes the analyte-active group does not necessarily have to be covalently grafted onto the diacetylene monomers. It could also be inserted into the diacetylenes and co-assembles with them to form the PDA-based sensor.185-187 In fact, such a physical doping method offers several advantages over the chemical attachment method, such as the absence of a chemical synthesis and the minimal interruption of the organization and selfassembly of the diacetylene monomers. Herein, for a better clarity, we sorted the affinochromism sensors based on which form they were constructed, namely in solution, in solid substrates or in other forms. The form has a significant influence on the sensing properties of PDA-based sensors. Cases in which PDAs were used to stabilize the sensors188-193 are not discussed here since the output signals were not generated from PDAs.

stituted diacetylene (CODA)(Figure 4ci), the resulting PDA vesicles became responsive to several external stimuli, including heat, light, and pH. More importantly, the fluorescence of the system could be switched on and off owing to the fluorescence resonance energy transfer between the red phase PDA and the open MC form of spiropyran. Furthermore, a set of logic gate operations based on the fluorescence switches of the designed sensor were realized (Figure 4cii), which could be a promising candidate to construct more complicated logic gates for the detection of analytes in biological and biomedical systems. Electrochromism Generally, PDAs are not electrochromic due to the poor electrical conductivity,174-175 which hampers the electrons from inducing a conformational change of PDA backbone. Also the difficulty in doping because of its rigid crystal microdomains makes it undesirable for electrochromic applications. However, it has been shown that the conductivity of a nano-structured PDA has reached 3 x 10-2 S/cm, which could be classified as a semiconductor.176 This high conductivity is likely caused by the scaling down of the sample size, which increased the surface to volume ratio, promoting the doping efficiency. The most common solution to remediate the conductivity problem is to integrate PDAs with a conductive material to form a nanocomposite. Carbon nanotubes (CNTs) are ideal candidates in this case because of their excellent conductivity177 and directional self-assembly of the diacetylenes on their surface due to the strong interactions between CNTs and the hydrophobic chains of diacetylenes.94, 178-182 The first current-induced reversible color change of PDAs on CNTs was reported by Peng et al. in 2009.183 This electrochromism was attributed to the 3D hopping conduction in the CNTs ribbon. In 2013, Zhang et al. described that in the presence of polymethylmethacrylate (PMMA), PDAs coated on a graphene surface showed electrical current-induced color transition and fluorescence enhancement.184 In that case, PDAs acted as the electrochromic material, graphene performed as the conductive matrix and PMMA ensured the mechanical strength of the composite and made the color transition visually more distinct. To the best of our knowledge, there were only two paper regarding the electrochromism of PDAs published in the past five years. Hanse et al. coated PDAs on the surface of multi-walled CNT by in situ photopolymerization, which led to an electrochromic composite.179 Raman spectra studies revealed that the stress in the PDA backbone was not crucial for the current-induced color transition. It was proposed that the destacking of π-orbitals of PDAs under current flow was the physical origin of electrochromism in CNT/PDA composites. In a follow up work, it was shown that the critical transition voltage for the electrochromism of CNT/PDA composites could be tuned by two methods: heat treatment of the diacetylene monomers prior to polymerization or anchoring diacetylene monomers to ZnO nanoparticles prior to polymerization.178 The previously discussed flexible and stretchable PDMS fibers equipped PDAs showed that the chromatic transition of the fiber was induced by the heat generated by the passing current, i.e., the color change was more a result of thermochromism than of electrochromism.157

Solution-based sensors Vesicles/liposomes prepared by the self-assembly of amphiphilic diacetylenes in aqueous solution are the most commonly used form of PDAs for sensing. These two terms have been used somewhat interchangeably when referring to the nanoparticles composed of bilayers with an enclosed volume. To avoid confusion and misunderstanding, in this review, the term “vesicle” is used uniformly instead of “liposome” considering the fact that the former is a more generic term, even though the latter has been used frequently in the cited literature. Vesicles prepared from the self-assembly of diacetylenes exhibited extensive applications as sensors based on the typical blueto-red color change and the fluorescence enhancement upon stimulation. A comprehensive but not exhaustive summary of PDA-based sensors reported in the past 5 years is provided in the Supporting Information (Table S1), including the type of analyte, the form of the sensor, and detection limit if available. It is worth mentioning that, as compared to PDA sensors in other forms, a critical advantage of such nanoparticles in aqueous solution is their mimicry of cell membrane and applications in biological systems. In this context, a certain molar percentage of phospholipids191, 194 or surfactants195-196 were often added to co-assemble with diacetylenes with the aim to increase the fluidity characteristic of the artificial membrane structure with preserved ability for polymerization. Membranes with a higher fluidity are suggested to react more sensitively to external perturbations. Also, it was reported that the phospholipids or surfactants affected the sensing performance of the PDAs.197-198 Metal ions are harmful to our bodies and environments once they are accumulated and surpass a certain concentration threshold. It is widely known that various metal ions, especially transition metal ions, can coordinate with certain ligands to form coordination complexes. Not surprisingly, many PDA-based metal ion sensors were designed on the basis of this mechanism. Depending on the chelation between the metal ion and the ligand, the chelation process affects the backbone of PDAs and/or results in the aggregation of PDAs leading to the optical transition. Lead ion (Pb2+) is a highly toxic heavy metal ion that causes severe health problems to multiple organs and physiological systems once accumulated in the human body.199 PDAs had also found numerous applications as Pb2+ sensors

Affinochromism As mentioned in the introduction, the disturbance of the backbone of PDAs leads to a chromatic change and the interactions between the head groups and the analytes significantly affect the overall conformation of the polymer backbone. Therefore, manipulating the af-

7

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 26

were rarely investigated. Xu et al. demonstrated that the enantioselective polymerization of diacetylenes could be fulfilled in the liquid crystal phase using linearly polarized light under magnetic field. These PDAs could serve as a direct visual probe for the enantioselective recognition of D- and L-lysine.209 However, the preparation of such PDAs remains difficult as it requires expensive equipment and long procedures. Introducing a chiral group into diacetylenes has been the most commonly used strategy to sense molecules with chirality. Generally, the molecular chirality can be transferred to the supramolecular system during the self-assembly process. Li et al. developed an enantioselective sensor that can efficiently distinguish chiral sulfonamide.210 A L-glutamic acid terminated diacetylene TCDAGlu was designed and synthesized (Figure 6a). It was found that upon polymerization, TCDA-Glu self-assembled into either vesicles or helical supramolecular gel with opposite chirality, depending on the solvent in which the self-assembly took place. Despite the contrary chiralities, both of them showed a specific blue-to-red chromatic transition upon interaction with S-tert-butylsulfinamide (STBSA) while being inert to R-TBSA (Figure 6b, c). It was found that the color transition was attributed to the formation of hydrogen bonds between L-glutamic acid and TBSA, which led to distortion of the PDA backbone and thus a color change for S-TBSA. While the sensing mechanism for the chiral TBSA remained unknown, it was believed that the enantioselective difference was caused by the different kinds of hydrogen bonding pattern between TBSA and PDAs. Furthermore, as a control experiment, vesicles prepared from the polymerization of the D-glumatic acid terminated diacetylene DTCDA-Glu were selectively responsive to R-TBSA over S-TBSA, indicating the importance of the chiral center in diacetylene to be able to fulfill enantioselective recognition towards chiral molecules.

Figure 5. Pb2+ induced chromatic transition of PDA vesicles. (a) Chemical structures of PCDA and crown ether-modified PCDA-L, and a schematic presentation of the sensing for Pb2+. (b) Photographs of the PDA vesicles in solution (250 mM) upon exposure to increasing concentrations of Pb2+. Adapted with permission from ref 202. Copyright 2015 Elsevier.

There are increasing demands for the fast and accurate sensing of bacteria because of worldwide incidents like food poisoning and bioterrorism alerts. Traditional culture-based method for bacterial sensing are time consuming, usually taking from hours to days. To overcome this problem, numerous work has been published with regard to the fast and chromatic detection of bacteria using PDA vesicles.34, 186, 211-213 Lee and co-workers reported the first example of PDAs acting as a sensor as well as an inhibitor for certain bacteria.214 In this case, imidazolium salt and imidazole were attached to the head groups of diacetylenes to prepare PDA-I-I vesicles (Figure 7a). A distinct blue-to-red color change and fluorescence enhancement of the system was observed upon interaction with various bacterial strains, such as MRSA (methicillin-resistant Staphylococcus aureus) and ESBL-EC (extended-spectrum β-lactamase-producing Escherichia coli) (Figure 7b). Furthermore, antibacterial experiments suggested that these PDA-I-I vesicles showed excellent inhibition towards both Gram-positive and Gram-negative pathogens. TEM and zeta potential analysis suggested that the inhibitive phenomenon was caused by the electrostatic interaction between the imidazolium moieties of PDA-I-I vesicles and the negatively charged bacterial surface. The interaction perturbed the membrane of the bacteria causing cell death (Figure 7c). Anti-infection experiments also showed that after the host cells were infected with bacteria, the addition of PDA-I-I vesicles lowered the numbers of bacterial cells by 88 %, while viability of host cells was not significantly affected, with a less than 10 % decrease in numbers. The result suggested that PDA-I-I vesicles have a great potential to be used as a dual-function kit for certain bacteria.

Figure 6. Chirality sensing of PDAs. (a) Chemical structure of TCDAGlu. (b) Photographs of PDAs in the presence of increasing concentrations of S-TBSA (up) and R-TBSA (down). From left to right: 0, 1 × 10−3, 5 × 10−3, 1 × 10−2, 2.5 × 10−2, 5 × 10−2, 7.5× 10−2, 1.0 × 10−1 M. (c) Absorbance of PDAs at 540 nm (red phase) with increasing concentrations of S-TBSA and R-TBSA. Adapted with permission from ref 210. Copyright 2017 American Chemical Society.

driven by affinochromism. Some ligands that had been explored to bind with Pb2+ specifically to induce the optical change of PDAs include di-(2-picolyl)amine,200 dopamine,201 crown ether,202-203 pentaethylene glycol,204-205 glycine,206 and galloyl group.185 Based on the fact that Pb2+ binds easily with specific crown ethers to form relatively stable coordination complex, Wang et al. synthesized a novel diacetylene monomer PCDA-L which contained 1-aza-18-crown-6ether as the head group (Figure 5a).202 Once co-assembled with PCDA and polymerized under UV light, the vesicles showed a specific color switch from blue to red upon interaction with Pb2+ (Figure 5b), while other metal ions did not induce an obvious color change, indicating excellent selectivity. Except for Pb2+, other metal ions such as mercury (Hg2+)207 and cesium (Cs+)208 had also been detected with a similar mechanism using PDA vesicles. Molecular chirality is ubiquitous in nature and chiral recognition plays a significant role in diverse fields of science and technology. Even though PDAs were extensively used to detect and distinguish numerous analytes, their applications for sensing chiral molecules 8

ACS Paragon Plus Environment

Page 9 of 26 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 9. PDA vesicles for the sensing of LPA. (a) Illustration of the electrostatic and hydrophobic interaction between LPA and iPDA. Photograph shows the color transition of iPDA from blue to red upon interaction with LPA. (b) Colorimetric response of iPDA towards LPA and other main components in blood plasma. (c) Time-dependent response of iPDA towards LPA (8 μM). (d) Schematic presentation of the diagnosis of ovarian cancer using a iPDA-based point-of-care device. Adapted with permission from ref 230. Copyright 2017 American Chemical Society.

Figure 7. PDA vesicles for sensing and inhibiting bacteria. (a) Polymerization of imidazolium salt and imidazole modified diacetylenes into PDA-I-I. (b) UV-vis spectra of PDA-I-I (100 μM) with increasing concentrations of E. coli O157:H7. Inset picture shows the color transition of PDA-I-I (100 μM) upon interaction with E. coli O157:H7 (1 x 107 CFU/mL). (c) Transmission electron microscopy (TEM) images of MRSA (1 x 107 CFU/mL) in the absence/presence of PDA-I-I. The arrow indicates the membrane integrity was disrupted by PDA-I-I vesicles. Reprinted with permission from ref 214. Copyright 2016 Elsevier.

(HA) lectin is anchored. HA binds with the α-glycosides of sialic acid on cell-surface glycoproteins and glycolipids to initiate the viral infection.215 Therefore, the most straightforward method to detect influenza virus is to modify the head groups of diacetylenes with sialic acid, as first illustrated by Charych et al.16 and later many others.216219 Alternatives include grafting influenza virus-specific anti-bodies,220-222 DNA,223 or even peptides224 to the head groups of diacetylenes. A highly specific and sensitive biosensor for the rapid detection of H1N1 influenza virus was developed by Song et al.224 The PEP-PDA vesicles were prepared by the co-assembly of PCDA and NHS-modified PCDA, followed by photopolymerization. A H1N1 influenza virus specific peptide was subsequently conjugated to the vesicles (Figure 8a). Once H1N1 influenza virus was introduced, the PEP-PDA solution turned from blue to red, attributed to the interaction between the virus and the peptides. A H1N1 virus non-specific peptide was introduced as a control study, which did not induce any obvious color switch, indicating the specificity of the peptide (Figure 8b). A detection limit for H1N1 influenza virus with 105 PFU was obtained using this naked-eye detection. Similar sensing strategies were also employed for other viruses, such as foot-andmouth disease virus.225 Biomolecules such as lipids, proteins, nucleic acids and amino acids are closely associated with various molecular interactions under physiological conditions, which macroscopically impact our health. Inappropriate concentrations of such biomolecules in certain parts of our bodies indicate the presence of a medical condition. Therefore, it is rather important to detect such biomolecules in an effective way. PDAs have gained considerable amount of attention for sensing biomolecules, as illustrated in numerous paper.42, 45, 226-229 Ovarian cancer is a worldwide common malignancy among females and the routine early stage diagnostic testing remains difficult. To remediate this challenge, Wang et al. developed a PDA-based biosensor for the rapid, sensitive and quantitative detection of lysophosphatidic acid (LPA), a well-known indicator for progression and metastasis of ovarian cancer tumor.230 By modifying the head group of PCDA with an imidazolium group and subjecting to UV light irradiation, a blue

Figure 8. Peptide-modified PDA vesicles for the detection of H1N1 influenza virus. (a)Schematic illustration of the sensing process of PEPPDA for H1N1 influenza virus. (b) Photographs of the PEP-PDA and control nanosensor after addition of different concentrations of H1N1 influenza virus. Adapted with permission from ref 224. Copyright 2016 The Royal Society of Chemistry. Influenza virus is likely the most common virus that greatly affects our lives. Influenza virus has a lipid bilayer structure where hemagglutinin

9

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 26

iPDA was obtained. It was found that upon exposure to LAP, the blue iPDA vesicles underwent fast chromatic transition from blue to red, with a detection limit of 0.5 μM in the presence of major blood plasma components (Figure 9a, b, c), which is lower than the average plasma LPA concentration of ovarian cancer patients. Zeta potential, TEM, Raman spectrum, and molecular dynamics simulation studies revealed that the electrostatic and hydrophobic interactions between iPDA and LPA induced the blue-to-red phase change synergistically. Furthermore, the level of LPA in plasma samples from mouse xenograft models and patients with ovarian cancer could be quantified using a portable point-of-care device (Figure 9d). Such a device allowed for the naked-eye detection with claimed 100% accuracy, opening up opportunities for early-stage ovarian cancer detection. As illustrated in the examples listed above, PDA vesicles are typically prepared using one or two different diacetylene monomers, one provides the functional head group to interact with the target analyte and the other one is a commercially available diacetylene, which promotes the co-assembly into a well-defined structure for the topochemical polymerization. However, this approach is inherently limiting their applications due to the monotony of the diacetylene monomers. In contrast, Wang et al. used three-component PDA vesicles to detect sialic acid (SA) in solution and fulfilled the in situ imaging of SA on living cells.231 Three elaborately designed diacetylene monomer PCDA-pBA, PCDA-Nap, and PCDA-Ea were synthesized(Figure 10a). Specifically, PCDA-pBA is decorated with a phenylboronic acid head group, which has been reported to have an anomalous binding profile with SA at physiological pH over other sugars. PCDA-Nap was grafted a 1,8-naphthalimide fluorophore as the head group, which could work as a fluorescence signaling moiety.

Figure 11. Detection of sodium benzoate based on a reversed color transition. (a) Schematic illustration of the reversed mechanism for sensing sodium benzoate. (b) Images of the PDA vesicles upon interaction with various concentrations of food additives. (c) The colorimetric response change of the vesicles with increasing concentration of sodium benzoate. Reprinted with permission from ref 232. Copyright 2018 Elsevier.

PCDA-Ea had an amine group at the end, which not only increased the binding affinity of phenylboronic acid with SA but also improved the fluorescence signaling of the system. Due to the energy transfer between the fluorophore on PCDA-Nap and the polymeric backbone of PDAs, the vesicles were non-fluorescent and showed the typical blue color of PDAs. Upon interaction with SA, the fluorescence of the 1,8-naphthalimide fluorophore was regained (Figure 10b) and the color switched from blue to red as a result of the disturbance of the PDA backbone. The potential of this approach was illustrated by the in situ imaging on living human breast adenocarcinoma (MCF-7) cell surface (Figure 10c). Most of the PDA vesicle sensors reported so far are based on the color transition and/or fluorescence enhancement of PDAs from undisturbed phase to disturbed phase upon interaction with external stimuli. In contrast, Li et al. achieved the detection of sodium benzoate, a commonly used food additive, with PDA vesicles employing the reversed mechanism, i.e., the target could promote the system to switch from the disturbed phase to the undisturbed phase (Figure 11a).232 They functionalized diacetylenes with a sodium benzoate analog 4-aminobenzioic acid as the hapten. In the presence of sodium benzoate antibodies, the system became red as a result of the interaction between hapten and the antibodies. In contrast to other food additives, with increasing concentration of sodium benzoate added, the antibodies interacted with sodium benzoate, thus the interaction between hapten and antibodies was inhibited, leading to the recovery of blue color (Figure 11b, c). This study suggested a new strategy to detect the target analyte, which could substantially broaden the design opportunities of new PDA sensors. Substrate-based sensors

Figure 10. Three-component PDAs for the detection of SA and cell imaging. (a) Chemical structures of the three diacetylene monomers used in this study: PCDA-EA, PCDA-pBA and PCDA-Nap, and the sensing mechanism for SA. (b) Emission spectra of the PDAs (100 μM) with increasing concentrations of SA in phosphate-buffered saline (PBS) buffer solution (10 mM, pH = 7.4). (c) Overlaid fluorescence and bright-field confocal images of MCF-7 cells after incubation with the PDA vesicles. Scale bar = 20 μm. Adapted with permission from ref 231. Copyright 2018 The Royal Society of Chemistry.

Immobilizing PDAs in a solid substrate is a common strategy to overcome the shortcomings of solution-based PDA sensors (e.g., low sensitivity due to the homogeneous dilution of vesicles and targets in solutions, poor portability, instability because of the intrinsic aggregation of the nanoparticles). The embedment of PDAs into matrices has been employed extensively for the fabrication of materials with well-defined nano/micro-scale structures and unique chromatic properties. Furthermore, it has been shown that the 10

ACS Paragon Plus Environment

Page 11 of 26 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials but only recently the first example of hydrochromic PDA was reported.246 However, the PDAs in that paper were sensitive to ambient moisture and hydrochromism did not occur when it was printed on paper, which restricted its long-term storage and applications. Park et al. improved the performance of such hydrochromic PDAs by introducing an imidazolium ion into the head groups of diacetylenes, making them compatible with inkjet printing (Figure 12a).247 Moreover, it was found that an instant blue-to-red colorimetric response took place when the patterned paper was exposed to water (Figure 12b), accompanied with a fluorescence turn-on. The same phenomenon was observed upon contact with a human palm, induced by the tiny amount of moisture secreted from a sweat pore (Figure 12c). Since some sweat pores are active and some are not, a precise map of active sweat pores on a palm was obtained by carefully examining each pore on the paper and its color change. Hydrogel Hydrogels are 3D cross-linked hydrophilic polymer networks that are capable of swelling and de-swelling reversibly in water. The 3D network, as compared to 1D or planar substrate, possesses higher stability, increased surface area, and multipoint interaction sites, which could greatly improve the sensing performance of the encapsulated sensor with amplified output signal and a lower detection limit. Hydrogels also show a degree of flexibility similar to natural tissue due to their large water content. Besides, hydrogels are usually non-toxic and bio-degradable, which are important features to fabricate sensors to be used for bio-purposes.

Figure 12. Paper-based hydrochromic PDA sensor. (a) Chemical structure of the imidazolium-modified diacetylene, and the color transition of the corresponding PDA from blue to red upon interaction with water. (b) Inkjet-printed pattern of the hydrochromic PDA on paper. (c) Fluorescence image of the sweat pores on a palm (left) and an amplified image of the boxed area on the left (right). Adapted with permission from ref 247. Copyright 2016 John Wiley and Sons.

proper incorporation not only affected polymerization behavior of the diacetylenes but also lead to an enhancement of the sensitivity of PDAs towards selective target analytes. Since the early reports of PDA-supporting matrices such as glass,63, 233 silica,65, 234-239 agarose,66 alginate fibers,67 layered double hydroxide nanosheets,64 various others were explored. Recent examples are discussed in the following section. Paper Paper is one of the most common materials that we use every day, and it has been employed to immobilize PDAs for multiple purposes. Except for its well-known advantages such as low-cost, light weight, flexibility, and disposability, which are advantageous for point-ofcare and in-field use, a superior feature of using paper as the substrate for sensing is that PDA sensors can be deposited onto paper by inkjet printing technique.240 This allows for the large-scale fabrication of desired patterns in an inexpensive manner, greatly broadening the potential applications of PDA sensors. The first report of inkjetprinted PDAs was in 2011,241 which illustrated that the chromatic properties of PDAs retained even after mixing with a non-ionic surfactant and being printed onto paper. Ever since then, inkjet printing was widely employed to fabricate paper-based PDA sensors.117, 242 Except for inkjet printing, other techniques such as drop-casting,115, 243 screen-printing,125 and wax-screen-printing244 had also been utilized to immobilize PDAs onto paper.

Figure 13. 3D printed PDA/hydrogel composites for detoxification. (a) Schematic illustration of the interaction between PDA vesicles and toxins. Photograph shows the centrifuged red blood cells after incubation with normal saline (control) or melittin (5 μg/mL) together with PDA vesicles in different concentrations. (b) Time-dependent fluorescence images of PEGDA hydrogel after incubation with a melittin solution (50 μg/mL) in the absence (up) or presence (down) of PDA vesicles. (c) SEM image showing the microscopic morphology of the 3D printed detoxifier. Scale bar = 50 μM. (d) Neutralized efficiency of the 3D printed detoxifier. An equivalent amount of PDA vesicles was added for comparison. Adapted from ref 248. Copyright 2014 Springer Nature.

Materials that undergo chromatic changes upon interaction with water (hydrochromism) have been widely explored as humidity sensors for applications like monitoring the water content in organic solvents, rewritable printing on paper and human sweat pore mapping.245 PDAs were designed to be responsive to numerous analytes

11

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 26

Figure 14. Membrane-supported PDA sensors. (a) i) Chemical structures of the five diacetylenes used to incorporate into PEO. ii) Histogram of the fluorescence emission profiles of the three PDA/PEO electrospun membranes against four organic solvents vapors (tetrahydrofuran, chloroform, methanol, and hexane) at room temperature for 1 h. iii) Optical microscope image of the polymerized PCDA-biotin membrane upon exposure to different amounts of STA (from top to bottom: 500 ng, 250 ng, 0 ng). iv) Images of the three PDA/PEO electrospun membranes upon interaction with different concentrations of hexylamine. (b) i) Chemical structures of C18-EDEA-T and PCDA, and a schematic illustration of the Hg2+-induced aggregation of the PDA vesicles. ii) Photographs of the membranes after filtration of the PDA vesicles with increasing concentrations of Hg2+. Panel a adapted with permission from ref 263. Copyright 2014 American Chemical Society. Panel b reprinted with permission from ref 187. Copyright 2014 The Royal Society of Chemistry.

spin-coating,114, 246, 255-257 mixing-drying,111, 132, 161, 258-260 and dropcasting.119, 184, 261-262 254

Gou et al. used 3D printing technology to install functional PDA particles into a hydrogel matrix, and the PDA/hydrogel composite was used to attract, capture and sense a pore-forming toxin melittin, which could be employed as a detoxification device.248 Experimental and computational simulations revealed that the hydrophobic and electrostatic interactions between PDAs and melittin lead to the capture of melittin by PDA vesicles, which could be monitored by the visual disappear of the red color of the supernatant (Figure 13a) and the fluorescence enhance of the precipitate. By incorporating PDAs into a 3D printed poly(ethylene glycol) diacrylate (PEGDA) hydrogel, the PDA/hydrogel composite allowed a faster diffusion of melittin into the 3D matrix and the capture of melittin became more efficient (Figure 13b). Inspired by the fact that the structure of liver contributes to fast substances’ exchange between blood stream and hepatic cells, the hydrogel was fabricated into a liver-mimetic struc-

In this context, an interesting example was reported by Davis et al.263 They synthesized a series of diacetylenes with different structures (Figure 14ai) and embedded them into poly(ethylene oxide) (PEO) to generate PDA/PEO nanofibers. The successful differentiation of four different organic solvents was achieved on the basis of the solvent-dependent fluorescence transition of the nanofibers (Figure 14aii). When biotin was introduced as a head group, the PDA/PEO nanofibers went through a clear blue-to-red color transition after streptavidin (STA) was added based on the biotin-STA interaction (Figure 14aiii), accompanied with a fluorescence enhancement, which could be used as a dual-mode sensor for STA. Finally, the determination and identification of 8 different amines using principal component analysis (PCA) was possible because of the different color patterns of PDAs upon interaction with different concentrations of amines (Figure 14aiv), indicating the multi-stimuli responsiveness of these PDA/PEO nanofiber.

ture with modified liver lobule configuration (Figure 13c). Such a porous structure facilitated the fast diffusion of melittin into the 3D matrix, which led to a superior detoxification performance of the hydrogel (Figure 13d). Results showed that the toxin-containing solution completely lost virulence in damaging cellular membrane after treatment with the hydrogel. This report provided an inspiring strategy for designing novel nanoparticle-based detoxification treatments.

Ma et al. reported a selective colorimetric sensor for naked eye detection of Hg2+ based on the unique coordination between thymine and Hg2+, and a simple filtration process using a cellulose acetate membrane.187 The co-assembly of PCDA and thymine-containing probe C18-EDEA-T yielded in functionalized PDA vesicles exposed thymine groups on the external surface. Upon interaction with Hg2+, the coordination led to the aggregation of the PDA vesicles (Figure 14bi). Using a facile filtration method, the aggregated particles stayed on top of the membrane due to their bigger sizes, which appeared as blue (Figure 14bii), while the colorless, single vesicles washed off. The visual detection limit for Hg2+ using this filtration method was as low as 0.1 μM.

Membrane Membranes are a class of materials that exhibits good mechanical flexibility, high thermal stability and good chemical resistance. Most importantly, the porous nature endows it with high specific surface area, which enables the fast diffusion of molecules in between. This feature made membrane as an ideal solid matrix to incorporate PDAs with improved sensitivity. PDA-embedded membranes were fabricated with various techniques, such as electrospinning,113, 116, 204, 206, 24912

ACS Paragon Plus Environment

Page 13 of 26 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Microbeads Microbeads are micrometer-sized particles that have many advantages over planar substrates with regard to multiplex detection and high throughput screening systems. The curved surfaces of microbeads increase the surface area for molecular attachment and thus enhance the sensing sensitivity. Also, the collective volumetric effect from microbeads promotes a better visual detection of analytes. Given the many advantages, microbeads embedded with PDAs have been actively explored for sensing applications185, 264-268 since the first report by Nie et al. in 2006.269 To obtain PDA-embedded microbeads with controllable and uniform sizes, microfluidic techniques have been frequently utilized.162, 201, 270 Figure 15. Detection of Pb2+ using PDA embedded microbeads. (a) i) Optical (blue phase, left) and fluorescent (red phase, right) microscope images of alginate hydrogel microbeads. ii) Plot of the relative fluorescence intensity of the PDA embedded microbeads as a function of added Pb2+ concentration. Inset shows fluorescence images of the microbeads upon incubation with different concentrations of Pb2+ for 1 h. (b) i) Optical (blue phase, left) and fluorescent (red phase, right) images of the Janus microbeads. Scale bar = 500 μm. ii) Collecting the Janus microbeads by employing magnetic field. Panel a reprinted with permission from ref 201. Copyright 2015 The Royal Society of Chemistry. Panel b reprinted with permission from ref 185. Copyright 2014 American Chemical Society.

Wang and co-workers modified diacetylenes with dopamine as the head group.201 Using a droplet microfluidic protocol, alginate hydrogel microbeads embedded with PDAs were obtained for the detection of Pb2+ on the basis of the coordination between dopamine and Pb2+ (Figure 15ai). As compared to the PDA vesicles in solution, the PDA microbeads showed a much higher sensitivity towards Pb2+ with excellent stability and a detection limit as low as 200 ppb compared to 10 μM in solution (Figure 15aii). Kang et al. reported the detection of Pb2+ using Janus-shaped PDA microbeads based on a similar coordination-induced sensing mechanism (Figure 15bi).185 More impressively, by introducing magnetic Fe3O4 nanoparticles into the microbeads, the system was endowed with additional features, such as convenient collection of the particles by applying magnetic field and easy manipulation (Figure 15bii). Barium ions (Ba2+),268 hydrocarbons,162 phosphinothricin acetyltransferase protein,270 and streptavidin266 were also detected using PDA microbeads.

nanoparticles at either high pH or low pH. Such PDA/ZnO nanocomposites were not only found responsive to pH, but also to temperature,274 organic acids,275 and organic amines.276 A wearable wrist strap sensor based on a PDA/molybdenum disulfide (MoS2) nanocomposite for the selective detection of N,N-dimethylformamide (DMF) vapor was fabricated by Wang et al.277 The incorporation of MoS2 into PDAs played dual roles in this case. On one hand, it supported the PDAs to enhance the porosity and the transparency of the film originating from its nano-flake morphology. On the other hand, its chelation ability towards atoms with lone electron pairs made the system more sensitive to DMF vapor than other organic vapors. Using a simple spin-coating method, the PDA/MoS2 nanocomposite was immobilized onto a polyethylene terephthalate (PET) film and a wearable wrist strap sensor was fabricated showing high transparency, sensitivity, flexibility and reliability. As the concentration of DMF vapor increased from 0.01% to 4%, a color change from blue to red could be clearly discerned on the wrist band, as confirmed in the UV-vis spectra (Figure 16ci, cii). This method offers a promising opportunity for the development of smart wearable sensing devices.

Other substrates In addition to the above mentioned substrates, a variety of other substrates were also utilized to immobilize PDAs. For instance, by immobilizing PDA vesicles into a 3D micropillar-structured CNTbased network (Figure 16ai, aii),271 as compared to a two-dimensional (2D) network in which PDA vesicles were immobilized on a silicon wafer, the sensitivity towards α-cyclodextrin (α-CD) was significantly enhanced, as the detection limit for α-CD decreased from 2.5 mM to 2.0 μM. It is presumed that the enhancement was caused by the increase of surface area of the 3D network, since a larger surface area means that more α-CD molecules interacted with PDAs through the large free volume existing in the sensor matrix, thus, ensuring a higher sensitivity. The color transition behavior of PDAs upon various stimuli could be altered by the incorporation of inorganic metallic materials, for example, zinc oxide (ZnO).147-149, 272 The strong hydrogen bonding and ionic interaction between PDAs and ZnO is responsible for the alteration. Amornsak et al. incorporated three diacetylene monomers TCDA, PCDA, and 5,7-hexadecadiynoic acid (HDDA) into ZnO nanoparticles and investigated the color transition behavior of the PDA/ZnO nanocomposites upon exposure to pH changes.273 As compared to PDA vesicles, the nanocomposites exhibited different color transition patterns towards pH changes (Figure 16b). For example, when decreasing the pH from neutral to acidic, all the nanocomposites changed from blue to red, while the PDA vesicles remained unchanged. When increasing the pH from neutral to basic, all of them showed clear color transition, but the color transition for the nanocomposites occurred at higher pH values. The proposed mechanism for this impact is attributed to the dissolution of ZnO

Microfluidic technology has been integrated with the preparation of PDA vesicles278 and the label-free detection of analytes owing to the fluorescence turn-on feature of PDA upon interaction with external stimulations279. One of the advantages of using this technology to prepare nanoparticles is its controllability, namely generating a more monodisperse system.280-281 The conventional hydration method inevitably results in poly-disperse particles since it is difficult to maintain a consistent mixing environment. The size of the nanoparticles affects the optical properties and the sensing performances of PDAs.278, 282-283 Whereas the label-free microfluidic sensors have several advantages over conventional bulk sensors, such as the consumption of nascent samples, large interfacial area, and rather short molecular diffusion distance, which accelerates the phase transition of PDAs upon interaction with analytes from non-fluorescent to fluorescent. Jang and co-workers developed a fish gill-like surface-modified microfluidic system for the specific detection of Staphylococcus 13

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 26

Figure 16. PDAs on various other substrates. (a) i) SEM image of the 3D micropillar-structured CNT network after Al2O3 was deposited. ii) Fluorescence microscopy images of the PDA-encapsulated CNT after exposure to increasing concentrations of α-CD. (b) Photographs of aqueous suspensions of different PDA/ZnO nanocomposites taken at different pH. (c) i) UV-vis spectra of the PDA/MoS2 film upon exposure to different concentrations of DMF. ii) Photographs of the flexible transparent wrist wrap upon exposure to DMF vapor. (d) Fluorescence images comparison of the channels without LST (left) and with LST(right) after flow of S. epidermidis. Dotted lines show the boundary of the channels, and arrows indicate the direction of bacterial flow. Scale bar = 500 μM. (e) i) Schematic presentation of the preparation of the PDA-embedded pen. ii) Colorimetric transition of the hand-painted images upon temperature treatments. (f) Photographs of the PDA-embedded Luria-Bertani-agar plate after incubation with different bacterial strains for 16 h. Panel a reprinted with permission from ref 271. Copyright 2015 The Royal Society of Chemistry. Panel b reprinted with permission from ref 273. Copyright 2014 Elsevier. Panel c reprinted with permission from ref 277. Copyright 2016 The Royal Society of Chemistry. Panel d reprinted with permission from ref 284. Copyright 2017 Springer Nature. Panel e reprinted with permission from ref 285. Copyright 2016 John Wiley and Sons. Panel f reprinted with permission from ref 286. Copyright 2016 The Royal Society of Chemistry.

pathogens based on the enzymatic reaction between lysostaphin (LST) and Staphylococcus.284 By functionalizing the microfluidic channels’ surface with amine groups, NHS-functionalized diacetylene monomers were immobilized on the channel surface via a typical EDC/NHS coupling reaction. As an amine-rich enzyme, LST was attached to the diacetylene monomers using the same coupling reaction. Upon UV light irradiation, a PDA/LST conjugated microfluidic channel was thus constructed. Due to the fact that LST cleaves the pentaglycine cross-bridge in the cell wall of certain Staphylococcus, once LST interacted with Staphylococcus, the interruption of the side chain led to the distinct fluorescence change of the system from non-fluorescent to red fluorescence (Figure 16d). It is noteworthy that while functioning as a chromatic sensor for Staphylococcus, the system also provides antibacterial activity since the cleavage of pentaglycine in the cell walls leads to the death of the bacteria.

(Figure 16eii). Moreover, the thermochromic transition temperature was dependent on the melting point of the wax. Therefore, it is plausible to control the transition temperature by varying the different kinds of wax, which might have potential applications in developing directly writable techniques. Park and co-workers modified diacetylene monomers with two different amine groups at the head group, and the resulting PDAs were incorporated into Luria-Bertani-agar.286 It was found that the agar showed a specific color change towards the bacteria strain Bacillus subtilis NCIB3610 rather than Bacillus subtilis SSB466 and Pseudomonas aeruginosa 14 (Figure 16f). The reason for this selectivity was due to the selective release of surfactin. The color change was induced by the specific interaction between surfactin and the amine-functionalized PDAs, as revealed by computational calculations. Other forms of PDA sensors

Chae et al. incorporated diacetylenes into paraffin wax using a facile mixing-molding method (Figure 16ei).285 The self-assembling behavior of diacetylenes were not affected by the large amount of hydrophobic wax molecules, thus creating a crayon-like PDA-wax composite pen. Such a pen could be deposited on a solid substrate, for example paper, which enables the fabrication of hand-writable patterned images. Such images on paper retained the chromatic properties of PDAs as they displayed thermochromic properties

Except for solution and substrate-based PDA sensors, in recent years, many other sensor types were fabricated on the basis of manipulating the self-assembly of diacetylenes. Hu and co-workers fabricated a PDA hollow microtube through a novel hierarchical self-assembly method, which could be used as an optical waveguide.287 The intensity of the out-coupled emission light and the waveguide performance of the PDA microtube could easily be affected by

14

ACS Paragon Plus Environment

Page 15 of 26 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials In addition to the above-mentioned examples, in which the color transition is induced by non-covalent interactions, the perturbation of PDA backbone is also achievable when the head groups of PDAs react chemically with the analytes. Requirements for this approach include fast chemical reactions, and mild conditions without extra heating. Successful examples include Cu2+ detection using click– chemistry,294 pH-induced deprotonation and protonation of the hydrazide head groups295 or tertiary amine,253 and CO2 detection based on amine-carbamate reaction.296-297 Lu et al. fulfilled the detection of hydrogen peroxide (H2O2) in an unusual way.298 It was found that in the presence of horse radish peroxide (HRP), the hydroxyl radicals generated from H2O2 could facilitate the polymerization of diacetylene vesicles into PDA. A clear color transition from colorless to blue was observed visually. When the head group of the diacetylene was modified with a phenylboronic acid, the system became more sensitive to H2O2 with a much lower detection limit.299 Kim et al. used this mechanism to detect glucose in solution with a detection limit down to 1 mM.300

LINEAR DIACETYLENE WITH TWO HEAD GROUPS OR TWO HYDROPHOBIC TAILS AT BOTH ENDS Figure 17. 1D PDA sensors. (a) i) Fluorescence spectra of the PDA microtubes upon increasing concentrations of TNP. ii) Schematic illustration of the PDA-based microtube waveguide sensor for the reversible detection of TNP. (b) i) Schematic illustration of the preparation of a free-standing PDA microtube. ii) SEM image of the PDA microtubes on a silicon substrate. iii) Optical (left) and fluorescence (right) images of a biotin-containing PDA microtube after interaction with a streptavidin solution (100 ng/mL). Scale bar = 8 μM. Panel a reprinted with permission from ref 288. Copyright 2014 The Royal Society of Chemistry. Panel b reprinted with permission from ref 293. Copyright 2016 American Chemical Society.

In contrast to diacetylenes with only one polar head group, diacetylenes modified at both ends are less studied partially because of the synthetic complexity. The obvious advantage of PDAs modified at both ends is that the non-covalent interactions between the side chains are strengthened at both ends, which not only promotes the facile self-assembly of the diacetylenes22, 301-303 but also improves the sensing performance of PDAs,304 in particular, reversible thermochromism, which is greatly determined by the strength of the intramolecular interactions among the side chains, as illustrated by Phollookin et al.,122 Ampornpun et al.,305 Lee et al.,131 and Yoon et al. 306 A recent example was reported by the group of Juyoung Yoon. They synthesized a diacetylene bis-PCDA-Ph, which had two diacetylenes tethered via an intervening p-phenylene group (Figure 18ai).307 The polymer bis-PDA-Ph exhibited excellent thermochromic reversibility between 20 to 100 °C in solution and 20 to 120 °C in PEO polymer matrix, even though there were no head groups that were capable of forming hydrogen bonds. Theoretical simulations revealed that the π-π interactions between the benzene rings along the rigid ene-yne backbone mainly promoted the reversibility, in agreement with other reports showing that the thermochromic reversibility of PDAs is largely dependent on the noncovalent intermolecular interactions. More recently, they elucidated the origins of the reversible thermochromism of bis-PDA-Ph in terms of electronic relaxation dynamics using UV-vis spectroscopy and frequency-resolved transient absorption (FRTA) spectroscopy.308 As compared to PCDA-PDA, which was derived from PCDA exhibiting reversible thermochromic behavior only below 60 °C, it was found that the existence of an intermediate state in the electronic relaxation was the direct evidence for the reversible thermochromic property of bis-PDA-Ph. As temperature increased, both PDAs showed a common intermediate state as an electronic structural marker for the recovery of their initial structures when the temperature decreased. However, if the temperature rises above 80°C, significant structural changes are induced, which led to the alteration of the electronic energy structure. As a result, the intermediate state of PCDA-PDA no longer existed while it remained for bis-PDA-Ph (Figure 18aii).

external perturbations due to the stimuli-responsive nature of PDAs. On the basis of such an optical waveguide, the highly selective and reproducible detection for picric acid (TNP) in aqueous media was accomplished.288 The strong complex interaction between TNP and the amine groups on the PDA microtubes facilitated the electron transfer from PDA to electron-deficient TNP, which resulted in the fluorescence quenching of the PDA microtubes (Figure 17ai). TNP captured on the waveguide could be removed by washing with ethanol, thus ensuring the reversibility of the waveguide (Figure 17aii). Further, the amine groups stretching out on the microtubes can be tailor-made to be sensitive to other analytes, such as miRNA-21,289 miRNA-215,290 light,291 and polarized light.292 In another report, Oh et al. presented a novel meniscus guided method to fabricate free-standing PDA microtubes.293 A single PDA wire with a tubular structure was feasibly fabricated by pulling of a micropipette containing a DMF solution of PCDA, followed by subsequent polymerization (Figure 17bi, bii). The sensing ability of such a PDA microtube was evaluated using biotinylated diacetylenes for the detection of streptavidin (Figure 17biii). Results showed that the sensitivity for streptavidin was about 4 orders of magnitude greater as compared to a traditional solution-based sensing platform, indicating excellent sensitivity. Also, this micro-tubular sensor only required a tiny volume of analyte (70-140 nL) because the microcapillary tube could act as a reservoir for the analyte, which is envisioned to be an advantageous feature. Chemical reactions between analyte and head group 15

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 26

Figure 18. Linear diacetylenes with two polar head groups and their corresponding PDAs. (a) i) Illustration of the formation of bis-PDA-Ph. ii) FRTA signals measured with bis-PDA-Ph and PCDA-PDA at 20 and 80 °C, respectively. The red circle indicates the intermediate state. (b) i) Chemical structures of diacetylene 1a, 1b and 1c. ii) Photographs of PDA 2a, 2b and 2c upon heating and cooling. (c) i) Chemical structure of diacetylene DGA. ii) Photographs showing the physical and color changes of hydrogels after being dipped in solutions of amino acids: aspartic acid (Asp), glutamic acid (Glu), lysine (Lys), histidine (His) and arginine (Arg) for 12 hours. iii) Photographs showing the physical and color changes of hydrogels after dipping in solutions of nucleobases: cytosine (C), thymine (T), adenine (A) and guanine (G) for 12 hours. ai) adapted with permission from ref 307. Copyright 2014 John Wiley and Sons. aii) adapted with permission from ref 308. Copyright 2016 American Chemical Society. Panel b adapted from ref 137. Copyright 2017 Elsevier. Panel c reprinted with permission from ref 309. Copyright 2017 The Royal Society of Chemistry.

Consequently, the color of PDAs cannot switch back as the temperature deceases, i.e., thermochromism becomes irreversible. The results are relevant beyond the fundamental understanding of PDAs but they can also serve as instructions when designing diacetylenes with specific thermochromic reversibility/irreversibility.

NON-LINEAR DIACETYLENES PDAs derived from non-linear diacetylenes are scarce and their applications for sensing are even rarer. Access to such diacetylenes requires more advanced synthesis, which likely limited their wider consideration. As a matter of fact, the few reports on this type of diacetylenes have discovered intriguing properties of PDAs, which are likely due to the different topological structures of diacetylenes that significantly affect their self-assembly behavior and the optical properties of PDAs.

Huo and co-workers synthesized three diacetylenes 1a, 1b and 1c with amino acid segment as linkers but with different head groups (Figure 18bi), and their corresponding polymers 2a, 2b and 2c were subsequently obtained.137 All three PDAs showed reversible thermochromism as a result of the hydrogen bonds among the amino acid linkers and the π-π interaction among the pyrenes. Especially, PDA 2c with a pyrene unit on both ends exhibited reversible color transition from 25 to 300 °C (Figure 18bii), which is as far as we know, the largest temperature range ever achieved of PDAs. Moreover, the increased temperature range clearly supported the theory that stronger intramolecular interactions leads to better thermochromic reversibility.

Macrocyclic diacetylenes are intriguing due to their structural novelty. Once they are pre-organized into an alignment that meets the structural requirements for the efficient topochemical polymerization, macrocycles could undergo polymerization upon heating or UV irradiation to form columnar structured PDAs.310-314 While these structures offer interesting opportunities to address fundamental questions such as the detailed polymerization process of diacetylenes, to the best of our knowledge, only one example was utilized for sensing applications so far. Heo et al. synthesized a series of macrocyclic diacetylenes MCDA 1-5.315 The 5 diacetylene monomers are similar structurally, with the only difference of the carbon chain length between the butadiyne and the ether linker (Figure 19ai). Monomers MCDA 1-4 underwent the typical 1,4-addition to form tubular-structured conjugated polymers poly(MCDA 1-4) via topochemical polymerization, while MCDA 5 failed to polymerize, indicating the significant influence of the slight structural variation. Furthermore, the four polymers poly(MCDA 1-4) showed different thermochromic (Figure 19aii) and solvatochromic (Figure 19aiii)

A bola-shaped diacetylene DGA with L-glutamic acid at both ends and a diacetylene moiety as the spacer was synthesized by Meng and co-workers (Figure 18ci).309 The amphiphile DGA could form a hydrogel in water upon a heating-and-cooling process. Microscopic studies revealed a nano-helix morphology of the hydrogel after polymerization with 254 nm UV light. These hydrogels exhibited selective shrinkage and chromatic change upon recognition of amino acids and nucleobases, such as glutamic acid, aspartic acid, lysine, histidine and cytosine (Figure 18cii, ciii). This phenomenon was attributed to the weakening of the interaction between the head group L-glutamic acid and the analytes via electrostatic interaction and hydrogen bonding. 16

ACS Paragon Plus Environment

Page 17 of 26 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 19. PDAs derived from non-linear diacetylenes and their sensing applications. (a) i) Chemical structures of macrocyclic diacetylenes MCDA 1-5 and the schematic illustration of the formation of the columnar PDAs. ii) UV-vis spectra of polymerized MCDA-3 upon heating and cooling. Inset: photographs of the polymer crystals at the designated temperature. iii) UV-vis spectra of polymerized MCDA-3 upon exposure to different aromatic solvents. Inset: photographs of the PDAs in different solvents. (b) i) A schematic representation of the co-assembly of diacetylene-modified calix[4]arene with PCDA. ii) Schematic representation of the colorimetric response of PCDA-5 towards Dy3+. Inset: color change and precipitation of PCDA-5 vesicles after the addition of Dy3+. (c) i) Chemical structures of monomers 1, 2 and 5. ii) UV-vis spectra of the PDA solution derived from monomer 5 upon interaction with different amounts of Con A. Inset: color transition of the solution after interaction with Con A. Panel a reprinted with permission from ref 315. Copyright 2017 American Chemical Society. Panel b reprinted with permission from ref 318. Copyright 2018 The Royal Society of Chemistry. Panel c adapted with permission from ref 325 and ref 326. Copyright 2016 John Wiley and Sons. Copyright 2017 John Wiley and Sons.

properties, which could be utilized to differentiate isomeric xylenes based on the color change.

ness to pH. When the carboxylic group of the garlic acid was replaced with poly(alkyl aryl ether) dendrons to obtained monomer 2, the reversible thermochromism was further improved to 300 °C, as a result of enhanced interactions provided by the dendrons.326 Alternatively, a linker was used to attach two α-D-mannopyranoside moieties at the position of the carboxylic group of the gallic acid to obtain monomer 5 that can be used for sensing, i.e., to assess the carbohydrate-protein interaction between sugar and lectin via chromic changes. It was found that the addition of the lectin concanavalin A (Con A) to PDA vesicles made from these monomers did not induce a color change of the system, even though DLS and AFM clearly showed an increase in vesicles’ sizes. However, if the polymerization of the monomers happened in the presence of Con A, the solution exhibited a distinct pink color, which was attributed to the ligand-lectin interaction (Figure 19cii).

Another group of non-linear diacetylenes is macrocycle-based, in which butadiynes are incorporated into the molecule as an appendix of the macrocycle, such as calixarene. Burilov et al. have conducted extensive work regarding the synthesis and applications of calix[4]arene macrocycles modified with diacetylene moieties.316-317 They successfully incorporated two diacetylenes into a p-tert-butylthiacalix[4]arene and two amphiphilic monomers were obtained (Figure 19bi).318 It is well known that due to its unique 3D structure, calix[4]arene is able to bind metal ions to form stable host-guest complex. Therefore it is anticipated that the PDAs that contain calix[4]arene are responsive to metal ions since the host-guest complexation distorts the cavity of calix[4]arene provoking the perturbation of the backbone of PDA. Indeed, upon co-assembly with PCDA, the system showed significant colorimetric response towards lanthanide ions with a detection limit of down to 8 µM, depending on the size of the metal ions (Figure 19bii).

CONCLUDING REMARKS

It is usually anticipated that anchoring multiple butadiyne moieties into one scaffold might enhance the stability 319-324 of the corresponding PDAs and endows the system with enhanced thermochromic properties,138 as a result of the strengthened interactions among the side chains. Singh et al. prepared a hyperbranched diacetylene monomer 1 with three diacetylenes covalently linked to a gallic acid (Figure 19ci). It was found that the PDA derived from this monomer showed excellent reversible colorimetric response to temperature in solution (20 to 95 °C), and in thin film (20 to 220 °C),325 accompanied with responsive-

Ever since the pioneering work conducted by Wegner in 1969, PDAs have been extensively explored as sensing materials due to their intriguing naked-eye observable color switch and fluorescence enhancement upon interaction with various stimulations. In this review, we discussed the most recent developments in PDA-based sensors focusing on the past five years. The progress made so far is encouraging, but challenges still exist. The underlying mechanism for the chromatic transition upon stimulation remains unclear. The sensitivity of solution-based PDA sensors often remains unsatisfactory. The reversibility of PDA sensors still remains a problem considering the fact that many of them showed excellent thermal, pH, or UV17

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 26

(6) Guo, Z.; Song, N. R.; Moon, J. H.; Kim, M.; Jun, E. J.; Choi, J.; Lee, J. Y.; Bielawski, C. W.; Sessler, J. L.; Yoon, J. A benzobisimidazolium-based fluorescent and colorimetric chemosensor for CO2. J. Am. Chem. Soc. 2012, 134, 1784617849. (7) Yang, Y. K.; Tae, J. Acridinium salt based fluorescent and colorimetric chemosensor for the detection of cyanide in water. Org. Lett. 2006, 8, 5721-5723. (8) Aldewachi, H.; Chalati, T.; Woodroofe, M. N.; Bricklebank, N.; Sharrack, B.; Gardiner, P. Gold nanoparticle-based colorimetric biosensors. Nanoscale 2017, 10, 18-33. (9) Bigdeli, A.; Ghasemi, F.; Golmohammadi, H.; Abbasi-Moayed, S.; Nejad, M. A. F.; Fahimi-Kashani, N.; Jafarinejad, S.; Shahrajabian, M.; Hormozi-Nezhad, M. R. Nanoparticle-based optical sensor arrays. Nanoscale 2017, 9, 16546-16563. (10) Kim, H. N.; Guo, Z.; Zhu, W.; Yoon, J.; Tian, H. Recent progress on polymer-based fluorescent and colorimetric chemosensors. Chem. Soc. Rev. 2011, 40, 79-93. (11) Li, C.; Numata, M.; Takeuchi, M.; Shinkai, S. A sensitive colorimetric and fluorescent probe based on a polythiophene derivative for the detection of ATP. Angew. Chem. Int. Ed. 2005, 44, 6371-6374. (12) Wegner, G. Topochemische Reaktionen von Monomeren mit konjug ierten Dreifachbindungen. Z. Naturforsch. B 1969, 24, 824-832. (13) Sheth, S. R.; Leckband, D. E. Direct Force Measurements of Polymerization-Dependent Changes in the Properties of Diacetylene Films. Langmuir 1997, 13, 5652-5662. (14) Carpick, R. W.; Sasaki, D. Y.; Marcus, M. S.; Eriksson, M. A.; Burns, A. R. Polydiacetylene films: a review of recent investigations into chromogenic transitions and nanomechanical properties. J. Phys.-Condes. Matter 2004, 16, R679R697. (15) Jelinek, R.; Ritenberg, M. Polydiacetylenes – recent molecular advances and applications. RSC Adv. 2013, 3, 21192–21201. (16) Charych, D.; Nagy, J.; Spevak, W.; Bednarski, M. Direct colorimetric detection of a receptor-ligand interaction by a polymerized bilayer assembly. Science 1993, 261, 585-588. (17) Chance, R. R. Chromism in Polydiacetylene Solutions and Crystals. Macromolecules 1980, 13, 396-398. (18) Chance, R. R.; Baughman, R. H.; Muller, H.; Eckhardt, C. J. Thermochromism in a Polydiacetylene Crystal. J. Chem. Phys. 1977, 67, 3616-3618. (19) Nallicheri, R. A.; Rubner, M. F. Investigations of the mechanochromic behavior of poly(urethane-diacetylene) segmented copolymers. Macromolecules 1991, 24, 517-525. (20) Chen, X.; Hong, L.; You, X.; Wang, Y.; Zou, G.; Su, W.; Zhang, Q. Photocontrolled molecular recognition of alpha-cyclodextrin with azobenzene containing polydiacetylene vesicles. Chem. Commun. 2009, 1356-1358. (21) Cheng, Q.; Stevens, R. C. Charge-Induced Chromatic Transition of Amino Acid-Derivatized Polydiacetylene Liposomes. Langmuir 1998, 14, 1974-1976. (22) Song, J.; Cheng, Q.; Kopta, S.; Stevens, R. C. Modulating Artificial Membrane Morphology: pH-Induced Chromatic Transition and Nanostructural Transformation of a Bolaamphiphilic Conjugated Polymer from Blue Helical Ribbons to Red Nanofibers. J. Am. Chem. Soc. 2001, 123, 3205-3213. (23) Kew, S. J.; Hall, E. A. pH response of carboxy-terminated colorimetric polydiacetylene vesicles. Anal. Chem. 2006, 78, 2231-2238. (24) Kolusheva, S.; Shahal, T.; Jelinek, R. Cation-Selective Color Sensors Composed of Ionophore−Phospholipid−Polydiacetylene Mixed Vesicles. J. Am. Chem. Soc. 2000, 122, 776-780. (25) Lee, J.; Kim, H. J.; Kim, J. Polydiacetylene liposome arrays for selective potassium detection. J. Am. Chem. Soc. 2008, 130, 5010-5011. (26) Lee, J.; Jun, H.; Kim, J. Polydiacetylene-Liposome Microarrays for Selective and Sensitive Mercury(II) Detection. Adv. Mater. 2009, 21, 3674-3677. (27) Jose, D. A.; Konig, B. Polydiacetylene vesicles functionalized with N-heterocyclic ligands for metal cation binding. Org. Biomol. Chem. 2010, 8, 655-662. (28) Xia, H.; Li, J.; Zou, G.; Zhang, Q.; Jia, C. A highly sensitive and reusable cyanide anion sensor based on spiropyran functionalized polydiacetylene vesicular receptors. J. Mater. Chem. A 2013, 1, 10713-10719. (29) Chen, X.; Lee, J.; Jou, M. J.; Kim, J. M.; Yoon, J. Colorimetric and fluorometric detection of cationic surfactants based on conjugated polydiacetylene supramolecules. Chem. Commun. 2009, 3434-3436. (30) Chen, X.; Kang, S.; Kim, M. J.; Kim, J.; Kim, Y. S.; Kim, H.; Chi, B.; Kim, S. J.; Lee, J. Y.; Yoon, J. Thin-film formation of imidazolium-based conjugated polydiacetylenes and their application for sensing anionic surfactants. Angew. Chem. Int. Ed. 2010, 49, 1422-1425. (31) Thongmalai, W.; Eaidkong, T.; Ampornpun, S.; Mungkarndee, R.; Tumcharern, G.; Sukwattanasinitt, M.; Wacharasindhu, S. Polydiacetylenes carrying

induced reversibility, but not for most of the chemical and biological analytes. Most of the PDAs are based on commercial linear diacetylenes, while other types of diacetylenes are rarely studied. Solvatochromism and affinochromism were frequently utilized to construct PDA sensors, while other sensing principles such as electrochromism and mechanochromism, especially the former had been scarcely exploited. These fundamental problems need to be addressed to reach the full potential of PDA sensors. With regard to their applications, from the early reports in which PDA-based sensors were confined to laboratory settings, the development has evolved into fabricating PDA-based sensors that could eventually be utilized in industry and in our daily life, which we believe is still the core of future explorations of PDA-based materials. To achieve this goal, the collaborative efforts from scientists and engineers from various disciplines are required to not only obtain highly sensitive PDA sensors, but also to facilitate their incorporation into low-cost, user-friendly devices. A few promising examples had already been presented in this review, such as smartphone-based differentiation of organic solvents,117 wearable wrist band for the detection of DMF vapor,277 3D printed PDA/hydrogel for detoxification,248 and point-of-care device for the diagnosis of ovarian cancer.230 With these embryo prototypes in hand, the practical applications of PDA-based sensors will continue to expand and thus benefits consumers worldwide in the future.

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS publications website at DOI:

A table of PDA-based sensors reported in the literature from 2014 to 2018, including type of analyte, form of the sensor, and detection limit if available (PDF)

AUTHOR INFORMATION Corresponding Author *E-mail: [email protected] ORCID Brigitte Städler 0000-0002-7335-3945 Notes The authors declare no competing financial interest.

ACKNOWLEDGMENT We gratefully acknowledge the financial support by the Innovation Fund Denmark.

REFERENCE (1) Jung, H. S.; Chen, X.; Kim, J. S.; Yoon, J. Recent progress in luminescent and colorimetric chemosensors for detection of thiols. Chem. Soc. Rev. 2013, 42, 6019-6031. (2) Zhou, Y.; Xu, Z.; Yoon, J. Fluorescent and colorimetric chemosensors for detection of nucleotides, FAD and NADH: highlighted research during 2004-2010. Chem. Soc. Rev. 2011, 40, 2222-2235. (3) Kim, H. N.; Ren, W. X.; Kim, J. S.; Yoon, J. Fluorescent and colorimetric sensors for detection of lead, cadmium, and mercury ions. Chem. Soc. Rev. 2012, 41, 3210-3244. (4) Zhou, Y.; Yoon, J. Recent progress in fluorescent and colorimetric chemosensors for detection of amino acids. Chem. Soc. Rev. 2012, 41, 52-67. (5) Peng, Y.; Zhang, A. J.; Dong, M.; Wang, Y. W. A colorimetric and fluorescent chemosensor for the detection of an explosive--2,4,6-trinitrophenol (TNP). Chem. Commun. 2011, 47, 4505-4507.

18

ACS Paragon Plus Environment

Page 19 of 26 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials (55) Xu, Y.; Li, J.; Hu, W.; Zou, G.; Zhang, Q. Thermochromism and supramolecular chirality of the coumarin-substituted polydiacetylene LB films. J. Colloid Interface Sci. 2013, 400, 116-122. (56) Zhu, Y.; Xu, Y.; Zou, G.; Zhang, Q. Chirality Transfer and Modulation in LB Films Derived from the Diacetylene/Melamine Hydrogen-Bonded Complex. Chirality 2015, 27, 492-499. (57) Araghi, H. Y.; Paige, M. F. Insight into diacetylene photopolymerization in Langmuir-Blodgett films using simultaneous AFM and fluorescence microscopy imaging. Surf. Interface Anal. 2017, 49, 1108-1114. (58) Berman, A.; Ahn, D. J.; Lio, A.; Salmeron, M.; Reichert, A.; Charych, D. Total alignment of calcite at acidic polydiacetylene films: cooperativity at the organic-inorganic interface. Science 1995, 269, 515-518. (59) Ahn, D. J.; Chae, E. H.; Lee, G. S.; Shim, H. Y.; Chang, T. E.; Ahn, K. D.; Kim, J. M. Colorimetric reversibility of polydiacetylene supramolecules having enhanced hydrogen-bonding under thermal and pH stimuli. J. Am. Chem. Soc. 2003, 125, 8976-8977. (60) Alekseev, A.; Ihalainen, P.; Ivanov, A.; Domnin, I.; Rosqvist, E.; Lemmetyinen, H.; Vuorimaa-Laukkanen, E.; Peltonen, J.; Vyaz'min, S. Stable blue phase polymeric Langmuir-Schaefer films based on unsymmetrical hydroxyalkadiynyl N -arylcarbamate derivatives. Thin Solid Films 2018, 645, 108-118. (61) Chu, B.; Xu, R. Chromatic transition of polydiacetylene in solution. Acc. Chem. Res. 1991, 24, 384-389. (62) Okada, S.; Peng, S.; Spevak, W.; Charych, D. Color and Chromism of Polydiacetylene Vesicles. Acc. Chem. Res. 1998, 31, 229-239. (63) Kim, J. M.; Ji, E. K.; Woo, S. M.; Lee, H.; Ahn, D. J. Immobilized Polydiacetylene Vesicles on Solid Substrates for Use as Chemosensors. Adv. Mater. 2003, 15, 1118-1121. (64) Itoh, T.; Shichi, T.; Yui, T.; Takahashi, H.; Inui, Y.; Takagi, K. Reversible color changes in lamella hybrids of poly(diacetylenecarboxylates) incorporated in layered double hydroxide nanosheets. J. Phys. Chem. B 2005, 109, 3199-3206. (65) Peng, H.; Tang, J.; Yang, L.; Pang, J.; Ashbaugh, H. S.; Brinker, C. J.; Yang, Z.; Lu, Y. Responsive periodic mesoporous polydiacetylene/silica nanocomposites. J. Am. Chem. Soc. 2006, 128, 5304-5305. (66) Meir, D.; Silbert, L.; Volinsky, R.; Kolusheva, S.; Weiser, I.; Jelinek, R. Colorimetric/fluorescent bacterial sensing by agarose-embedded lipid/polydiacetylene films. J. Appl. Microbiol. 2008, 104, 787-795. (67) Kauffman, J. S.; Ellerbrock, B. M.; Stevens, K. A.; Brown, P. J.; Pennington, W. T.; Hanks, T. W. Preparation, characterization, and sensing behavior of polydiacetylene liposomes embedded in alginate fibers. ACS Appl. Mater. Interfaces 2009, 1, 1287-1291. (68) Clarisse, D.; Prakash, P.; Geertsen, V.; Miserque, F.; Gravel, E.; Doris, E. Aqueous 1,3-dipolar cycloadditions promoted by copper nanoparticles in polydiacetylene micelles. Green Chem. 2017, 19, 3112-3115. (69) Villemin, E.; Gravel, E.; Jawale, D. V.; Prakash, P.; Namboothiri, I. N. N.; Doris, E. Polydiacetylene Nanotubes in Heterogeneous Catalysis: Application to the Gold-Mediated Oxidation of Silanes. Macromol. Chem. Phys. 2015, 216, 2398-2403. (70) Jawale, D. V.; Gravel, E.; Villemin, E.; Shah, N.; Geertsen, V.; Namboothiri, I. N.; Doris, E. Co-catalytic oxidative coupling of primary amines to imines using an organic nanotube-gold nanohybrid. Chem. Commun. 2014, 50, 15251-15254. (71) Geng, Q.; Du, J. Reduction of 4-nitrophenol catalyzed by silver nanoparticles supported on polymer micelles and vesicles. RSC Adv. 2014, 4, 1642516428. (72) Li, L.; An, X.; Yan, X. Folate-polydiacetylene-liposome for tumor targeted drug delivery and fluorescent tracing. Colloid Surf. B-Biointerfaces 2015, 134, 235239. (73) Gravel, E.; Ogier, J.; Arnauld, T.; Mackiewicz, N.; Duconge, F.; Doris, E. Drug delivery and imaging with polydiacetylene micelles. Chem. Eur. J. 2012, 18, 400-408. (74) Mackiewicz, N.; Gravel, E.; Garofalakis, A.; Ogier, J.; John, J.; Dupont, D. M.; Gombert, K.; Tavitian, B.; Doris, E.; Duconge, F. Tumor-targeted polydiacetylene micelles for in vivo imaging and drug delivery. Small 2011, 7, 27862792. (75) Yao, D.; Li, S.; Zhu, X.; Wu, J.; Tian, H. Tumor-cell targeting polydiacetylene micelles encapsulated with an antitumor drug for the treatment of ovarian cancer. Chem. Commun. 2017, 53, 1233-1236. (76) Yan, X.; An, X. Multifunctional polydiacetylene-liposome with controlled release and fluorescence tracing. RSC Adv. 2014, 4, 18604-18607. (77) Fang, J. H.; Chiu, T. L.; Huang, W. C.; Lai, Y. H.; Hu, S. H.; Chen, Y. Y.; Chen, S. Y. Dual-Targeting Lactoferrin-Conjugated Polymerized Magnetic

amino groups for colorimetric detection and identification of anionic surfactants. J. Mater. Chem. 2011, 21, 16391-16397. (32) Lee, S.; Lee, K. M.; Lee, M.; Yoon, J. Polydiacetylenes bearing boronic acid groups as colorimetric and fluorescence sensors for cationic surfactants. ACS Appl. Mater. Interfaces 2013, 5, 4521-4526. (33) Ma, Z.; Li, J.; Liu, M.; Cao, J.; Zou, Z.; Tu, J.; Jiang, L. Colorimetric Detection of Escherichia coli by Polydiacetylene Vesicles Functionalized with Glycolipid. J. Am. Chem. Soc. 1998, 120, 12678-12679. (34) Zhang, Y.; Ma, B.; Li, Y.; Li, J. Enhanced affinochromism of polydiacetylene monolayer in response to bacteria by incorporating CdS nano-crystallites. Colloid Surf. B-Biointerfaces 2004, 35, 41-44. (35) Pindzola, B. A.; Nguyen, A. T.; Reppy, M. A. Antibody-functionalized polydiacetylene coatings on nanoporous membranes for microorganism detection. Chem. Commun. 2006, 906-908. (36) Scindia, Y.; Silbert, L.; Volinsky, R.; Kolusheva, S.; Jelinek, R. Colorimetric detection and fingerprinting of bacteria by glass-supported lipid/polydiacetylene films. Langmuir 2007, 23, 4682-4687. (37) Park, C. K.; Kang, C. D.; Sim, S. J. Non-labeled detection of waterborne pathogen Cryptosporidium parvum using a polydiacetylene-based fluorescence chip. Biotechnol. J. 2008, 3, 687-693. (38) Jung, Y. K.; Kim, T. W.; Jung, C.; Cho, D. Y.; Park, H. G. A polydiacetylene microchip based on a biotin-streptavidin interaction for the diagnosis of pathogen infections. Small 2008, 4, 1778-1784. (39) Park, C. H.; Kim, J. P.; Lee, S. W.; Jeon, N. L.; Yoo, P. J.; Sim, S. J. A Direct, Multiplex Biosensor Platform for Pathogen Detection Based on Cross-linked Polydiacetylene (PDA) Supramolecules. Adv. Funct. Mater. 2009, 19, 37033710. (40) Cheng, Q.; Stevens, R. C. Coupling of an induced fit enzyme to polydiacetylene thin films: Colorimetric detection of glucose. Adv. Mater. 1997, 9, 481-483. (41) Geiger, E.; Hug, P.; Keller, B. A. Chromatic Transitions in Polydiacetylene Langmuir-Blodgett Films due to Molecular Recognition at the Film Surface Studied by Spectroscopic Methods and Surface Analysis. Macromol. Chem. Phys. 2002, 203, 2422-2431. (42) Ma, G.; Cheng, Q. Vesicular polydiacetylene sensor for colorimetric signaling of bacterial pore-forming toxin. Langmuir 2005, 21, 6123-6126. (43) Jung, Y. K.; Kim, T. W.; Kim, J.; Kim, J.-M.; Park, H. G. Universal Colorimetric Detection of Nucleic Acids Based on Polydiacetylene (PDA) Liposomes. Adv. Funct. Mater. 2008, 18, 701-708. (44) Jung, Y. K.; Kim, T. W.; Park, H. G.; Soh, H. T. Specific Colorimetric Detection of Proteins Using Bidentate Aptamer-Conjugated Polydiacetylene (PDA) Liposomes. Adv. Funct. Mater. 2010, 20, 3092-3097. (45) Wu, J.; Zawistowski, A.; Ehrmann, M.; Yi, T.; Schmuck, C. Peptide functionalized polydiacetylene liposomes act as a fluorescent turn-on sensor for bacterial lipopolysaccharide. J. Am. Chem. Soc. 2011, 133, 9720-9723. (46) Seo, D.; Kim, J. Effect of the Molecular Size of Analytes on Polydiacetylene Chromism. Adv. Funct. Mater. 2010, 20, 1397-1403. (47) Leal, M. P.; Assali, M.; Fernandez, I.; Khiar, N. Copper-catalyzed azide-alkyne cycloaddition in the synthesis of polydiacetylene: "click glycoliposome" as biosensors for the specific detection of lectins. Chem. Eur. J. 2011, 17, 1828-1836. (48) Kang, D. H.; Jung, H. S.; Ahn, N.; Lee, J.; Seo, S.; Suh, K. Y.; Kim, J.; Kim, K. Biomimetic detection of aminoglycosidic antibiotics using polydiacetylenephospholipids supramolecules. Chem. Commun. 2012, 48, 5313-5315. (49) Kwon, I. K.; Song, M. S.; Won, S. H.; Choi, S. P.; Kim, M.; Sim, S. J. Signal amplification by magnetic force on polydiacetylene supramolecules for detection of prostate cancer. Small 2012, 8, 209-213. (50) Zhou, G.; Wang, F.; Wang, H.; Kambam, S.; Chen, X. Colorimetric and fluorometric detection of neomycin based on conjugated polydiacetylene supramolecules. Macromol. Rapid Commun. 2013, 34, 944-948. (51) Zhou, G.; Wang, F.; Wang, H.; Kambam, S.; Chen, X.; Yoon, J. Colorimetric and fluorometric assays based on conjugated polydiacetylene supramolecules for screening acetylcholinesterase and its inhibitors. ACS Appl. Mater. Interfaces 2013, 5, 3275-3280. (52) Cho, Y.-S.; Ahn, K. H. Molecular interactions between charged macromolecules: colorimetric detection and quantification of heparin with a polydiacetylene liposome. J. Mater. Chem. B 2013, 1, 1182-1189. (53) Jeon, H.; Lee, S.; Li, Y.; Park, S.; Yoon, J. Conjugated polydiacetylenes bearing quaternary ammonium groups as a dual colorimetric and fluorescent sensor for ATP. J. Mater. Chem. 2012, 22, 3795-3799. (54) Tachibana, H.; Yamanaka, Y.; Sakai, H.; Abe, M.; Matsumoto, M. Effect of Position of Butadiyne Moiety in Amphiphilic Diacetylenes on the Polymerization in the Langmuir−Blodgett Films. Macromolecules 1999, 32, 8306-8309.

19

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Polydiacetylene-Assembled Nanocarriers with Self-Responsive Fluorescence/Magnetic Resonance Imaging for In Vivo Brain Tumor Therapy. Adv. Healthc. Mater. 2016, 5, 688-695. (78) Ripoll, M.; Neuberg, P.; Kichler, A.; Tounsi, N.; Wagner, A.; Remy, J. S. pHResponsive Nanometric Polydiacetylenic Micelles Allow for Efficient Intracellular siRNA Delivery. ACS Appl. Mater. Interfaces 2016, 8, 30665-30670. (79) Neuberg, P.; Perino, A.; Morin-Picardat, E.; Anton, N.; Darwich, Z.; Weltin, D.; Mely, Y.; Klymchenko, A. S.; Remy, J. S.; Wagner, A. Photopolymerized micelles of diacetylene amphiphile: physical characterization and cell delivery properties. Chem. Commun. 2015, 51, 11595-11598. (80) Morin, E.; Nothisen, M.; Wagner, A.; Remy, J. S. Cationic polydiacetylene micelles for gene delivery. Bioconjugate Chem. 2011, 22, 1916-1923. (81) Ma, B.; Zhang, S.; Jiang, H.; Zhao, B.; Lv, H. Lipoplex morphologies and their influences on transfection efficiency in gene delivery. J. Control. Release 2007, 123, 184-194. (82) Jiang, H.; Hu, X. Y.; Schlesiger, S.; Li, M.; Zellermann, E.; Knauer, S. K.; Schmuck, C. Morphology-Dependent Cell Imaging by Using a Self-Assembled Diacetylene Peptide Amphiphile. Angew. Chem. Int. Ed. 2017, 56, 14526-14530. (83) Jung, Y. K.; Woo, M. A.; Soh, H. T.; Park, H. G. Aptamer-based cell imaging reagents capable of fluorescence switching. Chem. Commun. 2014, 50, 1232912332. (84) Haridas, V.; Sadanandan, S.; Collart-Dutilleul, P. Y.; Gronthos, S.; Voelcker, N. H. Lysine-appended polydiacetylene scaffolds for human mesenchymal stem cells. Biomacromolecules 2014, 15, 582-590. (85) Theodorou, I.; Anilkumar, P.; Lelandais, B.; Clarisse, D.; Doerflinger, A.; Gravel, E.; Duconge, F.; Doris, E. Stable and compact zwitterionic polydiacetylene micelles with tumor-targeting properties. Chem. Commun. 2015, 51, 1493714940. (86) Desta, M. A.; Liao, C. W.; Sun, S. S. A General Strategy to Enhance the Performance of Dye-Sensitized Solar Cells by Incorporating a Light-Harvesting Dye with a Hydrophobic Polydiacetylene Electrolyte-Blocking Layer. Chem.-Asian J. 2017, 12, 690-697. (87) Pootrakulchote, N.; Reanprayoon, C.; Gasiorowski, J.; Sariciftci, N. S.; Thamyongkit, P. A polydiacetylene-nested porphyrin conjugate for dye-sensitized solar cells. New J. Chem. 2015, 39, 9228-9233. (88) Reanprayoon, C.; Gasiorowski, J.; Sukwattanasinitt, M.; Sariciftci, N. S.; Thamyongkit, P. Polydiacetylene-nested porphyrin as a potential light harvesting component in bulk heterojunction solar cells. RSC Adv. 2014, 4, 3045-3050. (89) Bhowmik, S.; Konda, M.; Das, A. K. Light induced construction of porous covalent organic polymeric networks for significant enhancement of CO2 gas sorption. RSC Adv. 2017, 7, 47695-47703. (90) Bhowmik, S.; Jadhav, R. G.; Das, A. K. Nanoporous Conducting Covalent Organic Polymer (COP) Nanostructures as Metal-Free High Performance Visible-Light Photocatalyst for Water Treatment and Enhanced CO2 Capture. J. Phys. Chem. C 2018, 122, 274-284. (91) Nishide, J.; Oyamada, T.; Akiyama, S.; Sasabe, H.; Adachi, C. High FieldEffect Mobility in an Organic Thin-Film Transistor with a Solid-State Polymerized Polydiacetylene Film as an Active Layer. Adv. Mater. 2006, 18, 3120-3124. (92) Cho, S.; Han, G.; Kim, K.; Sung, M. M. High-performance two-dimensional polydiacetylene with a hybrid inorganic-organic structure. Angew. Chem. Int. Ed. 2011, 50, 2742-2746. (93) Liang, J.; Huang, L.; Li, N.; Huang, Y.; Wu, Y.; Fang, S.; Oh, J.; Kozlov, M.; Ma, Y.; Li, F.; Baughman, R.; Chen, Y. Electromechanical actuator with controllable motion, fast response rate, and high-frequency resonance based on graphene and polydiacetylene. ACS Nano 2012, 6, 4508-4519. (94) Ulaganathan, M.; Hansen, R. V.; Drayton, N.; Hingorani, H.; Kutty, R. G.; Joshi, H.; Sreejith, S.; Liu, Z.; Yang, J.; Zhao, Y. Photopolymerization of Diacetylene on Aligned Multiwall Carbon Nanotube Microfibers for High-Performance Energy Devices. ACS Appl. Mater. Interfaces 2016, 8, 32643-32648. (95) Choi, Y. K.; Kim, H. J.; Kim, S. R.; Cho, Y. M.; Ahn, D. J. Enhanced Thermal Stability of Polyaniline with Polymerizable Dopants. Macromolecules 2017, 50, 3164-3170. (96) Lebegue, E.; Farre, C.; Jose, C.; Saulnier, J.; Lagarde, F.; Chevalier, Y.; Chaix, C.; Jaffrezic-Renault, N. Responsive Polydiacetylene Vesicles for Biosensing Microorganisms. Sensors 2018, 18, 599. (97) Lee, S.; Kim, J. Y.; Chen, X.; Yoon, J. Recent progress in stimuli-induced polydiacetylenes for sensing temperature, chemical and biological targets. Chem. Commun. 2016, 52, 9178-9196. (98) Diegelmann, S. R.; Tovar, J. D. Polydiacetylene-Peptide 1D Nanomaterials. Macromol. Rapid Commun. 2013, 34, 1343-1350.

Page 20 of 26

(99) Chen, X.; Zhou, G.; Peng, X.; Yoon, J. Biosensors and chemosensors based on the optical responses of polydiacetylenes. Chem. Soc. Rev. 2012, 41, 46104630. (100) Yarimaga, O.; Jaworski, J.; Yoon, B.; Kim, J. M. Polydiacetylenes: supramolecular smart materials with a structural hierarchy for sensing, imaging and display applications. Chem. Commun. 2012, 48, 2469-2485. (101) Alami, M.; Ferri, F. A convenient route to unsymmetrical conjugated diynes. Tetrahedron Lett. 1996, 37, 2763-2766. (102) Mata, A.; Hsu, L.; Capito, R.; Aparicio, C.; Henrikson, K.; Stupp, S. I. Micropatterning of bioactive self-assembling gels. Soft Matter 2009, 5, 1228-1236. (103) Nieuwland, M.; van Gijzel, N.; van Hest, J. C.; Lowik, D. W. The influence of amino acid sequence on structure and morphology of polydiacetylene containing peptide fibres. Soft Matter 2015, 11, 1335-1344. (104) Yang, D.; Zou, R.; Zhu, Y.; Liu, B.; Yao, D.; Jiang, J.; Wu, J.; Tian, H. Magainin II modified polydiacetylene micelles for cancer therapy. Nanoscale 2014, 6, 14772-14783. (105) Diegelmann, S. R.; Hartman, N.; Markovic, N.; Tovar, J. D. Synthesis and alignment of discrete polydiacetylene-peptide nanostructures. J. Am. Chem. Soc. 2012, 134, 2028-2031. (106) Menzel, H.; Horstmann, S.; Mowery, M. D.; Cai, M.; Evans, C. E. Diacetylene polymerization in self-assembled monolayers: influence of the odd/even nature of the methylene spacer. Polymer 2000, 41, 8113-8119. (107) Aoki, K.; Kudo, M.; Tamaoki, N. Novel odd/even effect of alkylene chain length on the photopolymerizability of organogelators. Org. Lett. 2004, 6, 40094012. (108) Charoenthai, N.; Pattanatornchai, T.; Wacharasindhu, S.; Sukwattanasinitt, M.; Traiphol, R. Roles of head group architecture and side chain length on colorimetric response of polydiacetylene vesicles to temperature, ethanol and pH. J. Colloid Interface Sci. 2011, 360, 565-573. (109) Khanantong, C.; Charoenthai, N.; Phuangkaew, T.; Kielar, F.; Traiphol, N.; Traiphol, R. Phase transition, structure and color-transition behaviors of monocarboxylic diacetylene and polydiacetylene assemblies: The opposite effects of alkyl chain length. Colloid Surf. A-Physicochem. Eng. Asp. 2018, 553, 337348. (110) Park, D.-H.; Kim, B.; Kim, J.-M. A Tetrahydrofuran-selective Optical Solvent Sensor Based on Solvatochromic Polydiacetylene. Bull. Korean Chem. Soc. 2016, 37, 793-794. (111) Wang, X.; Sun, X.; Hu, P. A.; Zhang, J.; Wang, L.; Feng, W.; Lei, S.; Yang, B.; Cao, W. Colorimetric Sensor Based on Self-Assembled Polydiacetylene/Graphene-Stacked Composite Film for Vapor-Phase Volatile Organic Compounds. Adv. Funct. Mater. 2013, 23, 6044-6050. (112) Lee, J.; Chang, H. T.; An, H.; Ahn, S.; Shim, J.; Kim, J. M. A protective layer approach to solvatochromic sensors. Nat. Commun. 2013, 4, 2461. (113) Yoon, J.; Chae, S. K.; Kim, J. M. Colorimetric sensors for volatile organic compounds (VOCs) based on conjugated polymer-embedded electrospun fibers. J. Am. Chem. Soc. 2007, 129, 3038-3039. (114) Jiang, H.; Wang, Y.; Ye, Q.; Zou, G.; Su, W.; Zhang, Q. Polydiacetylenebased colorimetric sensor microarray for volatile organic compounds. Sens. Actuator B-Chem. 2010, 143, 789-794. (115) Eaidkong, T.; Mungkarndee, R.; Phollookin, C.; Tumcharern, G.; Sukwattanasinitt, M.; Wacharasindhu, S. Polydiacetylene paper-based colorimetric sensor array for vapor phase detection and identification of volatile organic compounds. J. Mater. Chem. 2012, 22, 5970-5977. (116) Yoon, J.; Jung, Y.-S.; Kim, J.-M. A Combinatorial Approach for Colorimetric Differentiation of Organic Solvents Based on Conjugated Polymer-Embedded Electrospun Fibers. Adv. Funct. Mater. 2009, 19, 209-214. (117) Park, D. H.; Heo, J. M.; Jeong, W.; Yoo, Y. H.; Park, B. J.; Kim, J. M. Smartphone-Based VOC Sensor Using Colorimetric Polydiacetylenes. ACS Appl. Mater. Interfaces 2018, 10, 5014-5021. (118) Dolai, S.; Bhunia, S. K.; Beglaryan, S. S.; Kolusheva, S.; Zeiri, L.; Jelinek, R. Colorimetric Polydiacetylene-Aerogel Detector for Volatile Organic Compounds (VOCs). ACS Appl. Mater. Interfaces 2017, 9, 2891-2898. (119) Tu, M.-C.; Cheema, J. A.; Yildiz, U. H.; Palaniappan, A.; Liedberg, B. Vapor phase solvatochromic responses of polydiacetylene embedded matrix polymers. J. Mater. Chem. C 2017, 5, 1803-1809. (120) Exarhos, G. J.; Risen, W. M.; Baughman, R. H. Resonance Raman study of the thermochromic phase transition of a polydiacetylene. J. Am. Chem. Soc. 1976, 98, 481-487. (121) Yu, X.; Luo, Y.; Wu, W.; Yan, Q.; Zou, G.; Zhang, Q. Synthesis and reversible thermochromism of azobenzene-containing polydiacetylenes. Eur. Polym. J. 2008, 44, 3015-3021.

20

ACS Paragon Plus Environment

Page 21 of 26 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials nanocomposites: Effects of PDA side chain length and PVP molecular weight. J. Ind. Eng. Chem. 2017, 46, 130-138. (144) Gu, Y.; Cao, W.; Zhu, L.; Chen, D.; Jiang, M. Polymer Mortar Assisted SelfAssembly of Nanocrystalline Polydiacetylene Bricks Showing Reversible Thermochromism. Macromolecules 2008, 41, 2299-2303. (145) Oaki, Y.; Ishijima, Y.; Imai, H. Emergence of temperature-dependent and reversible color-changing properties by the stabilization of layered polydiacetylene through intercalation. Polym. J. 2018, 50, 319-326. (146) Shimogaki, T.; Matsumoto, A. Structural and Chromatic Changes of Host Polydiacetylene Crystals during Intercalation with Guest Alkylamines. Macromolecules 2011, 44, 3323-3327. (147) Traiphol, N.; Rungruangviriya, N.; Potai, R.; Traiphol, R. Stable polydiacetylene/ZnO nanocomposites with two-steps reversible and irreversible thermochromism: the influence of strong surface anchoring. J. Colloid Interface Sci. 2011, 356, 481-489. (148) Chanakul, A.; Traiphol, N.; Traiphol, R. Controlling the reversible thermochromism of polydiacetylene/zinc oxide nanocomposites by varying alkyl chain length. J. Colloid Interface Sci. 2013, 389, 106-114. (149) Traiphol, N.; Chanakul, A.; Kamphan, A.; Traiphol, R. Role of Zn2+ ion on the formation of reversible thermochromic polydiacetylene/zinc oxide nanocomposites. Thin Solid Films 2017, 622, 122-129. (150) Takeuchi, M.; Imai, H.; Oaki, Y. Real-Time Imaging of 2D and 3D Temperature Distribution: Coating of Metal-Ion-Intercalated Organic Layered Composites with Tunable Stimuli-Responsive Properties. ACS Appl. Mater. Interfaces 2017, 9, 16546-16552. (151) Okaniwa, M.; Oaki, Y.; Kaneko, S.; Ishida, K.; Maki, H.; Imai, H. Advanced Biomimetic Approach for Crystal Growth in Nonaqueous Media: Morphology and Orientation Control of Pentacosadiynoic Acid and Applications. Chem. Mater. 2015, 27, 2627-2632. (152) Wu, A.; Beck, C.; Ying, Y.; Federici, J.; Iqbal, Z. Thermochromism in Polydiacetylene–ZnO Nanocomposites. J. Phys. Chem. C 2013, 19593-19600. (153) Patlolla, A.; Zunino, J.; Frenkel, A. I.; Iqbal, Z. Thermochromism in polydiacetylene-metal oxide nanocomposites. J. Mater. Chem. 2012, 22, 7028-7035. (154) Yao, Y.; Fu, K.; Huang, X.; Chen, D. Polydiacetylene-Tb3+ Nanosheets of Which Both the Color and the Fluorescence Can Be Reversibly Switched between Two Colors. Chin. J. Chem . 2017, 35, 1678-1686. (155) Guo, J.; Fu, K.; Zhang, Z.; Yang, L.; Huang, Y.-C.; Huang, C.-I.; Zhu, L.; Chen, D. Reversible thermochromism via hydrogen-bonded cocrystals of polydiacetylene and melamine. Polymer 2016, 105, 440-448. (156) Peng, J.; Cheng, Y.; Tomsia, A. P.; Jiang, L.; Cheng, Q. Thermochromic Artificial Nacre Based on Montmorillonite. ACS Appl. Mater. Interfaces 2017, 9, 24993-24998. (157) Lu, X.; Zhang, Z.; Sun, X.; Chen, P.; Zhang, J.; Guo, H.; Shao, Z.; Peng, H. Flexible and stretchable chromatic fibers with high sensing reversibility. Chem. Sci. 2016, 7, 5113-5117. (158) Tomioka, Y.; Tanaka, N.; Imazeki, S. Surface‐pressure‐induced reversible color change of a polydiacetylene monolayer at a gas–water interface. J. Chem. Phys. 1989, 91, 5694-5700. (159) Carpick, R. W.; Sasaki, D. Y.; Burns, A. R. First Observation of Mechanochromism at the Nanometer Scale. Langmuir 2000, 16, 1270-1278. (160) Feng, H.; Lu, J.; Li, J.; Tsow, F.; Forzani, E.; Tao, N. Hybrid mechanoresponsive polymer wires under force activation. Adv. Mater. 2013, 25, 17291733. (161) Park, D.-H.; Hong, J.; Park, I. S.; Lee, C. W.; Kim, J.-M. A Colorimetric Hydrocarbon Sensor Employing a Swelling-Induced Mechanochromic Polydiacetylene. Adv. Funct. Mater. 2014, 24, 5186-5193. (162) Hong, J.; Park, D. H.; Baek, S.; Song, S.; Lee, C. W.; Kim, J. M. Polydiacetylene-embedded microbeads for colorimetric and volumetric sensing of hydrocarbons. ACS Appl. Mater. Interfaces 2015, 7, 8339-8343. (163) Seo, S.; Lee, J.; Kwon, M. S.; Seo, D.; Kim, J. Stimuli-Responsive MatrixAssisted Colorimetric Water Indicator of Polydiacetylene Nanofibers. ACS Appl. Mater. Interfaces 2015, 7, 20342-20348. (164) Ishijima, Y.; Imai, H.; Oaki, Y. Tunable Mechano-responsive ColorChange Properties of Organic Layered Material by Intercalation. Chem 2017, 3, 509-521. (165) Terada, H.; Imai, H.; Oaki, Y. Visualization and Quantitative Detection of Friction Force by Self-Organized Organic Layered Composites. Adv. Mater. 2018, 30, 1801121. (166) You, X.; Chen, X.; Zou, G.; Su, W.; Zhang, Q.; He, P. Colorimetric response of azobenzene-terminated polydiacetylene vesicles under thermal and photic stimuli. Chem. Phys. Lett. 2009, 482, 129-133.

(122) Phollookin, C.; Wacharasindhu, S.; Ajavakom, A.; Tumcharern, G.; Ampornpun, S.; Eaidkong, T.; Sukwattanasinitt, M. Tuning Down of Color Transition Temperature of Thermochromically Reversible Bisdiynamide Polydiacetylenes. Macromolecules 2010, 43, 7540-7548. (123) Yu, L.; Hsu, S. L. A Spectroscopic Analysis of the Role of Side Chains in Controlling Thermochromic Transitions in Polydiacetylenes. Macromolecules 2011, 45, 420-429. (124) Tanioku, C.; Matsukawa, K.; Matsumoto, A. Thermochromism and structural change in polydiacetylenes including carboxy and 4-carboxyphenyl groups as the intermolecular hydrogen bond linkages in the side chain. ACS Appl. Mater. Interfaces 2013, 5, 940-948. (125) Shin, H.; Yoon, B.; Park, I. S.; Kim, J. M. An electrothermochromic paper display based on colorimetrically reversible polydiacetylenes. Nanotechnology 2014, 25, 094011. (126) Park, I. S.; Park, H. J.; Jeong, W.; Nam, J.; Kang, Y.; Shin, K.; Chung, H.; Kim, J.-M. Low Temperature Thermochromic Polydiacetylenes: Design, Colorimetric Properties, and Nanofiber Formation. Macromolecules 2016, 49, 12701278. (127) Han, N.; Woo, H. J.; Kim, S. E.; Jung, S.; Shin, M. J.; Kim, M.; Shin, J. S. Systemized organic functional group controls in polydiacetylenes and their effects on color changes. J. Appl. Polym. Sci. 2017, 134, 45011. (128) Park, H.; Lee, J. S.; Choi, H.; Ahn, D. J.; Kim, J. M. Rational Design of Supramolecular Conjugated Polymers Displaying Unusual Colorimetric Stability upon Thermal Stress. Adv. Funct. Mater. 2007, 17, 3447-3455. (129) Park, I. S.; Park, H. J.; Kim, J. M. A soluble, low-temperature thermochromic and chemically reactive polydiacetylene. ACS Appl. Mater. Interfaces 2013, 5, 8805-8812. (130) Lu, J.; Zhou, J.; Li, J. Tuned chromic process for polydiacetylenes vesicles: the influence of polymer matrices. Soft Matter 2011, 7, 6529-6531. (131) Lee, S.; Lee, J.; Kim, H. N.; Kim, M. H.; Yoon, J. Thermally reversible polydiacetylenes derived from ethylene oxide-containing bisdiacetylenes. Sens. Actuator B-Chem. 2012, 173, 419-425. (132) Chen, X.; Yoon, J. A thermally reversible temperature sensor based on polydiacetylene: Synthesis and thermochromic properties. Dyes Pigment. 2011, 89, 194-198. (133) Kim, J.-M.; Lee, J.-S.; Choi, H.; Sohn, D.; Ahn, D. J. Rational Design and in-Situ FTIR Analyses of Colorimetrically Reversibe Polydiacetylene Supramolecules. Macromolecules 2005, 38, 9366-9376. (134) Wacharasindhu, S.; Montha, S.; Boonyiseng, J.; Potisatityuenyong, A.; Phollookin, C.; Tumcharern, G.; Sukwattanasinitt, M. Tuning of Thermochromic Properties of Polydiacetylene toward Universal Temperature Sensing Materials through Amido Hydrogen Bonding. Macromolecules 2010, 43, 716724. (135) Hu, W.; Hao, J.; Li, J.; Zou, G.; Zhang, Q. Novel Chromatic Transitions of Azobenzene-Functionalized Polydiacetylene Aggregates in 1,2-Dichlorobenzene Solution. Macromol. Chem. Phys. 2012, 213, 2582-2589. (136) Park, S. H.; Roh, J.; Ahn, D. J. Optimal photoluminescence achieved by control of photopolymerization for diacetylene derivatives that induce reversible, partially reversible, and irreversible responses. Macromol. Res. 2017, 25, 960-962. (137) Huo, J.; Hu, Z.; He, G.; Hong, X.; Yang, Z.; Luo, S.; Ye, X.; Li, Y.; Zhang, Y.; Zhang, M.; Chen, H.; Fan, T.; Zhang, Y.; Xiong, B.; Wang, Z.; Zhu, Z.; Chen, D. High temperature thermochromic polydiacetylenes: Design and colorimetric properties. Appl. Surf. Sci. 2017, 423, 951-956. (138) Dong, W.; Lin, G.; Wang, H.; Lu, W. New Dendritic Polydiacetylene Sensor with Good Reversible Thermochromic Ability in Aqueous Solution and Solid Film. ACS Appl. Mater. Interfaces 2017, 9, 11918-11923. (139) Zhang, L.; Yuan, Y.-Z.; Tian, X.-H.; Sun, J.-Y. A thermally reversible supramolecular system based on biphenyl polydiacetylene. Chin. Chem. Lett. 2015, 26, 1133-1136. (140) Niu, R.; Meng, X.-l.; Yang, D.-d.; Chang, Y.; Zha, F. Preparation of Reversible Thermochromism Supramolecules of 4-Aminophenol-Modified Polydiacetylene. Arab. J. Sci. Eng. 2015, 40, 2867-2872. (141) Guo, H.; Zhang, J.; Porter, D.; Peng, H.; Löwik, D. W. P. M.; Wang, Y.; Zhang, Z.; Chen, X.; Shao, Z. Ultrafast and reversible thermochromism of a conjugated polymer material based on the assembly of peptide amphiphiles. Chem. Sci. 2014, 5, 4189-4195. (142) Kamphan, A.; Traiphol, N.; Traiphol, R. Versatile route to prepare reversible thermochromic polydiacetylene nanocomposite using low molecular weight poly(vinylpyrrolidone). Colloid Surf. A-Physicochem. Eng. Asp. 2016, 497, 370377. (143) Kamphan, A.; Khanantong, C.; Traiphol, N.; Traiphol, R. Structural-thermochromic relationship of polydiacetylene (PDA)/polyvinylpyrrolidone (PVP)

21

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 26

(191) Thet, N. T.; Jamieson, W. D.; Laabei, M.; Mercer-Chalmers, J. D.; Jenkins, A. T. Photopolymerization of polydiacetylene in hybrid liposomes: effect of polymerization on stability and response to pathogenic bacterial toxins. J. Phys. Chem. B 2014, 118, 5418-5427. (192) Thet, N. T.; Alves, D. R.; Bean, J. E.; Booth, S.; Nzakizwanayo, J.; Young, A. E.; Jones, B. V.; Jenkins, A. T. Prototype Development of the Intelligent Hydrogel Wound Dressing and Its Efficacy in the Detection of Model Pathogenic Wound Biofilms. ACS Appl. Mater. Interfaces 2016, 8, 14909-14919. (193) Zhou, J.; Yao, D.; Qian, Z.; Hou, S.; Li, L.; Jenkins, A. T. A.; Fan, Y. Bacteria-responsive intelligent wound dressing: Simultaneous In situ detection and inhibition of bacterial infection for accelerated wound healing. Biomaterials 2018, 161, 11-23. (194) Kolusheva, S.; Shahal, T.; Jelinek, R. Peptide−Membrane Interactions Studied by a New Phospholipid/Polydiacetylene Colorimetric Vesicle Assay†. Biochemistry 2000, 39, 15851-15859. (195) Shin, M. J.; Kim, Y. J.; Kim, J. D. Chromatic response of polydiacetylene vesicle induced by the permeation of methotrexate. Soft Matter 2015, 11, 50375043. (196) Shin, Y. J.; Shin, M. J.; Shin, J. S. Permeation-induced chromatic change of a polydiacetylene vesicle with nonionic surfactant. Colloid Surf. A-Physicochem. Eng. Asp. 2017, 520, 459-466. (197) Yadav, M. K.; Kumar, V.; Singh, B.; Tiwari, S. K. Phospholipid/Polydiacetylene Vesicle-Based Colorimetric Assay for High-Throughput Screening of Bacteriocins and Halocins. Appl. Biochem. Biotechnol. 2017, 182, 142-154. (198) Kang, D. H.; Jung, H. S.; Lee, J.; Seo, S.; Kim, J.; Kim, K.; Suh, K. Y. Design of polydiacetylene-phospholipid supramolecules for enhanced stability and sensitivity. Langmuir 2012, 28, 7551-7556. (199) Needleman, H. Lead poisoning. Annu. Rev. Med. 2004, 55, 209-222. (200) Lee, K. M.; Chen, X.; Fang, W.; Kim, J. M.; Yoon, J. A dual colorimetric and fluorometric sensor for lead ion based on conjugated polydiacetylenes. Macromol. Rapid Commun. 2011, 32, 497-500. (201) Wang, D.-E.; Wang, Y.; Tian, C.; Zhang, L.; Han, X.; Tu, Q.; Yuan, M.; Chen, S.; Wang, J. Polydiacetylene liposome-encapsulated alginate hydrogel beads for Pb2+ detection with enhanced sensitivity. J. Mater. Chem. A 2015, 3, 21690-21698. (202) Wang, M.; Wang, F.; Wang, Y.; Zhang, W.; Chen, X. Polydiacetylenebased sensor for highly sensitive and selective Pb2+ detection. Dyes Pigment. 2015, 120, 307-313. (203) Pan, X.; Wang, Y.; Jiang, H.; Zou, G.; Zhang, Q. Benzo-15-crown-5 functionalized polydiacetylene-based colorimetric self-assembled vesicular receptors for lead ion recognition. J. Mater. Chem. 2011, 21, 3604-3610. (204) Li, Y.; Wang, L.; Yin, X.; Ding, B.; Sun, G.; Ke, T.; Chen, J.; Yu, J. Colorimetric strips for visual lead ion recognition utilizing polydiacetylene embedded nanofibers. J. Mater. Chem. A 2014, 2, 18304-18312. (205) Narkwiboonwong, P.; Tumcharern, G.; Potisatityuenyong, A.; Wacharasindhu, S.; Sukwattanasinitt, M. Aqueous sols of oligo(ethylene glycol) surface decorated polydiacetylene vesicles for colorimetric detection of Pb2+. Talanta 2011, 83, 872-878. (206) Li, Y.; Wang, L.; Wen, Y.; Ding, B.; Sun, G.; Ke, T.; Chen, J.; Yu, J. Constitution of a visual detection system for lead(ii) on polydiacetylene–glycine embedded nanofibrous membranes. J. Mater. Chem. A 2015, 3, 9722-9730. (207) Lee, C. G.; Kang, S.; Oh, J.; Eom, M. S.; Oh, J.; Kim, M.-G.; Lee, W. S.; Hong, S.; Han, M. S. A colorimetric and fluorescent chemosensor for detection of Hg2+ using counterion exchange of cationic polydiacetylene. Tetrahedron Lett. 2017, 58, 4340-4343. (208) Gwon, Y. J.; Kim, C.; Lee, T. S. Chromatic detection of Cs ions using polydiacetylene-based vesicles containing crown-ether-like ethylene glycol units. Sens. Actuator B-Chem. 2019, 281, 343-349. (209) Xu, Y.; Yang, G.; Xia, H.; Zou, G.; Zhang, Q.; Gao, J. Enantioselective synthesis of helical polydiacetylene by application of linearly polarized light and magnetic field. Nat. Commun. 2014, 5, 5050. (210) Li, S.; Zhang, L.; Jiang, J.; Meng, Y.; Liu, M. Self-Assembled Polydiacetylene Vesicle and Helix with Chiral Interface for Visualized Enantioselective Recognition of Sulfinamide. ACS Appl. Mater. Interfaces 2017, 9, 37386-37394. (211) Silbert, L.; Ben Shlush, I.; Israel, E.; Porgador, A.; Kolusheva, S.; Jelinek, R. Rapid chromatic detection of bacteria by use of a new biomimetic polymer sensor. Appl. Environ. Microbiol. 2006, 72, 7339-7344. (212) Zhang, Y.; Fan, Y.; Sun, C.; Shen, D.; Li, Y.; Li, J. Functionalized polydiacetylene-glycolipid vesicles interacted with Escherichia coli under the TiO2 colloid. Colloid Surf. B-Biointerfaces 2005, 40, 137-142.

(167) Li, J.; Jiang, H.; Hu, W.; Xia, H.; Zou, G.; Zhang, Q. Photo-controlled hierarchical assembly and fusion of coumarin-containing polydiacetylene vesicles. Macromol. Rapid Commun. 2013, 34, 274-279. (168) Sun, X.; Chen, T.; Huang, S.; Cai, F.; Chen, X.; Yang, Z.; Li, L.; Cao, H.; Lu, Y.; Peng, H. UV-induced chromatism of polydiacetylenic assemblies. J. Phys. Chem. B 2010, 114, 2379-2382. (169) Yan, X.; An, X. Thermal and photic stimuli-responsive polydiacetylene liposomes with reversible fluorescence. Nanoscale 2013, 5, 6280-6283. (170) Lee, J.; Seo, S.; Kim, J. Rapid Light-Driven Color Transition of Novel Photoresponsive Polydiacetylene Molecules. ACS Appl. Mater. Interfaces 2018, 10, 3164-3169. (171) Baek, W.; Heo, J. M.; Oh, S.; Lee, S. H.; Kim, J.; Joung, J. F.; Park, S.; Chung, H.; Kim, J. M. Photoinduced reversible phase transition of azobenzenecontaining polydiacetylene crystals. Chem. Commun. 2016, 52, 14059-14062. (172) Xia, H.; Xu, Y.; Yang, G.; Jiang, H.; Zou, G.; Zhang, Q. A reversible multistimuli-responsive fluorescence probe and the design for combinational logic gate operations. Macromol. Rapid Commun. 2014, 35, 303-308. (173) Klajn, R. Spiropyran-based dynamic materials. Chem. Soc. Rev. 2014, 43, 148-184. (174) Baughman, R. H.; Chance, R. R. Fully Conjugated Polymer Crystals: Solid-State Synthesis and Properties of the Polydiacetylenes. Ann. N.Y. Acad. Sci. 1978, 313, 705-724. (175) Takami, K.; Mizuno, J.; Akai-kasaya, M.; Saito, A.; Aono, M.; Kuwahara, Y. Conductivity Measurement of Polydiacetylene Thin Films by Double-Tip Scanning Tunneling Microscopy. J. Phys. Chem. B 2004, 108, 16353-16356. (176) Marikhin, V. A.; Guk, E. G.; Myasnikova, L. P. New approach to achieving the potentially high conductivity of polydiacetylene. Phys. Solid State 1997, 39, 686-689. (177) Wu, Z.; Chen, Z.; Du, X.; Logan, J. M.; Sippel, J.; Nikolou, M.; Kamaras, K.; Reynolds, J. R.; Tanner, D. B.; Hebard, A. F.; Rinzler, A. G. Transparent, conductive carbon nanotube films. Science 2004, 305, 1273-1276. (178) Varghese Hansen, R.; Zhong, L.; Khor, K. A.; Zheng, L.; Yang, J. Tuneable electrochromism in weavable carbon nanotube/polydiacetylene yarns. Carbon 2016, 106, 110-117. (179) Varghese Hansen, R.; Huang, M.; Zhan, Z.; Sun, G.; Yang, J.; Zheng, L. On the study of electrochromism in multiwalled carbon nanotube–polydiacetylene composites. Carbon 2015, 90, 222-230. (180) Sun, X.; Lu, X.; Qiu, L.; Peng, H. Orienting polydiacetylene using aligned carbon nanotubes. J. Mater. Chem. C 2015, 3, 2642-2649. (181) Yang, K. S.; Yun, J. S.; Kim, J. C.; Min, J.; Park, T. J.; Ahn, J. K.; Kim, D. H. Polydiacetylene Single-Walled Carbon Nanotubes Nano-Hybrid for Cellular Imaging Applications. J. Nanosci. Nanotechnol. 2012, 12, 377-385. (182) Thauvin, C.; Rickling, S.; Schultz, P.; Celia, H.; Meunier, S.; Mioskowski, C. Carbon nanotubes as templates for polymerized lipid assemblies. Nat. Nanotechnol. 2008, 3, 743-748. (183) Peng, H.; Sun, X.; Cai, F.; Chen, X.; Zhu, Y.; Liao, G.; Chen, D.; Li, Q.; Lu, Y.; Zhu, Y.; Jia, Q. Electrochromatic carbon nanotube/polydiacetylene nanocomposite fibres. Nat. Nanotechnol. 2009, 4, 738-741. (184) Zhang, W.; Xu, H.; Chen, Y.; Cheng, S.; Fan, L. J. Polydiacetylenepolymethylmethacrylate/graphene composites as one-shot, visually observable, and semiquantative electrical current sensing materials. ACS Appl. Mater. Interfaces 2013, 5, 4603-4606. (185) Kang, D. H.; Jung, H. S.; Ahn, N.; Yang, S. M.; Seo, S.; Suh, K. Y.; Chang, P. S.; Jeon, N. L.; Kim, J.; Kim, K. Janus-compartmental alginate microbeads having polydiacetylene liposomes and magnetic nanoparticles for visual lead(II) detection. ACS Appl. Mater. Interfaces 2014, 6, 10631-10637. (186) de Oliveira, T. V.; Soares Nde, F.; de Andrade, N. J.; Silva, D. J.; Medeiros, E. A.; Badaro, A. T. Application of PCDA/SPH/CHO/Lysine vesicles to detect pathogenic bacteria in chicken. Food Chem. 2015, 172, 428-432. (187) Ma, X.; Sheng, Z.; Jiang, L. Sensitive naked-eye detection of Hg2+ based on the aggregation and filtration of thymine functionalized vesicles caused by selective interaction between thymine and Hg2+. Analyst 2014, 139, 3365-3368. (188) Seo, S.; Kwon, M. S.; Phillips, A. W.; Seo, D.; Kim, J. Highly sensitive turnon biosensors by regulating fluorescent dye assembly on liposome surfaces. Chem. Commun. 2015, 51, 10229-10232. (189) Zhou, J.; Tun, T. N.; Hong, S. H.; Mercer-Chalmers, J. D.; Laabei, M.; Young, A. E.; Jenkins, A. T. Development of a prototype wound dressing technology which can detect and report colonization by pathogenic bacteria. Biosens. Bioelectron. 2011, 30, 67-72. (190) Zhou, J.; Loftus, A. L.; Mulley, G.; Jenkins, A. T. A thin film detection/response system for pathogenic bacteria. J. Am. Chem. Soc. 2010, 132, 6566-6570.

22

ACS Paragon Plus Environment

Page 23 of 26 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials (234) Lu, Y.; Yang, Y.; Sellinger, A.; Lu, M.; Huang, J.; Fan, H.; Haddad, R.; Lopez, G.; Burns, A. R.; Sasaki, D. Y.; Shelnutt, J.; Brinker, C. J. Self-assembly of mesoscopically ordered chromatic polydiacetylene/silica nanocomposites. Nature 2001, 410, 913-917. (235) Yang, Y.; Lu, Y.; Lu, M.; Huang, J.; Haddad, R.; Xomeritakis, G.; Liu, N.; Malanoski, A. P.; Sturmayr, D.; Fan, H.; Sasaki, D. Y.; Assink, R. A.; Shelnutt, J. A.; van Swol, F.; Lopez, G. P.; Burns, A. R.; Brinker, C. J. Functional nanocomposites prepared by self-assembly and polymerization of diacetylene surfactants and silicic acid. J. Am. Chem. Soc. 2003, 125, 1269-1277. (236) Gill, I.; Ballesteros, A. Immunoglobulin-polydiacetylene sol-gel nanocomposites as solid-state chromatic biosensors. Angew. Chem. Int. Ed. 2003, 42, 32643267. (237) Yamanaka, S. A.; Charych, D. H.; Loy, D. A.; Sasaki, D. Y. Solid Phase Immobilization of Optically Responsive Liposomes in Sol-Gel Materials for Chemical and Biological Sensing. Langmuir 1997, 13, 5049-5053. (238) Peng, H.; Tang, J.; Pang, J.; Chen, D.; Yang, L.; Ashbaugh, H. S.; Brinker, C. J.; Yang, Z.; Lu, Y. Polydiacetylene/silica nanocomposites with tunable mesostructure and thermochromatism from diacetylenic assembling molecules. J. Am. Chem. Soc. 2005, 127, 12782-12783. (239) Nopwinyuwong, A.; Boonsupthip, W.; Pechyen, C.; Suppakul, P. Formation of Polydiacetylene/Silica Nanocomposite as a Colorimetric Indicator: Effect of Time and Temperature. Adv. Polym. Tech. 2013, 32, E724-E731. (240) Park, D.-H.; Park, B. J.; Kim, J.-M. Creation of functional polydiacetylene images on paper using inkjet printing technology. Macromol. Res. 2016, 24, 943950. (241) Yoon, B.; Ham, D. Y.; Yarimaga, O.; An, H.; Lee, C. W.; Kim, J. M. Inkjet printing of conjugated polymer precursors on paper substrates for colorimetric sensing and flexible electrothermochromic display. Adv. Mater. 2011, 23, 54925497. (242) Yoon, B.; Shin, H.; Yarimaga, O.; Ham, D.-Y.; Kim, J.; Park, I. S.; Kim, J.M. An inkjet-printable microemulsion system for colorimetric polydiacetylene supramolecules on paper substrates. J. Mater. Chem. 2012, 22, 8680-8686. (243) Pumtang, S.; Siripornnoppakhun, W.; Sukwattanasinitt, M.; Ajavakom, A. Solvent colorimetric paper-based polydiacetylene sensors from diacetylene lipids. J. Colloid Interface Sci. 2011, 364, 366-372. (244) Zhang, P.; Zhang, C.; Shu, B. Micropatterned paper devices using amineterminated polydiacetylene vesicles as colorimetric probes for enhanced detection of double-stranded DNA. Sens. Actuator B-Chem. 2016, 236, 27-34. (245) Jung, H. S.; Verwilst, P.; Kim, W. Y.; Kim, J. S. Fluorescent and colorimetric sensors for the detection of humidity or water content. Chem. Soc. Rev. 2016, 45, 1242-1256. (246) Lee, J.; Pyo, M.; Lee, S. H.; Kim, J.; Ra, M.; Kim, W. Y.; Park, B. J.; Lee, C. W.; Kim, J. M. Hydrochromic conjugated polymers for human sweat pore mapping. Nat. Commun. 2014, 5, 3736. (247) Park, D.-H.; Jeong, W.; Seo, M.; Park, B. J.; Kim, J.-M. Inkjet-Printable Amphiphilic Polydiacetylene Precursor for Hydrochromic Imaging on Paper. Adv. Funct. Mater. 2016, 26, 498-506. (248) Gou, M.; Qu, X.; Zhu, W.; Xiang, M.; Yang, J.; Zhang, K.; Wei, Y.; Chen, S. Bio-inspired detoxification using 3D-printed hydrogel nanocomposites. Nat. Commun. 2014, 5, 3774. (249) Chae, S. K.; Park, H.; Yoon, J.; Lee, C. H.; Ahn, D. J.; Kim, J. M. Polydiacetylene Supramolecules in Electrospun Microfibers: Fabrication, Micropatterning, and Sensor Applications. Adv. Mater. 2007, 19, 521-524. (250) Yapor, J. P.; Alharby, A.; Gentry-Weeks, C.; Reynolds, M. M.; Alam, A.; Li, Y. V. Polydiacetylene Nanofiber Composites as a Colorimetric Sensor Responding To Escherichia coli and pH. ACS Omega 2017, 2, 7334-7342. (251) Moazeni, N.; Merati, A. A.; Latifi, M.; Sadrjahani, M.; Rouhani, S. Fabrication and characterization of polydiacetylene supramolecules in electrospun polyvinylidene fluoride nanofibers with dual colorimetric and piezoelectric responses. Polymer 2018, 134, 211-220. (252) Wu, J.; Lu, X.; Shan, F.; Guan, J.; Lu, Q. Polydiacetylene-embedded supramolecular electrospun fibres for a colourimetric sensor of organic amine vapour. RSC Adv. 2013, 3, 22841-228444. (253) Jeon, H.; Lee, J.; Kim, M. H.; Yoon, J. Polydiacetylene-based electrospun fibers for detection of HCl gas. Macromol. Rapid Commun. 2012, 33, 972-976. (254) Lee, S.; Cho, Y.; Ye, B. U.; Baik, J. M.; Kim, M. H.; Yoon, J. Unprecedented colorimetric responses of polydiacetylenes driven by plasma induced polymerization and their patterning applications. Chem. Commun. 2014, 50, 1244712449. (255) Parambath Kootery, K.; Jiang, H.; Kolusheva, S.; Vinod, T. P.; Ritenberg, M.; Zeiri, L.; Volinsky, R.; Malferrari, D.; Galletti, P.; Tagliavini, E.; Jelinek, R. Poly(methyl methacrylate)-supported polydiacetylene films: unique chromatic

(213) Wu, W.; Zhang, J.; Zheng, M.; Zhong, Y.; Yang, J.; Zhao, Y.; Wu, W.; Ye, W.; Wen, J.; Wang, Q.; Lu, J. An aptamer-based biosensor for colorimetric detection of Escherichia coli O157:H7. PLoS One 2012, 7, e48999. (214) Lee, S.; Cheng, H.; Chi, M.; Xu, Q.; Chen, X.; Eom, C. Y.; James, T. D.; Park, S.; Yoon, J. Sensing and antibacterial activity of imidazolium-based conjugated polydiacetylenes. Biosens. Bioelectron. 2016, 77, 1016-1019. (215) Stencel-Baerenwald, J. E.; Reiss, K.; Reiter, D. M.; Stehle, T.; Dermody, T. S. The sweet spot: defining virus-sialic acid interactions. Nat. Rev. Microbiol. 2014, 12, 739-749. (216) Reichert, A.; Nagy, J. O.; Spevak, W.; Charych, D. Polydiacetylene Liposomes Functionalized with Sialic Acid Bind and Colorimetrically Detect Influenza Virus. J. Am. Chem. Soc. 1995, 117, 829-830. (217) Charych, D.; Cheng, Q.; Reichert, A.; Kuziemko, G.; Stroh, M.; Nagy, J. O.; Spevak, W.; Stevens, R. C. A ‘litmus test’ for molecular recognition using artificial membranes. Chem. Biol. 1996, 3, 113-120. (218) Deng, J.; Sheng, Z.; Zhou, K.; Duan, M.; Yu, C. Y.; Jiang, L. Construction of effective receptor for recognition of avian influenza H5N1 protein HA1 by assembly of monohead glycolipids on polydiacetylene vesicle surface. Bioconjugate Chem. 2009, 20, 533-537. (219) Song, J.; Cheng, Q.; Zhu, S.; Stevens, R. C. “Smart” Materials for Biosensing Devices: Cell-Mimicking Supramolecular Assemblies and Colorimetric Detection of Pathogenic Agents. Biomed. Microdevices 2002, 4, 213-221. (220) Dong, W.; Luo, J.; He, H.; Jiang, L. A reinforced composite structure composed of polydiacetylene assemblies deposited on polystyrene microspheres and its application to H5N1 virus detection. Int. J. Nanomed. 2013, 8, 221-232. (221) Jeong, J.-p.; Cho, E.; Yun, D.; Kim, T.; Lee, I.-S.; Jung, S. Label-Free Colorimetric Detection of Influenza Antigen Based on an Antibody-Polydiacetylene Conjugate and Its Coated Polyvinylidene Difluoride Membrane. Polymers 2017, 9, 127. (222) Jiang, L.; Luo, J.; Dong, W.; Wang, C.; Jin, W.; Xia, Y.; Wang, H.; Ding, H.; Jiang, L.; He, H. Development and evaluation of a polydiacetylene based biosensor for the detection of H5 influenza virus. J. Virol. Methods 2015, 219, 38-45. (223) Park, M. K.; Kim, K. W.; Ahn, D. J.; Oh, M. K. Label-free detection of bacterial RNA using polydiacetylene-based biochip. Biosens. Bioelectron. 2012, 35, 44-49. (224) Song, S.; Ha, K.; Guk, K.; Hwang, S.-G.; Choi, J. M.; Kang, T.; Bae, P.; Jung, J.; Lim, E.-K. Colorimetric detection of influenza A (H1N1) virus by a peptide-functionalized polydiacetylene (PEP-PDA) nanosensor. RSC Adv. 2016, 6, 48566-48570. (225) Jeong, J.-P.; Cho, E.; Lee, S.-C.; Kim, T.; Song, B.; Lee, I.-S.; Jung, S. Detection of Foot-and-Mouth Disease Virus Using a Polydiacetylene Immunosensor on Solid-Liquid Phase. Macromol. Mater. Eng. 2018, 303, 1700640. (226) Rangin, M.; Basu, A. Lipopolysaccharide identification with functionalized polydiacetylene liposome sensors. J. Am. Chem. Soc. 2004, 126, 5038-5039. (227) Kolusheva, S.; Molt, O.; Herm, M.; Schrader, T.; Jelinek, R. Selective detection of catecholamines by synthetic receptors embedded in chromatic polydiacetylene vesicles. J. Am. Chem. Soc. 2005, 127, 10000-10001. (228) Kamphan, A.; Gong, C.; Maiti, K.; Sur, S.; Traiphol, R.; Arya, D. P. Utilization of chromic polydiacetylene assemblies as a platform to probe specific binding between drug and RNA. RSC Adv. 2017, 7, 41435-41443. (229) Pernia Leal, M.; Assali, M.; Cid, J. J.; Valdivia, V.; Franco, J. M.; Fernandez, I.; Pozo, D.; Khiar, N. Synthesis of 1D-glyconanomaterials by a hybrid noncovalent-covalent functionalization of single wall carbon nanotubes: a study of their selective interactions with lectins and with live cells. Nanoscale 2015, 7, 1925919272. (230) Wang, Y.; Pei, H.; Jia, Y.; Liu, J.; Li, Z.; Ai, K.; Lu, Z.; Lu, L. Synergistic Tailoring of Electrostatic and Hydrophobic Interactions for Rapid and Specific Recognition of Lysophosphatidic Acid, an Early-Stage Ovarian Cancer Biomarker. J. Am. Chem. Soc. 2017, 139, 11616-11621. (231) Wang, D. E.; Yan, J.; Jiang, J.; Liu, X.; Tian, C.; Xu, J.; Yuan, M. S.; Han, X.; Wang, J. Polydiacetylene liposomes with phenylboronic acid tags: a fluorescence turn-on sensor for sialic acid detection and cell-surface glycan imaging. Nanoscale 2018, 10, 4570-4578. (232) Li, X.; Liu, W.; Yue, X.; Song, P.; Yin, Y.; Meng, M.; Xi, R. A competitive immunoassay using hapten-modified polydiacetylene vesicles for homogeneous and sensitive detection of sodium benzoate. Sens. Actuator B-Chem. 2018, 258, 1060-1065. (233) Choi, J.-M.; Yoon, B.; Choi, K.; Seol, M.-L.; Kim, J.-M.; Choi, Y.-K. Micropatterning Polydiacetylene Supramolecular Vesicles on Glass Substrates using a Pre-Patterned Hydrophobic Thin Film. Macromol. Chem. Phys. 2012, 213, 610616.

23

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 26

(277) Wang, T.; Guo, Y.; Wan, P.; Sun, X.; Zhang, H.; Yu, Z.; Chen, X. A flexible transparent colorimetric wrist strap sensor. Nanoscale 2017, 9, 869-874. (278) Baek, S.; Song, S.; Lee, J.; Kim, J.-M. Nanoscale diameter control of sensory polydiacetylene nanoparticles on microfluidic chip for enhanced fluorescence signal. Sens. Actuator B-Chem. 2016, 230, 623-629. (279) Eo, S.-H.; Song, S.; Yoon, B.; Kim, J.-M. A Microfluidic Conjugated-Polymer Sensor Chip. Adv. Mater. 2008, 20, 1690-1694. (280) Ma, J.; Lee, S. M.; Yi, C.; Li, C. W. Controllable synthesis of functional nanoparticles by microfluidic platforms for biomedical applications - a review. Lab Chip 2017, 17, 209-226. (281) Yoo, I.; Song, S.; Yoon, B.; Kim, J. M. Size-controlled fabrication of polydiacetylene-embedded microfibers on a microfluidic chip. Macromol. Rapid Commun. 2012, 33, 1256-1261. (282) Guo, C. X.; Boullanger, P.; Liu, T.; Jiang, L. Size effect of polydiacetylene vesicles functionalized with glycolipids on their colorimetric detection ability. J. Phys. Chem. B 2005, 109, 18765-18771. (283) de Oliveira, T. V.; Soares, N. d. F. F.; Coimbra, J. S. d. R.; de Andrade, N. J.; Moura, L. G.; Medeiros, E. A. A.; de Medeiros, H. S. Stability and sensitivity of polydiacetylene vesicles to detect Salmonella. Sens. Actuator B-Chem. 2015, 221, 653-658. (284) Jang, H.; Lee, P.; Kim, S.; Kim, S. M.; Jeon, T.-J. An antibacterial microfluidic system with fish gill structure for the detection of Staphylococcus via enzymatic reaction on a chromatic polydiacetylene material caused by lysostaphin. Microchim. Acta 2017, 184, 4563-4569. (285) Chae, S.; Lee, J. P.; Kim, J.-M. Mechanically Drawable Thermochromic and Mechanothermochromic Polydiacetylene Sensors. Adv. Funct. Mater. 2016, 26, 1769-1776. (286) Park, J.; Ku, S. K.; Seo, D.; Hur, K.; Jeon, H.; Shvartsman, D.; Seok, H. K.; Mooney, D. J.; Lee, K. Label-free bacterial detection using polydiacetylene liposomes. Chem. Commun. 2016, 52, 10346-10349. (287) Hu, W.; Chen, Y.; Jiang, H.; Li, J.; Zou, G.; Zhang, Q.; Zhang, D.; Wang, P.; Ming, H. Optical waveguide based on a polarized polydiacetylene microtube. Adv. Mater. 2014, 26, 3136-3141. (288) Yang, G.; Hu, W.; Xia, H.; Zou, G.; Zhang, Q. Highly selective and reproducible detection of picric acid in aqueous media, based on a polydiacetylene microtube optical waveguide. J. Mater. Chem. A 2014, 2, 15560-15565. (289) Zhu, Y.; Qiu, D.; Yang, G.; Wang, M.; Zhang, Q.; Wang, P.; Ming, H.; Zhang, D.; Yu, Y.; Zou, G.; Badugu, R.; Lakowicz, J. R. Selective and sensitive detection of MiRNA-21 based on gold-nanorod functionalized polydiacetylene microtube waveguide. Biosens. Bioelectron. 2016, 85, 198-204. (290) Wang, M.; Yu, Y.; Liu, F.; Ren, L.; Zhang, Q.; Zou, G. Single polydiacetylene microtube waveguide platform for discriminating microRNA-215 expression levels in clinical gastric cancerous, paracancerous and normal tissues. Talanta 2018, 188, 27-34. (291) Xia, H.; Chen, Y.; Yang, G.; Zou, G.; Zhang, Q.; Zhang, D.; Wang, P.; Ming, H. Optical modulation of waveguiding in spiropyran-functionalized polydiacetylene microtube. ACS Appl. Mater. Interfaces 2014, 6, 15466-15471. (292) Yang, G.; Zhang, Y.; Xia, H.; Zou, G.; Zhang, Q. Multiconfigurable logic gate operation in 1D polydiacetylene microtube waveguide. RSC Adv. 2016, 6, 53794-53799. (293) Oh, S.; Uh, K.; Jeon, S.; Kim, J.-M. A Free-Standing Self-Assembled Tubular Conjugated Polymer Sensor. Macromolecules 2016, 49, 5841-5848. (294) Xu, Q.; Lee, K. M.; Wang, F.; Yoon, J. Visual detection of copper ions based on azide- and alkyne-functionalized polydiacetylene vesicles. J. Mater. Chem. 2011, 21, 15214-15217. (295) Jonas, U.; Shah, K.; Norvez, S.; Charych, D. H. Reversible Color Switching and Unusual Solution Polymerization of Hydrazide-Modified Diacetylene Lipids. J. Am. Chem. Soc. 1999, 121, 4580-4588. (296) Xu, Q.; Lee, S.; Cho, Y.; Kim, M. H.; Bouffard, J.; Yoon, J. Polydiacetylenebased colorimetric and fluorescent chemosensor for the detection of carbon dioxide. J. Am. Chem. Soc. 2013, 135, 17751-17754. (297) Kim, K. W.; Lee, J. M.; Kwon, Y. M.; Choi, T.-Y.; Kim, J. Y. H.; Bae, S.; Song, J.-A. Polyamine-Functionalized Polydiacetylene (PDA) Vesicles for Colorimetric Sensing of Carbon Dioxide. Macromol. Res. 2018, 26, 284-290. (298) Lu, S.; Jia, C.; Duan, X.; Zhang, X.; Luo, F.; Han, Y.; Huang, H. Polydiacetylene vesicles for hydrogen peroxide detection. Colloid Surf. A-Physicochem. Eng. Asp. 2014, 443, 488-491. (299) Jia, C.; Tang, J.; Lu, S.; Han, Y.; Huang, H. Enhanced Sensitivity for Hydrogen Peroxide Detection: Polydiacetylene Vesicles with Phenylboronic Acid Head Group. J. Fluoresc. 2016, 26, 121-127.

transitions and molecular sensing. ACS Appl. Mater. Interfaces 2014, 6, 86138620. (256) Lee, B.-y.; Kim, J.; Kim, W. J.; Kim, J. K. Dual functional membrane capable of both visual sensing and blocking of waterborne virus. J. Membr. Sci. 2018, 549, 680-685. (257) Trachtenberg, A.; Malka, O.; Kootery, K. P.; Beglaryan, S.; Malferrari, D.; Galletti, P.; Prati, S.; Mazzeo, R.; Tagliavini, E.; Jelinek, R. Colorimetric analysis of painting materials using polymer-supported polydiacetylene films. New J. Chem. 2016, 40, 9054-9059. (258) Kim, J. M.; Lee, Y. B.; Chae, S. K.; Ahn, D. J. Patterned Color and Fluorescent Images with Polydiacetylene Supramolecules Embedded in Poly(vinyl alcohol) Films. Adv. Funct. Mater. 2006, 16, 2103-2109. (259) Seo, M.; Park, D.-H.; Park, B. J.; Kim, J.-M. Flexible patch-type hydrochromic polydiacetylene sensor for human sweat pore mapping. J. Appl. Polym. Sci. 2017, 134, 44419. (260) Park, E. Y.; Kim, J. W.; Ahn, D. J.; Kim, J.-M. A Polydiacetylene Supramolecular System That Emits Red, Green, and Blue Fluorescence. Macromol. Rapid Commun. 2007, 28, 171-175. (261) Yarimaga, O.; Lee, S.; Kim, J. M.; Choi, Y. K. Fabrication of Patterned Polydiacetylene Composite Films Using a Replica-Molding (REM) Technique. Macromol. Rapid Commun. 2010, 31, 270-274. (262) Lee, J.; Yarimaga, O.; Lee, C. H.; Choi, Y.-K.; Kim, J.-M. Network Polydiacetylene Films: Preparation, Patterning, and Sensor Applications. Adv. Funct. Mater. 2011, 21, 1032-1039. (263) Davis, B. W.; Burris, A. J.; Niamnont, N.; Hare, C. D.; Chen, C. Y.; Sukwattanasinitt, M.; Cheng, Q. Dual-mode optical sensing of organic vapors and proteins with polydiacetylene (PDA)-embedded electrospun nanofibers. Langmuir 2014, 30, 9616-9622. (264) Jun, B. H.; Baek, J.; Kang, H.; Park, Y. J.; Jeong, D. H.; Lee, Y. S. Preparation of polydiacetylene immobilized optically encoded beads. J. Colloid Interface Sci. 2011, 355, 29-34. (265) Lee, J.; Kim, J. Multiphasic Sensory Alginate Particle Having Polydiacetylene Liposome for Selective and More Sensitive Multitargeting Detection. Chem. Mater. 2012, 24, 2817-2822. (266) Kim, H.-M.; Kang, Y.-L.; Chung, W.-J.; Kyeong, S.; Jeong, S.; Kang, H.; Jeong, C.; Rho, W.-Y.; Kim, D.-H.; Jeong, D. H.; Lee, Y.-S.; Jun, B.-H. Ligand immobilization on polydiacetylene-coated and surface-enhanced Raman scattering-encoded beads for label-free detection. J. Ind. Eng. Chem. 2015, 21, 158-162. (267) Lee, H.-Y.; Tiwari, K. R.; Raghavan, S. R. Biopolymer capsules bearing polydiacetylenic vesicles as colorimetric sensors of pH and temperature. Soft Matter 2011, 7, 3273-3276. (268) Yun, D.; Cho, E.; Dindulkar, S. D.; Jung, S. Succinoglycan Octasaccharide Conjugated Polydiacetylene-Doped Alginate Beads for Barium (II) Detection. Macromol. Mater. Eng. 2016, 301, 805-811. (269) Nie, Q.; Zhang, Y.; Zhang, J.; Zhang, M. Immobilization of polydiacetylene onto silica microbeads for colorimetric detection. J. Mater. Chem. 2006, 16, 546549. (270) Jung, S. H.; Jang, H.; Lim, M. C.; Kim, J. H.; Shin, K. S.; Kim, S. M.; Kim, H. Y.; Kim, Y. R.; Jeon, T. J. Chromatic biosensor for detection of phosphinothricin acetyltransferase by use of polydiacetylene vesicles encapsulated within automatically generated immunohydrogel beads. Anal. Chem. 2015, 87, 2072-2078. (271) Lee, S.; Lee, J.; Lee, D. W.; Kim, J. M.; Lee, H. A 3D networked polydiacetylene sensor for enhanced sensitivity. Chem. Commun. 2016, 52, 926-929. (272) Potai, R.; Faisadcha, K.; Traiphol, R.; Traiphol, N. Controllable thermochromic and phase transition behaviors of polydiacetylene/zinc(II) ion/zinc oxide nanocomposites via photopolymerization: An insight into the molecular level. Colloid Surf. A-Physicochem. Eng. Asp. 2018, 555, 27-36. (273) Chanakul, A.; Traiphol, N.; Faisadcha, K.; Traiphol, R. Dual colorimetric response of polydiacetylene/zinc oxide nanocomposites to low and high pH. J. Colloid Interface Sci. 2014, 418, 43-51. (274) Toommee, S.; Traiphol, R.; Traiphol, N. High color stability and reversible thermochromism of polydiacetylene/zinc oxide nanocomposite in various organic solvents and polymer matrices. Colloid Surf. A-Physicochem. Eng. Asp. 2015, 468, 252-261. (275) Chanakul, A.; Traiphol, R.; Traiphol, N. Colorimetric sensing of various organic acids by using polydiacetylene/zinc oxide nanocomposites: Effects of polydiacetylene and acid structures. Colloid Surf. A-Physicochem. Eng. Asp. 2016, 489, 9-18. (276) Chanakul, A.; Traiphol, R.; Traiphol, N. Utilization of polydiacetylene/zinc oxide nanocomposites to detect and differentiate organic bases in various media. J. Ind. Eng. Chem. 2017, 45, 215-222.

24

ACS Paragon Plus Environment

Page 25 of 26 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials materials formed by cross-linking fullerene derivatives. Angew. Chem. Int. Ed. 2009, 48, 2166-2170. (320) Kim, C.; Lee, S. J.; Lee, I. H.; Kim, K. T.; Song, H. H.; Jeon, H.-J. Stabilization of Supramolecular Nanostructures Induced by Self-Assembly of Dendritic Building Blocks. Chem. Mater. 2003, 15, 3638-3642. (321) Sui, G.; Micic, M.; Huo, Q.; Leblanc, R. M. Studies of a novel polymerizable amphiphilic dendrimer. Colloid Surf. A-Physicochem. Eng. Asp. 2000, 171, 185197. (322) Shirakawa, M.; Fujita, N.; Shinkai, S. A stable single piece of unimolecularly pi-stacked porphyrin aggregate in a thixotropic low molecular weight gel: a one-dimensional molecular template for polydiacetylene wiring up to several tens of micrometers in length. J. Am. Chem. Soc. 2005, 127, 4164-4165. (323) Sutapin, C.; Mantaranon, N.; Chirachanchai, S. Eight-armed polydiacetylene under benzoxazine dimer branched polylactide: a structural combination for reversible thermochromic effects and a model case for free-standing poly(lactic acid) films. J. Mater. Chem. C 2017, 5, 8288-8294. (324) Nierengarten, J.-F.; Setayesh, S. Towards polymerizable fullerene derivatives to stabilize the initially formed phases in bulk-heterojunction solar cells. New J. Chem. 2006, 30, 313-316. (325) Singh, Y.; Jayaraman, N. Visual Detection of pH and Biomolecular Interactions at Micromolar Concentrations Aided by a Trivalent Diacetylene-Based Vesicle. Macromol. Chem. Phys. 2017, 218, 1700039. (326) Singh, Y.; Jayaraman, N. Multicolor Reversible Thermochromic Properties of Gallic Acid-Cored Polydiacetylenes Appended with Poly(alkyl aryl ether) Dendrons. Macromol. Chem. Phys. 2016, 217, 940-950.

(300) Kim, M.; Shin, Y. J.; Hwang, S. W.; Shin, M. J.; Shin, J. S. Chromatic detection of glucose using polymerization of diacetylene vesicle. J. Appl. Polym. Sci. 2018, 135, 46394. (301) Kim, D. Y.; Lee, S. A.; Jung, D.; Koo, J.; Soo Kim, J.; Yu, Y. T.; Lee, C. R.; Jeong, K. U. Topochemical polymerization of dumbbell-shaped diacetylene monomers: relationship between chemical structure, molecular packing structure, and gelation property. Soft Matter 2017, 13, 5759-5766. (302) Kim, B. G.; Kim, S.; Seo, J.; Oh, N. K.; Zin, W. C.; Park, S. Y. Supramolecular assembly of fluorescent phasmidic diacetylene and its photopolymerization. Chem. Commun. 2003, 2306-2307. (303) Haridas, V.; Sharma, Y. K.; Creasey, R.; Sahu, S.; Gibson, C. T.; Voelcker, N. H. Gelation and topochemical polymerization of peptide dendrimers. New J. Chem. 2011, 35, 303-309. (304) Yoon, B.; Park, I. S.; Shin, H.; Park, H. J.; Lee, C. W.; Kim, J. M. A litmustype colorimetric and fluorometric volatile organic compound sensor based on inkjet-printed polydiacetylenes on paper substrates. Macromol. Rapid Commun. 2013, 34, 731-735. (305) Ampornpun, S.; Montha, S.; Tumcharern, G.; Vchirawongkwin, V.; Sukwattanasinitt, M.; Wacharasindhu, S. Odd–Even and Hydrophobicity Effects of Diacetylene Alkyl Chains on Thermochromic Reversibility of Symmetrical and Unsymmetrical Diyndiamide Polydiacetylenes. Macromolecules 2012, 45, 90389045. (306) Yoon, B.; Shin, H.; Kang, E. M.; Cho, D. W.; Shin, K.; Chung, H.; Lee, C. W.; Kim, J. M. Inkjet-compatible single-component polydiacetylene precursors for thermochromic paper sensors. ACS Appl. Mater. Interfaces 2013, 5, 45274535. (307) Lee, S.; Lee, J.; Lee, M.; Cho, Y. K.; Baek, J.; Kim, J.; Park, S.; Kim, M. H.; Chang, R.; Yoon, J. Construction and Molecular Understanding of an Unprecedented, Reversibly Thermochromic Bis-Polydiacetylene. Adv. Funct. Mater. 2014, 24, 3699-3705. (308) Baek, J.; Joung, J. F.; Lee, S.; Rhee, H.; Kim, M. H.; Park, S.; Yoon, J. Origin of the Reversible Thermochromic Properties of Polydiacetylenes Revealed by Ultrafast Spectroscopy. J. Phys. Chem. Lett. 2016, 7, 259-265. (309) Meng, Y.; Jiang, J.; Liu, M. Self-assembled nanohelix from a bolaamphiphilic diacetylene via hydrogelation and selective responsiveness towards amino acids and nucleobases. Nanoscale 2017, 9, 7199-7206. (310) Xu, Y.; Smith, M. D.; Geer, M. F.; Pellechia, P. J.; Brown, J. C.; Wibowo, A. C.; Shimizu, L. S. Thermal reaction of a columnar assembled diacetylene macrocycle. J. Am. Chem. Soc. 2010, 132, 5334-5335. (311) Xu, W. L.; Smith, M. D.; Krause, J. A.; Greytak, A. B.; Ma, S.; Read, C. M.; Shimizu, L. S. Single Crystal to Single Crystal Polymerization of a Self-Assembled Diacetylene Macrocycle Affords Columnar Polydiacetylenes. Cryst. Growth Des. 2014, 14, 993-1002. (312) Zhang, X.; Deng, C.; Wang, M.; Liu, X.; Lin, C.; Peng, L.; Wang, L. Topochemical polymerisation of assembled diacetylene macrocycle bearing dibenzylphosphine oxide in solid state. Supramol. Chem. 2016, 29, 94-101. (313) Kikuchi, K.; Tatewaki, Y.; Okada, S. Synthesis and Solid-State Polymerization of a Macrocyclic Compound with Two Butadiyne Units. Bull. Chem. Soc. Jpn. 2017, 90, 387-394. (314) Kantha, C.; Kim, H.; Kim, Y.; Heo, J.-M.; Joung, J. F.; Park, S.; Kim, J.-M. Topochemical polymerization of macrocyclic diacetylene with a naphthalene moiety for a tubular-shaped polydiacetylene chromophore. Dyes Pigment. 2018, 154, 199-204. (315) Heo, J.-M.; Kim, Y.; Han, S.; Joung, J. F.; Lee, S.-h.; Han, S.; Noh, J.; Kim, J.; Park, S.; Lee, H.; Choi, Y. M.; Jung, Y.-S.; Kim, J.-M. Chromogenic Tubular Polydiacetylenes from Topochemical Polymerization of Self-Assembled Macrocyclic Diacetylenes. Macromolecules 2017, 50, 900-913. (316) Burilov, V. A.; Valiyakhmetova, A. M.; Aukhadieva, R. I.; Solovieva, S. E.; Antipin, I. S. Synthesis of new p-tert-butylcalix[4]arene derivatives containing photopolymerizable 1,3-butadiyne fragments. Russ. J. Gen. Chem. 2017, 87, 1946-1951. (317) Burilov, V.; Valiyakhmetova, A.; Mironova, D.; Safiullin, R.; Kadirov, M.; Ivshin, K.; Kataeva, O.; Solovieva, S.; Antipin, I. “Clickable” thiacalix[4]arene derivatives bearing polymerizable 1,3-butadiyne fragments: synthesis and incorporation into polydiacetylene vesicles. RSC Adv. 2016, 6, 44873-44877. (318) Burilov, V.; Valiyakhmetova, A.; Mironova, D.; Sultanova, E.; Evtugyn, V.; Osin, Y.; Katsyuba, S.; Burganov, T.; Solovieva, S.; Antipin, I. Novel amphiphilic conjugates of p-tert-butylthiacalix[4]arene with 10,12-pentacosadiynoic acid in 1,3-alternate stereoisomeric form. Synthesis and chromatic properties in the presence of metal ions. New J. Chem. 2018, 42, 2942-2951. (319) Wang, J.; Shen, Y.; Kessel, S.; Fernandes, P.; Yoshida, K.; Yagai, S.; Kurth, D. G.; Mohwald, H.; Nakanishi, T. Self-assembly made durable: water-repellent

25

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TOC

ACS Paragon Plus Environment

Page 26 of 26