Photo-responsive Behavior of Wormlike Micelles Constructed by

Mar 11, 2019 - The photo-responsive wormlike micelles constructed by Gemini surfactants and cinnamate derivatives play a great role in the field of sm...
0 downloads 0 Views 2MB Size
Subscriber access provided by ECU Libraries

Interface Components: Nanoparticles, Colloids, Emulsions, Surfactants, Proteins, Polymers

Photo-responsive Behavior of Wormlike Micelles Constructed by Gemini Surfactant 12-3-12·2Br- and Different Cinnamate Derivatives Yan Tu, Qizhou Chen, Yazhuo Shang, Hongni Teng, and Honglai Liu Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.8b04290 • Publication Date (Web): 11 Mar 2019 Downloaded from http://pubs.acs.org on March 12, 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 31 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

Langmuir

Photo-responsive Behavior of Wormlike Micelles Constructed by Gemini Surfactant 12-3-12·2Br- and Different Cinnamate Derivatives Yan Tu†

†Key

Qizhou Chen†

Yazhuo Shang†* Hongni Teng††

Honglai Liu†

Laboratory for Advanced Materials, School of Chemistry & Molecular

Engineering, East China University of Science and Technology, Shanghai 200237, China ††Department

of Applied Chemistry, College of Chemical and Environmental

Engineering, Shandong University of Science and Technology, Qingdao 266510, China

Yazhuo Shang†* (Corresponding Author): [email protected] (Y. Shang)

ACS Paragon Plus Environment

Langmuir 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

ABSTRACT The photo-responsive wormlike micelles constructed by Gemini surfactants and cinnamate derivatives play a great role in the field of smart materials. However, how the structure of cinnamate derivatives to adjust the photo-responsive behavior of micelles is still a hot spot for scientists to research. Here, three kinds of aromatic salts with different ortho substituted groups including trans-o-methoxy cinnamate (transOMCA), trans-o-hydroxy cinnamate (trans-OHCA) and trans-cinnamate (trans-CA) were introduced into Gemini surfactant 12-3-12·2Br- aqueous solutions to construct photo-responsive wormlike micelles through their non-covalent interactions. Their properties were researched using the rheological method, cryo-TEM, 1HNMR and 2D NOESY spectrum. The results show that these cinnamate derivatives could well construct wormlike micelles with 12-3-12·2Br-. Furthermore, subtle differences in the ortho substitutes’ structure make a significant effect on the photo-responsive behavior of formed wormlike micelles. Specifically, the zero viscosity (0) of 40mM 12-312·2Br-/24mM trans-OHCA mixed solution decreases from 26.72Pa·s to 2.6Pa·s with the shorter of the length of wormlike micelles after UV irradiation. Corresponding, the

0 of the same ratio of 12-3-12·2Br-/trans-OMCA decreases from 2.42Pa·s to 0.06Pa·s as well as the wormlike micelles are transited into rod-like micelles and even spherical micelles after the same UV irradiation time. While the variation of wormlike micelles in 12-3-12·2Br-/trans-CA system induced by UV light is not obvious with 0 maintaining around 2.89Pa·s. This study will help us better understand the effects of chemical groups on macro-phenomena and micro-interaction for micellar systems. It provides a theoretical basis for the construction of photo-responsive micelles, thus widening their application in the field of soft materials.

Keywords: Photo-responsive; Wormlike micelles; Gemini surfactant; Cinnamate derivatives

ACS Paragon Plus Environment

Page 2 of 31

Page 3 of 31 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

Langmuir

INTRODUCTION Wormlike micelles constructed with surfactants have attracted great interests of scientists due to their unique rheological behavior and extensive applications1-4. Based on the excellent and adjustable viscoelastic properties, stimuli-responsive wormlike micelles response to external triggers such as pH 5, CO2 6-7, temperature8, light9-12, redox reaction13 or magnetic field14 have attracted great attention in the field of soft materials 15-17.

Among them, the photo-responsive viscoelastic fluids composed of wormlike

micelles are of particular interest because the UV light are easily accessible, clean and pollution-free, non-interference and controllability11, 18. These viscoelastic fluids with light-tunable rheology can find widely applications in microscale or nanoscale devices such as micro flow sensors19, drag-reducing fluids20, rheology control21, drug release 22 and so on. Since surfactants have well self-assembly abilities for its amphipathy, it is expected to be an efficient approach to construct photo-responsive wormlike micelles through the non-covalent interaction between surfactants and photosensitive molecules23-24. Up to now, a number of studies on the photo-responsive wormlike micelles composed of surfactants and photosensitive molecules had been reported 9, 25-28. Previous research on the surfactants and photosensitive molecules binary systems mainly focused on conventional surfactants. Compared with single-chain surfactants, Gemini surfactants consist of two amphiphilic chains connected by a spacer group show more excellent self-assembly capability in aqueous solutions 29-33. For example, the Gemini surfactants can self-assemble into wormlike micelles by themselves at a sufficiently high concentration34-35. Up to now, a number of photo-responsive self-assemblies constructed by Gemini surfactants and photosensitive molecules had been reported36-39. It is expected that the binary systems composed of Gemini surfactants and photoresponsive molecules would observe stronger self-assembly capability and need a smaller amount of surfactants, thus saving the cost, reducing the environmental pollution and expanding their application due to their synergistic reaction.

ACS Paragon Plus Environment

Langmuir 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

It is worth noting that subtle variation in the architecture and position of the substituent group presented in the phenyl ring of aromatics molecules evidently affect the interaction between surfactants and aromatic salts40-41. Up to now, many comparative studies on self-assembly aggregates constructed by aromatic salts and surfactants had been reported

40-43.

It is clear from the literature that the influence of

aromatic additives on a micellar surfactant solution can be rather complex. Small variation in the structure and concentration of counter ions can alter the length and flexibility of micelles drastically and affect the macroscopic rheological properties as well. However, as for photo-responsive fluids, most of the researchers only focused on a single photo-responsive molecule44-49, while few systematic comparative studies on photosensitive molecules had been carried out36-37. For example, Dong’s group discussed the effect of substituent position on phase behavior of trans-coumaric acid isomers (trans-OCA, trans-MCA, trans-PCA) and alkyldimethylamine oxide (CmDMAO) mixed systems. It is noticed that trans-OCA and trans-MCA can be inserted into C12DMAO micelle more efficiently than that of trans-PCA. However, the precise role of the structure of the photo-responsive molecules in dictating the surfactants’ arrangement in micellar layer is not well understood, there are still many problems to break through. The lack of systematic research on Gemini surfactants/ aromatic salts mixed systems significantly limits their applications in different industrial areas. Therefore,the effects of the structure of photosensitive aromatic salts on the construction and performance of viscoelastic wormlike micelles in Gemini surfactant solutions are worth studying. The objective of this work is to research the impact of the structure of ortho substitutes of cinnamate derivatives on the construction of wormlike micelles containing Gemini surfactants. It could be further identify the differences of photo-responsive self-assembly behavior induced by the differences in cis-trans isomerism of photo-responsive molecules. In this work, we select three aromatic salts with cis-trans isomerism, including trans-o-methoxy cinnamate (trans-OMCA), trans-o-hydroxy cinnamate (trans-OHCA), trans-cinnamate (trans-CA). It is expected that these molecules with cis-trans isomerism act as regulators in intelligent response system to construct smart-stimuli responsive

ACS Paragon Plus Environment

Page 4 of 31

Page 5 of 31 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

Langmuir

wormlike micelles. Then we compare the differences of effect of three kinds of photoresponsive molecules with different ortho substituents in Gemini surfactant aqueous solutions. All of them are effective to construct photo-responsive fluids with Gemini surfactant 12-3-12·2Br-. Their photo-responsive behavior and mechanism were researched using the rheological method, cryo-TEM, 1HNMR and 2D NOESY spectrum. Investigation on the influence of ortho substituents of aromatic salts on wormlike micelles constructed by the 12-3-12·2Br-/cinnamate derivatives not only could illustrate the effect of the difference in structures of ortho substituents on the construction of wormlike micelles. It also provides detailed information about the mechanism of switch on/off molecules participating in the construction of wormlike micelles. This research will help us to optimize the intelligent micelles constructed by the binary systems, which is significantly important in designing smart responsive viscoelastic systems and extending their applications in related fields.

EXPERIMENTAL SECTION Materials. Propanediyl-α,ω-bis (N-dodecyl-N, N-dimethylammonium bromide;12-3-12·2Br-) was synthesized according to our previous work50. Trans-o-methoxy cinnamic acid (99%), trans-o-hydroxycinnamic acid (99%), trans-cinnamic acid (99%) were purchased from TCI Chemical Industry Development Co., Ltd and used without further purification. The trans-o-methoxy cinnamate (trans-OMCA), trans-o-hydroxyl cinnamate (trans-OHCA) and trans-cinnamate (trans-CA) used in this paper were prepared by mixing trans-cinnamic acid derivatives with equal molar sodium hydroxide (NaOH) in solution, their FT-IR results are shown in Figure S6 in supporting information. The chemical structures of 12-3-12·2Br-, trans-OMCA, trans-OHCA as well as trans-CA are shown in scheme1. Ultrapure water from Millipore system was used to prepare all samples.

ACS Paragon Plus Environment

Langmuir 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

12-3-12·2Br-

Scheme 1. Chemical structures of Gemini surfactant12-3-12·2Br- and trans-cinnamate derivatives (transOMCA, trans-OHCA and trans-CA)

UV-Light Irradiation. Ultra high-pressure short arc mercury lamp (CHF-XM35-500W) with 365 nm optical filter was used to illuminate samples. The prepared samples were put in a 10ml transparent sample bottle and kept at 25℃ in a jacketed beaker connected with low temperature thermostat bath. Rheological Measurements. Steady dynamic rheological measurements were performed using a cone plate system (CP50-1) with a radius of 50 mm and a taper angle of 1° using physical MCR 302 rheometer (Austrian Antonio Co., Ltd). Before measurement, samples were kept at 298.15K for at least 48h to ensure equilibration. Dynamic stress sweep measurements were used to identify the linear viscoelastic region of the samples. The rheology properties of samples were measured with a fixed stress value in the linear viscoelastic region. The sample was added to peltier plate and equilibrated at 298.15K about 5 min before any measurements. Microstructure measurement. TEM: The microstructures of aggregates were obtained by a transmission electron microscope (Jeol JEM-1400, Japan). Firstly, an appropriate solution was dropped on the side of the copper mesh containing the carbon support film and waited about several minutes for the samples adsorbing on the copper mesh. The excess solution was sucked

ACS Paragon Plus Environment

Page 6 of 31

Page 7 of 31 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

Langmuir

with filter paper. The samples were stained with phosphotungstate and then dried the excess solution with filter paper for observation. Cryo-TEM: Samples was dropped on onto a TEM carbon grid using a micropipette. The TEM carbon grid was blotted with filter papers in order to form a thin film on the mesh hole. And samples were plunged into a reservoir of liquid enhance at -165 °C after 4 s. Then the grid was transferred to a cryo-genic sample holder (Gatan 626) and this holder was put into a JEOL JEM-1400 TEM (120 kV) instrument at approximately -174oC. The microstructure results were obtained. NMR Measurements. 1HNMR

spectra was recorded using a Bruker AVANCE Ⅲ 400MHz FT-NMR

spectrometer at room temperature of 298.15K. Deuterium oxide (99.9%) was used to prepare the stock solutions of the cationic Gemini surfactants with and without transcinnamate derivatives. Two-dimensional (2D) nuclear overhauser effect spectrometry (NOESY) NMR measurements for the equal proportion of 12-3-12·2Br-/ trans-OMCA, 12-3-12·2Br/trans-OHCA and 12-3-12·2Br-/trans-CA mixed solutions were performed on the Bruker AVANCE-500 MHz NMR spectrometer at 298.15K.

RESULTS AND DISCUSSION Rheological behavior of 12-3-12·2Br- / cinnamate derivatives mixed systems.

ACS Paragon Plus Environment

Langmuir 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

Figure 1. Steady shear rheological behavior for mixed systems containing 40mM 12-3-12·2Br- and transCA (0-36mM, a), trans-OHCA (0-24mM, b); Variation of zero-shear viscosity (0) of samples with concentration of cinnamate derivatives ( trans-CA, trans-OMCA, trans-OHCA) at 298.15K (c).

Rheology measurement is a significant method to probe the viscoelastic behavior of fluids. Figure 1 depicts the variation of the rheological behavior for mixed systems containing 40mM 12-3-12·2Br- and a series of different concentrations of cinnamate derivatives. As the similar results of the steady state shear curves for 12-3-12·2Br- and trans-OMCA mixed system have been mentioned in our previous published articles, we mainly show the steady-state shear curves for 12-3-12·2Br-/trans-OHCA and 12-312·2Br-/trans-CA mixed systems51. In addition, the 0 comparison of the three cinnamate derivatives are shown in Figure 1c. As can be seen, the pure surfactant solution appears as a Newtonian fluid52. Then the viscosity of the samples increases with the addition of the cinnamate derivatives. Finally, it is apparent that all plots for samples display a plateau at lower shear rate and appear shear-thinning phenomenon at higher shear rate when a certain amount of cinnamate derivatives (≥12mM) were introduced into 12-3-12·2Br- solutions. The results indicate that the cinnamate derivatives can promote the growth of surfactant micelles along one-dimensional direction and the micelles can entangle with each other to form a network structure, while entangled micelles are arranged in the shear direction at higher shear rate. This is a typical character of wormlike micelles53. This phenomenon is attribute to the intermolecular noncovalent bond interaction, leading to the change of critical packing parameters (CPP) of surfactants41, 54-56. Cross and Carreau models were used to obtain the 057. The variation of 0 for 40mM 12-3-12·2Br- with concentration of the

ACS Paragon Plus Environment

Page 8 of 31

Page 9 of 31 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

Langmuir

cinnamate derivatives is shown in Figure 1d. It is clear from the results that similar trends in 0 are observed in 12-3-12·2Br-/cinnamate derivatives mixed systems. Under the condition of our study, the maximum value of 0 of the system appears at 40mM: 24mM. And the maximum values of 0 for 12-3-12·2Br-/ trans-OHCA, 12-3-12·2Br-/ trans-CA and 12-3-12·2Br-/ trans-OMCA are 26.72Pa·s, 2.89Pa·s and 2.42Pa·s respectively. It can be concluded that the 0 of 12-3-12·2Br-/ trans-OHCA is about 10 times higher than that of 12-3-12·2Br-/ trans-CA and 12-3-12·2Br-/ trans-OMCA systems for the same proportion. Further increase the concentration of cinnamate derivatives, the 12-3-12·2Br-/trans-OHCA appears phase separation phenomenon. For the 12-3-12·2Br-/trans-CA and 12-3-12·2Br-/trans-OMCA systems, the solutions still maintain clear and transparent except the decrease of 0, which should be attributed to the micellar branching58-59.

Figure 2. Dynamic viscoelastic behavior (a) and Cole-Cole plots (b) for 40mM12-3-12·2Br-/24mM trans-OHCA, 40mM 12-3-12·2Br-/24mM trans-CA and 40mM 12-3-12·2Br-/24mM trans-OMCA mixed system.

In order to further compare their rheological behavior, we systematically studied the dynamic viscoelasticity behavior of 12-3-12·2Br-/trans-OHCA, 12-3-12·2Br-/ trans-CA, and 12-3-12·2Br-/trans-OMCA of the same proportion. Figure 2 shows the plots of the elastic modulus (G') and the viscous modulus (G'') as a function of angular frequency (ω) for different mixed systems. It can be seen that G" is greater than G' (G' < G") for all the systems when the oscillation frequency is lower, and the solution primarily exhibits a viscous character. Then G' and G" intersect

ACS Paragon Plus Environment

Langmuir 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 31

at a specific frequency (G' = G") with the further increase of G' and G". And finally the G' exceed G" (G' > G") when the oscillation frequency is higher. This viscoelastic property of fluids coincides well with the typical Maxwell model

60,

indicating the

aggregates formed in 40mM 12-3-12·2Br-/24mM trans-OHCA, 40mM 12-3-12·2Br/24mM trans-CA and 40mM 12-3-12·2Br-/24mM trans-OMCA are typical wormlike micelles. Data points of G″ versus G′ for different systems are shown in Figure 2b. The Cole-Cole plots can also reveal the formation of wormlike micelles directly. The related parameters can be expressed by the following formulas3: ω2τR2

'

G (ω ) = G 0

1 + ω2τR2

G''(ω) = G0 τR =

ωτR 1 + ω2τR2

1 ω*

(1) (2) (3)

Here, G0, ω, τR are the plateau modulus of G′, oscillation frequency and relaxation time respectively. ω* and G* are angular frequency and modulus respectively when G′ intersects with G″. We can calculate from the Figure 2a that the G0 is 99.1Pa, 24.6Pa and 20.9Pa and τR is 0.316s, 0.124s and 0.117s for 40mM 12-3-12·2Br-/24mM trans-OHCA, 40mM 123-12·2Br-/24mM trans-CA, 40mM 12-3-12·2Br-/24mM trans-OMCA. Wormlike micelles are viscoelastic fluids that are in a dynamic equilibrium and can be continuously destroyed and reconstituted. The relaxation time (τR) of wormlike micelles consists of two parts, one part is the relaxation time (τrep) of the micelle growing along the one-dimensional direction, that is, the time required for the micelles to grow to length L. The other part is the relaxation time of the micelles breaking (τb), that is, the time required for the micelles of length L to break into two parts. For a linear viscoelastic fluid that conforms to the Maxwell model, τbtrans-OMCA in 12-3-12·2Br-/ cinnamate derivatives systems. This means that the the length of wormlike micelles in 12-3-12·2Br-/transOHCA mixed system is larger than the corresponding 12-3-12·2Br-/trans-OMCA and 12-3-12·2Br-/trans-CA mixed systems. It because that the electrostatic interaction and hydrophobic interaction between 12-3-12·2Br- and trans-OHCA is stronger, which leads to the closer arrangement of the interfacial layer of micelle molecules and the denser mesh structure. This is caused by greater ionization and smaller steric hindrances of trans-OHCA molecules. The absolute value of the solvation free energy of OHCA in Table 2 is the largest, which can also demonstrate that the degree of ionization in aqueous solution is the largest. The formation of wormlike micelles is attributed to the alternation of CPP at interface through the synergistic interaction between the 12-312·2Br- and cinnamate derivatives. Aromatic anions, in addition to their capacity to charge-neutralize the head groups, can penetrate deeper into the hydrophobic interior of the micelles as a result of favorable hydrophobic forces, thus increasing the CPP by affecting both V (the hydrophobic chain volume) and a (the head group area) of surfactants63. Photo-responsive behavior of 12-3-12·2Br-/cinnamate derivatives mixed systems. Rheology behaviors.

ACS Paragon Plus Environment

Langmuir 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

Figure 3. Effect of UV irradiation time on steady shear rheological behavior of 40mM 12-3-12·2Br/24mM cinnamate derivatives at 298.15K: trans-OHCA (a), trans-OMCA (b), trans-CA (c).

Base on the wormlike micelles constructed by 40mM12-3-12·2Br-/24mM cinnamate derivatives, we further compare the photo-responsive capability of formed wormlike micelles. Figure 3 depicts the variation of steady shear curve of 40mM12-312·2Br-/24mM trans-OHCA (a), 40mM 12-3-12·2Br-/24mM trans-CA (b), 40mM 123-12·2Br-/24mM trans-OMCA (c) with the prolongation of UV irradiation time. For the system containing trans-OHCA, the shear thinning phenomenon still exists with significantly decrease of 0 under UV irradiation (within 150min), indicating that the morphology of aggregates still maintains a three-dimensional network structure except the wormlike micelles becomes shorter in length. However, for the 12-3-12·2Br-/transOMCA system, the shearing thinning phenomenon tends to disappear after the same UV irradiation time. Combined with the TEM observation results (the high definition version is shown in Figure S3 in supporting information), it indicates that the selfassembled morphology of aggregates constructed by 12-3-12·2Br-/trans-OMCA mixed system changes from wormlike micelles to short rod-like micelles after UV irradiation. For the 12-3-12·2Br-/trans-CA mixed system, the shear rheological curve of the system hardly changes after the same UV irradiation time, indicating that the formed wormlike micelles have almost no light response under the same UV irradiated condition. The variation of 0 with UV irradiation time can also be seen directly from the insertion diagram. The 0 of 12-3-12·2Br-/trans-OHCA mixed system changes from 26.72Pa·s to 2.6Pa·s with the UV irradiation time. Accordingly, the 0 of 12-3-12 · 2Br-/transOMCA mixed system is from 2.42 Pa·s to 0.06 Pa·s. For the 12-3-12·2Br-/trans-CA

ACS Paragon Plus Environment

Page 12 of 31

Page 13 of 31 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

Langmuir

system, the η0 nearly keeps at around 2.89Pa·s. In order to further explain the effect of UV light on the morphology and rheological behavior of these systems. We further research their dynamic viscoelastic behavior.

Figure 4. Effect of UV irradiation time on dynamic viscoelastic behavior of 40mM 12-3-12·2Br-/24mM cinnamate derivatives at 298.15K: trans-OHCA (a), trans-OMCA (b), trans-CA (c).

Figure 4 depicts the effect of UV irradiation time on dynamic viscoelastic behavior for 12-3-12·2Br-/trans-OHCA, 12-3-12·2Br-/trans-CA, 12-3-12·2Br-/trans-OMCA mixed systems with a fixed ratio (40mM12-3-12·2Br-: 24mM cinnamate derivatives) at 298.15K. It can be seen from the Figure 4a that the viscoelastic curves of 12-312·2Br-/ trans-OHCA are typical wormlike micelles before and after UV irradiation, which is consistent with the steady shear rheological results. However, for the 12-312·2Br-/trans-OMCA mixed system, the intersection point of viscoelastic curve gradually disappears with the increase of UV irradiation time, and the sample displays a pure viscous fluid (G''>G') in the entire angular frequency range. It reveals that

ACS Paragon Plus Environment

Langmuir 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 31

aggregate morphology gradually changes from wormlike micelles to rod-like micelles after UV irradiation. The viscoelastic curves of 12-3-12·2Br-/trans-CA have almost no change, which is consistent with the steady shear rheological results. We calculate a series of rheological parameters from dynamic frequency sweep results to quantitatively compare the differences of the 12-3-12·2Br-/trans-OHCA and 12-3-12·2Br-/trans-OMCA mixed systems. As can be seen from the Table 1, both of the τ, G* and G0 for 12-3-12·2Br-/trans-OHCA or 12-3-12·2Br-/trans-OMCA system decrease with the prolongation of UV irradiation time. The smaller the value of τR, the faster the diffusion of the wormlike micelles, and the shorter the micellar length. This means that the length of the wormlike micelles formed in the system becomes shorter and the structure becomes looser with the prolongation of UV irradiation time. However, for the 12-3-12·2Br-/trans-CA mixed system, the degree of isomerization of CA is very low under this condition as concluded form the UV-vis results in Figure S1 in supporting information, the spatial steric hindrance and hydrophobicity of cis- do not change much, so there is no obvious photo isomerization. Table 1. Variation of rheological parameters for 40mM 12-3-12·2Br-/24mM trans-OHCA and 40mM 12-3-12·2Br-/24mM trans-OMCA mixed systems with the prolongation of UV irradiation time.

Time(min) 0 20 40 80 110

τR 0.316 0.232 0.193 0.134 0.099

trans-OHCA G* 48.5 36.5 29.4 25.3 21.7

G0 99.1 80.9 63.9 51.8 37.9

ACS Paragon Plus Environment

trans-OMCA τR G* 0.124 10.2 0.081 8.4 0.038 7.6 0.027 7.3 -

G0 24.6 16.2 15.5 13.8 -

Page 15 of 31 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

Langmuir

Cryo-TEM. Before UV: 12-3-12·2Br-/trans-OHCA

a-1

b-1

After UV: 12-3-12·2Br-/trans-OHCA

a-2

b-2

Before UV: 12-3-12·2Br-/trans-OMCA

a-3

a-4

b-3

After UV: 12-3-12·2Br-/trans-OMCA

b-4

Figure 5. Cryo-TEM results of 12-3-12·2Br-/trans-OHCA and 12-3-12·2Br-/trans-OMCA before and after UV irradiation. The scale bars are 500nm for (a) and 200nm for (b).

To further verify the change of the structures formed in the system, the micellar solution was observed using cryo-TEM. The cryo-TEM results of 12-3-12·2Br-/trans-

ACS Paragon Plus Environment

Langmuir 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

OHCA and 12-3-12·2Br-/trans-OMCA before and after UV irradiation are shown in Figure 5. As can be seen from the cryo-TEM results of 12-3-12·2Br-/ trans-OHCA, before UV irradiation, many elongated and flexible structures are entangled with each other. The diameters of aggregates are hundreds of nanometers in the sample and they are difficult to identify where the aggregates begin and end in the cryo-TEM images. After UV irradiation, most of the elongated and flexible aggregates become shorter and the length of the aggregates can be obtained (head and tail marked by arrowheads). As for 12-3-12·2Br-/ trans-OMCA, many linear aggregates are existed and intertwined with each other in (a-3) and (b-3). The length of micelles is obviously shorter than the 12-3-12·2Br-/trans-OHCA of the same proportion. After UV irradiation, most of the the longer flexible aggregates become shorter rigid rod-like aggregates, and some spherical micelle aggregates are presented at the same time. The results indicate that the morphology of aggregates in 12-3-12·2Br-/trans-OHCA system still maintain a threedimensional network structure except the wormlike micelles becomes shorter in length. However, the self-assembled morphology of aggregates constructed by 12-3-12·2Br/trans-OMCA mixed system changes from wormlike micelles to short rod-like micelles or even spherical micelles after UV irradiation. The results are in good agreement with the rheological results. Effect of the ortho substitutes on the interaction of 12-3-12·2Br-/ cinnamate derivatives mixed systems. From the above experimental research, we can conclude that the nature of the substituent group present in the phenyl ring of an aromatic co-solute can have considerable influence on its interactions with a surfactant. The capacity to participate in these interactions directly affects the micellar aggregation morphology of the mixed solution and thus the rheological properties. In order to further understand how the differences of molecular structure of trans-OMCA, trans-OHCA, trans-CA affect the self-assembly between 12-3-12·2Br- and cinnamate derivatives, we studied the interaction between 12-3-12Br- and cinnamate derivatives by 1HNMR and 2D NOESY spectra. 1HNMR

spectra.

ACS Paragon Plus Environment

Page 16 of 31

Page 17 of 31 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

Langmuir

The concentration of 12-3-12·2Br- we choose (40mM) in the above experiments is too high for 1HNMR measurement for it is difficult to operate in nuclear magnetic tube and strongly shield the proton peaks. Therefore, we reduce the concentration of 12-312·2Br- by half. And the concentration of cinnamate derivatives is 0mM, 4mM, 8mM and 12mM respectively. It is worth noting that the diluted system in D2O is still display high viscosity fluid with the increasing of the concentration of cinnamate derivatives before UV irradiation as same as the aqueous system we studied before. The corresponding rheological behavior and photo-responsive behavior of a typical sample in D2O have been provided in Figure S4 in supporting information).

a

b

Figure 6. 1HNMR measurement of 12-3-12·2Br-/trans-OMCA, 12-3-12·2Br-/trans-OHCA and12-312·2Br-/trans-CA(c) before and after UV irradiation: (a) before UV irradiation; (b) after UV irradiation.

It is clear from the above results that the influence of aromatic additives on a micellar solution is worth researching. Therefore, to understand the precise role of the co-solutes, the interaction between surfactants with positive charge and aromatic additives with opposite charge must be taken into account. 1HNMR is one of the most

ACS Paragon Plus Environment

Langmuir 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

powerful techniques for explaining intermolecular interaction between surfactants and additives as for the chemical shift can reflect the micro environmental change of the system64. Firstly, 1H NMR spectra of cinnamate derivatives (12mM) at trans and cis forms are obtained in Figure S5 in supporting information. The signal peaks of cinnamate derivatives at trans forms can be well separated. In the existence of 12-312·Br-, the Figure 6 shows the 1HNMR spectra of 12-3-12·2Br-/ trans-OMCA (a), 123-12·Br-/ trans-OHCA (b), 12-3-12·2Br-/trans-CA (c) before and after UV irradiation. Before UV irradiation, the signal peaks of Hb, Hc, Hf, He protons for 12-3-12·2Br- shift to upfield and all the peaks for 12-3-12·2Br- gradually become broadened and compressed with the increase of the cinnamate derivatives’ concentration. The interaction of cinnamate derivatives with ammonium groups may shield the charge near the headgroups, leading to the upfield shift of the protons near the ammonium headgroups. The phenyl group of cinnamate derivatives may penetrate into the hydrophobic cores of the surfactant aggregates, which makes the protons in the alkyl chains be shield and the protons shifts to upfield42. The broadening and compression of resonance signal peaks for 12-3-12·Br- indicate that the motion of the surfactant molecules is highly restricted, which is usually attributed to the formation of large aggregates65. It means that the electrostatic and hydrophobic interaction existed in 123-12·Br-/cinnamate derivatives systems make the orientation and place of the additives within or at the micellar interface, resulting in the formation of large size aggregates. The shift of signal peaks to upfield of trans-OCMA, trans-OCHA and trans-CA reveals that the cinnamate derivatives exist in a relatively nonpolar environment, indicating that some trans form molecules infiltrate the palisade layer of aggregates and take part in the formation of micelles66. Furthermore, compared with the 12-3-12·2Br-/OMCA and 12-3-12·2Br-/CA system, the signal peaks in 12-3-12·2Br-/OHCA system exhibit more obvious broadening and compression. It means that the 12-3-12·2Br-/OHCA system shows a stronger intermolecular interaction and form more dense wormlike micelles, which are in good agreement with the previous rheological results. After UV irradiation, for 12-3-12·2Br- /OMCA mixed system, the change of proton signal peaks of OMCA molecules indicates that trans-OMCA is isomerized to

ACS Paragon Plus Environment

Page 18 of 31

Page 19 of 31 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

Langmuir

cis-OMCA. The signal peaks of cis-OMCA only change the height with little shifting and broadening. It indicates that the intermolecular interaction between 12-3-12·Br- and cis-OMCA weakens as the OMCA molecules escape from the micellar layer, and then the aggregates are transformed into smaller aggregates67. For 12-3-12·Br-/OHCA system, there are still obvious shift and slight broadening of peaks in the system, indicating that there still are strong interaction and large aggregates in the 12-3-12·Br/OHCA system after UV irradiation. The signal peaks of the 12-3-12·Br- and trans-CA systems in Figure 6 is almost unchanged after the same UV irradiation time. Combined with the UV-vis results in Table S1 and Figure S2 in supporting information, the photoisomerism degree of cinnamate derivatives is OMCA>OHCA>CA under the same UV irradiation time. The trans-CA molecules are difficult to convert into cis-CA, resulting in the formed wormlike micelles almost unchanged before and after UV irradiation. To facilitate a comparison, we selected several groups that exhibited obvious changes in the proton resonances. As shown in Table 2, the order of the displacement difference of the peaks (Δδ) before and after the UV irradiation in 12-312·Br- and cinnamate derivatives mixed systems is obviously OMCA>OHCA>CA. The Δδ of cinnamate derivatives also have the same tendency in Table S2. By studying the intermolecular interaction from a microscopic perspective, the conclusion is in good agreement with the macroscopic rheological phenomena embodied in the system. Table 2. 1HNMR chemical shifts (δ, ppm) of 20mM 12-3-12·2Br- with the existence of 12mM cinnamate derivatives (trans-OMCA, trans-OHCA, trans-CA) before and after UV irradiation at 298.15K.

ppm

Ha

Hc

Hf

He

12-3-12·2Br-

Pure system(δ)

0.8091

1.7041

2.2238

3.1754

12-3-12·2Br-

before UV(Δδ)

0.0415

-0.0486

-0.0319

-0.0513

/trans-OMCA

after UV(Δδ)

0.0081

-0.0226

-0.0041

-0.0786

12-3-12·2Br-

before UV(Δδ)

0.0354

-0.0547

-0.0333

-0.0560

/trans-OHCA

after UV(Δδ)

0.0192

-0.0318

-0.0235

-0.0571

12-3-12·2Br-

before UV(Δδ)

0.0477

-0.0532

-0.0296

-0.0571

/trans-CA

after UV(Δδ)

0.0425

-0.0532

-0.0304

-0.0571

ACS Paragon Plus Environment

Langmuir 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

H-H NOESY spectra. To complement the spectroscopic studies above, we have examined the internal structural change of the micelles through the two-dimensional nuclear overhauser effect spectroscopy (2D NOESY). This technique can provide detailed information on the spatial vicinity ( |△Es|OHCA> |△Es|CA. These results suggest that the ability of OMCA to escape from micellar layer to aqueous solution through UV light is stronger than that of OHCA and CA molecules. This is consistent with the macroscopic phenomena exhibited by the corresponding system. The Vm of the trans form is trans-OMCA> trans-CA> trans-OHCA before the UV irradiation, it shows that the trans-OHCA molecules are arranged more closely in the micelle layer because of the smaller steric resistance. After the UV irradiation, the steric resistance of trans-OMCA and trans-OHCA increases, leading to the micelles to become looser and looser. The photo-responsive mechanism is shown in Figure 8. For

ACS Paragon Plus Environment

Langmuir 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 31

trans form molecules, the aromatic ring penetrates into the non-polar core of the surfactant micelles, while the carboxyl terminus is located in the micellar barrier layer of the aggregate. The electrostatic interaction and hydrophobic interaction between surfactants and additives lead to micelle growth. Besides, due to greater ionization and smaller steric resistance, the trans-OHCA has stronger interaction with 12-3-12·2Brthan that of trans-OMCA. After UV irradiation, for 12-3-12·2Br-/OHCA or 12-312·2Br-/OMCA system, cis form is more hydrophilic than trans form, so cis form is more likely to exist in the aqueous phase and tends to desorb from the aggregates. As a result, the structure of self-assembled aggregates in 12-3-12·2Br-/OHCA and 12-312·2Br-/OMCA systems is different due to the different ability of molecules escaping to polar solution and the different adsorption of isomerized molecules at the interface. In addition, the isomerization ability of photo-responsive molecules in Figure S2 also significantly affected the degree of isomerization of wormlike micelles under the same UV irradiation time. This is especially prominent in the 12-3-12·2Br-/CA system. The differences of noncovalent interaction between 12-3-12·2Br- and cinnamate derivatives lead to the variation of CPP at the interface, resulting to the smart-stimuli self-assembly induced by UV light. Table 3. Solvation free energy and molar volume of trans-OMCA, trans-CA, trans-OHCA before and after UV irradiation.

transOMCA

cisOMCA

transOHCA

cisOHCA

transCA

cisCA

Solvation free energy(KJ/mol)

-37.68

-40.56

-53.97

-56.39

-35.28

-37.16

Molar volume(cm3/mol)

122.23

125.16

107.35

110.60

116.16

104.50

△E(KJ/mol)

-2.88

ACS Paragon Plus Environment

-2.42

-1.88

Page 23 of 31 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

Langmuir

Figure 8. Mechanism of photo-responsive behavior of wormlike micelles formed in 12-3-12·2Br-/transOHCA, 12-3-12·2Br-/trans-OMCA and 12-3-12·2Br-/trans-CA mixed systems.

CONCLUSIONS Three wormlike micelles systems based on the variation of the cinnamate derivatives were developed by mixing 12-3-12·2Br- with trans-OHCA, trans-OMCA, trans-CA through intermolecular noncovalent interaction. The results show that the 123-12·2Br-/trans-OHCA can construct stronger three-dimensional network structure compared with the same proportion of 12-3-12·2Br-/trans-OMCA and 12-3-12·2Br/trans-CA. The high viscosity of 12-3-12·2Br-/trans-OHCA may attribute to two reasons: on one hand, the stronger electronic interaction between trans-OHCA and 123-12·2Br- due to the larger degree of ionization of carboxylate. On the other hand, the hydrophobic interaction between trans-OHCA and 12-3-12·2Br- is stronger because it is embedded more closely in the micelle layer. The strong interaction between 12-312·2Br- and trans-OHCA leads to the closer arrangement of the interfacial layer of micelle molecules and the denser mesh structure. Besides, the 0 of 12-3-12·2Br-/transOHCA decreases from 26.72Pa·s to 2.6Pa·s with the shorter of wormlike micelles after UV irradiation. Corresponding, the 0 of 12-3-12·2Br-/trans-OMCA decreases from 2.42Pa·s to 0.06Pa·s as well as the wormlike micelles are transited into rod-like micelles after the same UV irradiation time. However, the light responsiveness of the12-3-12·2Br-/trans-CA system is not obvious with the 0 nearly maintained at 2.89Pa·s. This is due to the different degree of molecular ionization and isomerization, as well as the change of hydrophilic ability and steric resistance before and after UV

ACS Paragon Plus Environment

Langmuir 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

irradiation, resulting in the variation of aggregate morphology and macroscopic rheology. This study indicates that differences in chemical groups lead to differences in molecular properties, thereby significantly affecting the interactions between molecules. It will help us better understand the effects of chemical groups on micelle systems, thus enriching photo-responsive fluids constructed by surfactants and additives binary system and widening its application in the field of soft matter.

ASSOCIATED CONTENT Supporting Information : Additional experimental and numerical results: UV−Vis Spectroscopy, 1HNMR chemical shifts, FT-IR, TEM.

ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (Project No. 21476072) and the Fundamental Research Funds for the Central Universities.

REFERENCES 1.

Chu, Z.; Dreiss, C. A.; Feng, Y., Smart Wormlike Micelles. Chemical Society

Reviews 2013, 42, 7174-7203. 2.

Wang, A.; Shi, W.; Huang, J.; Yan, Y., Adaptive Soft Molecular Self-Assemblies.

Soft Matter 2016, 12, 337-357. 3.

Rogers, S. A.; Calabrese, M. A.; Wagner, N. J., Rheology of Branched Wormlike

Micelles. Current Opinion in Colloid & Interface Science 2014, 19, 530-535. 4.

Kelleppan, V. T.; Moore, J. E.; McCoy, T. M.; Sokolova, A. V.; Campo, L. d.;

Wilkinson, B. L.; Tabor, R. F., Self-Assembly of Long-Chain Betaine Surfactants: Effect of Tailgroup Structure on Wormlike Micelle Formation. Langmuir 2017, 34, 970-977. 5.

Zhang, Y.; Kong, W.; An, P.; He, S.; Liu, X., CO2/pH-Controllable Viscoelastic

ACS Paragon Plus Environment

Page 24 of 31

Page 25 of 31 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

Langmuir

Nanostructured Fluid Based on Stearic Acid Soap and Bola-Type Quaternary Ammonium Salt. Langmuir 2016, 32, 2311-2320. 6.

Zhang, Y.; Feng, Y.; Wang, Y.; Li, X., CO2-Switchable Viscoelastic Fluids Based

on a Pseudogemini Surfactant. Langmuir 2013, 29, 4187-4192. 7.

Su, X.; Cunningham, M. F.; Jessop, P. G., Switchable Viscosity Triggered by CO2

Using Smart Worm-Like Micelles. Chemical Communications 2013, 49, 2655-2657. 8.

Lin, Y.; Qiao, Y.; Yan, Y.; Huang, J., Thermo-Responsive Viscoelastic Wormlike

Micelle to Elastic Hydrogel Transition in Dual-Component Systems. Soft Matter 2009, 5, 3047-3053. 9.

Oh, H.; Javvaji, V.; Yaraghi, N. A.; Abezgauz, L.; Danino, D.; Raghavan, S. R.,

Light-Induced Transformation of Vesicles to Micelles and Vesicle-Gels to Sols. Soft Matter 2013, 9, 11576-11584. 10. Jiang, J.; Xu, Q.; Wang, G.; Cui, Z., Light and CO2/N2 Dual Stimuli-Responsive Wormlike Micelles Based on a Zwitterionic Surfactant and an Azobenzene Surfactant. Soft Matter 2018, 14, 773-779. 11. Lee, H. Y.; Diehn, K. K.; Sun, K.; Chen, T.; Raghavan, S. R., Reversible Photorheological Fluids Based on Spiropyran-Doped Reverse Micelles. Journal of the American Chemical Society 2011, 133, 8461-8463. 12. Zhang, D.; Lu, X.; Li, Y.; Wang, G.; Chen, Y.; Jiang, J., Dual Stimuli-Responsive Wormlike Micelles Base on Cationic Azobenzene Surfactant and Sodium Azophenol. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2018, 543, 155162. 13. Tsuchiya, K.; Orihara, Y.; Kondo, Y.; Yoshino, N.; Ohkubo, T.; Sakai, H.; Abe, M., Control of Viscoelasticity Using Redox Reaction. Journal of the American Chemical Society 2004, 126, 12282-12283. 14. Wang, L.; Dong, S.; Hao, J., Recent Progress of Magnetic Surfactants: SelfAssembly, Properties and Functions. Current Opinion in Colloid & Interface Science 2018, 35, 81-90. 15. Zhang, Y.; Yang, C.; Guo, S.; Chen, H.; Liu, X., Tandem Triggering of Wormlike Micelles Using Co2 and Redox. Chemical Communications 2016, 52, 12717-12720.

ACS Paragon Plus Environment

Langmuir 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 26 of 31

16. Zhang, Y.; Feng, Y.; Wang, J.; He, S.; Guo, Z.; Chu, Z.; Dreiss, C. A., CO2Switchable Wormlike Micelles. Chemical Communications 2013, 49, 4902-4904. 17. Danov, K. D.; Kralchevsky, P. A.; Stoyanov, S. D.; Cook, J. L.; Stott, I. P.; Pelan, E. G., Growth of Wormlike Micelles in Nonionic Surfactant Solutions: Quantitative Theory Vs. Experiment. Advances in colloid and interface science 2018, 256, 1-22. 18. Bansal, A.; Zhang, Y., Photocontrolled Nanoparticle Delivery Systems for Biomedical Applications. Accounts of chemical research 2014, 47, 3052-3060. 19. Sershen S R; Mensing G A; Ng M, Independent Optical Control of Microfluidic Valves Formed from Optomechanically Responsive Nanocomposite Hydrogels. Advanced Materials 2005, 17, 1366-1368. 20. Shi, H. f.; Wang, Y.; Fang, B.; Yeshayahu, Light-Responsive Threadlike Micelles as Drag Reducing Fluids with Enhanced Heat-Transfer Capabilities. Langmuir 2011, 27, 5806-5813. 21. Sullivan, P. F., Panga, M. K., Lafitte, V. Applications of Wormlike Micelles in the Oilfield Industry. Wormlike Micelles 2017, 6, 330-352. 22. Lone, M. S.; Bhat, P. A.; Shah, R. A.; Chat, O. A.; Dar, A. A., A Green pHSwitchable Amino Acid Based Smart Wormlike Micellar System for Efficient and Controlled Drug Delivery. ChemistrySelect 2017, 2, 1144-1148. 23. Tu, Y.; Gao, M.; Teng, H.; Shang, Y.; Fang, B.; Liu, H., A Gemini SurfactantContaining System with Abundant Self-Assembly Morphology and Rheological Behaviors Tunable by Photoinduction. RSC Advances 2018, 8, 16004-16012. 24. Dai, C.; Yang, Q.; Gao, M.; Zhao, M.; Jiang, J., The Mechanism Difference between CO2 and pH Stimuli for a Dual Responsive Wormlike Micellar System. Physical chemistry chemical physics 2018, 20, 19900-19905. 25. Peng, S.; Guo, Q.; Hughes, T. C.; Hartley, P. G., Reversible Photorheological Lyotropic Liquid Crystals. Langmuir 2013, 30, 866-872. 26. Martin, N.; Sharma, K. P.; Harniman, R. L.; Richardson, R. M.; Hutchings, R. J.; Alibhai, D.; Li, M.; Mann, S., Light-Induced Dynamic Shaping and Self-Division of Multipodal

Polyelectrolyte-Surfactant

Microarchitectures

Photomechanics. Scientific reports 2017, 7, 41327.

ACS Paragon Plus Environment

Via

Azobenzene

Page 27 of 31 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

Langmuir

27. Geng, S.; Wang, Y.; Wang, L.; Kouyama, T.; Gotoh, T.; Wada, S.; Wang, J.Y., A Light-Responsive Self-Assembly Formed by a Cationic Azobenzene Derivative and Sds as a Drug Delivery System. Scientific reports 2017, 7, 39202. 28. Basílio, N.; García-Río, L., Photoswitchable Vesicles. Current Opinion in Colloid & Interface Science 2017, 32, 29-38. 29. Deng, S.; Zhao, J., Self-Assembly of Cationic Gemini Surfactants, Alkanediyl-Bis(Dimethyldodecyl-Ammonium Bromide), in Cyclohexane: Effects of Spacer Length on Their Association into Reverse Lyotropic Liquid Crystalline or Reverse Vesicles. Soft Matter 2018, 14, 734-741. 30. Liu, S.; Wang, X.; Chen, L.; Hou, L.; Zhou, T., Aggregation Morphologies of a Series of Heterogemini Surfactants with a Hydroxyl Head Group in Aqueous Solution. Soft Matter 2014, 10, 9177-9186. 31. Lu, H.; Zheng, C.; Xue, M.; Huang, Z., pH-Regulated Surface Properties and pHReversible Micelle Transition of a Zwitterionic Gemini Surfactant in Aqueous Solution. Physical chemistry chemical physics 2016, 18, 32192-32197. 32. Han, C.; Guo, Y.; Chen, X.; Yao, M.; Zhang, Y.; Zhang, Q.; Wei, X., Phase Behaviour and Temperature-Responsive Properties of a Gemini Surfactant/Brij30/Water System. Soft Matter 2017, 13, 1171-1181. 33. Sharma, R.; Kamal, A.; Abdinejad, M.; Mahajan, R. K.; Kraatz, H.B., Advances in the Synthesis, Molecular Architectures and Potential Applications of Gemini Surfactants. Advances in Colloid and Interface Science 2017, 248, 35-68. 34. Zana, R.; Talmon, Y., Dependence of Aggregate Morphology on Structure of Dimeric Surfactants. Nature 1993, 362, 228-230. 35. Borse, M. S.; Devi, S., Importance of Head Group Polarity in Controlling Aggregation Properties of Cationic Gemini Surfactants. Advances in Colloid and Interface Science 2006, 123, 387-399. 36. Ito, T. H.; Miranda, P. C.; Morgon, N. H.; Heerdt, G.; Dreiss, C. A.; Sabadini, E., Molecular Variations in Aromatic Cosolutes: Critical Role in the Rheology of Cationic Wormlike Micelles. Langmuir 2014, 30, 11535-11542. 37. Jia, K.; Hu, J.; Dong, J.; Li, X., Light-Responsive Multillamellar Vesicles in

ACS Paragon Plus Environment

Langmuir 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

Coumaric Acid/Alkyldimethylamine Oxide Binary Systems: Effects of Surfactant and Hydrotrope Structures. Journal of colloid and interface science 2016, 477, 156-165. 38. Liu Q; Jin H; Fang B; Xu, Z., Heological Property and Rheokinetics of Novel Photogelling Fluids. Journal of Dispersion Science and Technology 2018, 39, 15651571. 39. Yu, L.; Fang, B.; Jin, H.; Chen, J.; Li, K.; Tian, M.; Yang, M.; Jin, L., Rheological Properties of a Novel Photosensitive Micelle Composed of Cationic Gemini Surfactant and 4-Phenylazo Benzoic Acid. Journal of Dispersion Science and Technology 2015, 36, 1770-1776. 40. Kang, W.; Zhao, Y.; Wang, P.; Li, Z.; Hou, X.; Huang, Z.; Yang, H., Rheological Behavior and Mechanism of pH-Responsive Wormlike Micelle Variations Induced by Isomers of Phthalic Acid. Soft Matter 2018, 14, 4445-4452. 41. Chen, X. X.; Liu, Q.; Guo, Y.; Yan, H.; Li, J.; Lv, D. M.; Zhang, J. H.; Liu, M.; Yao, C. H.; Wei, X. L., Effects of Additives on the Viscoelastic Responses of Cationic Gemini Surfactant Solutions. Soft Matter 2018, 14, 6696-6707. 42. Yu, D.; Huang, X.; Deng, M.; Lin, Y.; Jiang, L.; Huang, J.; Wang, Y., Effects of Inorganic and Organic Salts on Aggregation Behavior of Cationic Gemini Surfactants. Journal of Physical Chemistry B 2010, 114, 14955–14964. 43. Kang, W.; Wang, P.; Fan, H.; Yang, H.; Dai, C.; Yin, X.; Zhao, Y.; Guo, S., A pHResponsive Wormlike Micellar System of a Noncovalent Interaction-Based Surfactant with a Tunable Molecular Structure. Soft Matter 2017, 13, 1182-1189. 44. Zhao, M.; Gao, M.; Dai, C.; Du, M.; Liu, Y.; Zou, C.; Tao, J.; Wang, X.; Wang, T., Investigation on a Novel Photo-Responsive System Formed by N-Methyl-NCetylpyrrolidinium Bromide and Ortho-Methoxycinnamic. Journal of Molecular Liquids 2016, 223, 329-334. 45. Du, M.; Dai, C.; Chen, A.; Wu, X.; Li, Y.; Liu, Y.; Li, W.; Zhao, M., Investigation on the Aggregation Behavior of Photo-Responsive System Composed of 1-Hexadecyl3-Methylimidazolium Bromide and 2-Methoxycinnamic Acid. RSC Advances 2015, 5, 68369-68377. 46. Zhao, M.; Gao, M.; Dai, C.; Zou, C.; Yang, Z.; Wu, X.; Liu, Y.; Wu, Y.; Fang, S.;

ACS Paragon Plus Environment

Page 28 of 31

Page 29 of 31 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

Langmuir

Lv, W., Investigation of Novel Triple-Responsive Wormlike Micelles. Langmuir 2017, 33, 4319-4327. 47. Ketner, A. M.; Kumar, R.; Davies, T. S.; Elder, P. W.; Raghavan, S. R., A Simple Class of Photorheological Fluids: Surfactant Solutions with Viscosity Tunable by Light. Journal of the American Chemical Society 2007, 129, 1553-1559. 48. Wang, D.; Dong, R.; Long, P.; Hao, J., Photo-Induced Phase Transition from Multilamellar Vesicles to Wormlike Micelles. Soft Matter 2011, 7, 10713-10719. 49. Li, J.; Zhao, M.; Zhou, H.; Gao, H.; Zheng, L., Photo-Induced Transformation of Wormlike Micelles to Spherical Micelles in Aqueous Solution. Soft Matter 2012, 8, 7858. 50. Chen, L.; Shang, Y.; Liu, H.; Hu, Y., Effect of the Spacer Group of Cationic Gemini Surfactant on Microemulsion Phase Behavior. Journal of colloid and interface science 2006, 301, 644-650. 51. Pletneva, V. A.; Molchanov, V. S.; Philippova, O. E., Viscoelasticity of Smart Fluids Based on Wormlike Surfactant Micelles and Oppositely Charged Magnetic Particles. Langmuir 2015, 31, 110-119. 52. Dreiss, C. A., Wormlike Micelles: Where Do We Stand? Recent Developments, Linear Rheology and Scattering Techniques. Soft Matter 2007, 3, 956. 53. Macías, E. R.; Bautista, F.; Pérez-López, J. H.; Schulz, P. C.; Gradzielski, M.; Manero, O.; Puig, J. E.; Escalante, J. I., Effect of Ionic Strength on Rheological Behavior of Polymer-Like Cetyltrimethylammonium Tosylate Micellar Solutions. Soft Matter 2011, 7, 2094-2102. 54. Siddiqui, U. S.; Khan, F.; Khan, I. A.; Dar, A. A.; Kabir ud, D., Role of Added Counterions in the Micellar Growth of Bisquaternary Ammonium Halide Surfactant (14-S-14): 1h Nmr and Viscometric Studies. Journal of colloid and interface science 2011, 355, 131-139. 55. Shrestha, R. G.; Nomura, K.; Yamamoto, M.; Yamawaki, Y.; Tamura, Y.; Sakai, K.; Sakamoto, K.; Sakai, H.; Abe, M., Peptide-Based Gemini Amphiphiles: Phase Behavior and Rheology of Wormlike Micelles. Langmuir 2012, 28, 15472-15481. 56. Wang, J.; Feng, Y.; Agrawal, N. R.; Raghavan, S. R., Wormlike Micelles Versus

ACS Paragon Plus Environment

Langmuir 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

Water-Soluble Polymers as Rheology-Modifiers: Similarities and Differences. Physical chemistry chemical physics 2017, 19, 24458-24466. 57. Raghavan, S. R.; Kaler, E. W., Highly Viscoelastic Wormlike Micellar Solutions Formed by Cationic Surfactants with Long Unsaturated Tails. Langmuir 2001, 17, 300306. 58. Ziserman, L.; Abezgauz, L.; Ramon, O.; Raghavan, S. R.; Danino, D., Origins of the Viscosity Peak in Wormlike Micellar Solutions. 1. Mixed Catanionic Surfactants. A Cryo-Transmission Electron Microscopy Study. Langmuir 2009, 25, 10483-10489. 59. Zhang, Y.; Feng, Y.; Wang, J.; He, S.; Guo, Z.; Chu, Z.; Dreiss, C. A., CO2Switchable Wormlike Micelles. Chemical Communications 2013, 49, 4902-4904. 60. Davies, T. S.; Ketner, A. M.; Raghavan, S. R., Self-Assembly of Surfactant Vesicles That Transform into Viscoelastic Wormlike Micelles Upon Heating. Journal of the American Chemical Society 2006, 128, 6669-6675. 61. Acharya D P, Kunieda H. Wormlike micelles in mixed surfactant solutions. Advances in colloid and interface science 2006, 123, 401-413. 62. Lu, Y.; Zhou, T.; Fan, Q.; Dong, J.; Li, X., Light-Responsive Viscoelastic Fluids Based on Anionic Wormlike Micelles. Journal of colloid and interface science 2013, 412, 107-111. 63. Alonso, B.; Harris, R. K.; Kenwright, A. M., Micellar Solubilization: Structural and Conformational Changes Investigated by 1h and 13c Liquid-State Nmr. Journal of colloid and interface science 2002, 251, 366-375. 64. Yang, Y.; Dong, J.; Li, X., Novel Viscoelastic Systems from Azobenzene Dye and Cationic Surfactant Binary Mixtures: Effect of Surfactant Chain Length. Journal of Dispersion Science and Technology 2013, 34, 47-54. 65. Chen, A. J.; Hsu, I. J.; Wu, W. Y.; Su, Y. T.; Tsai, F. Y.; Mou, C. Y., A Fluorescent Organic Nanotube Assembled from Novel P-Phenylene Ethynylene-Based Dicationic Amphiphiles. Langmuir 2013, 29, 2580-2587. 66. Jiang, Z.; Jia, K.; Liu, X.; Dong, J.; Li, X., Multiple Responsive Fluids Based on Vesicle to Wormlike Micelle Transitions by Single-Tailed Pyrrolidone Surfactants. Langmuir 2015, 31, 11760-11768.

ACS Paragon Plus Environment

Page 30 of 31

Page 31 of 31 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

Langmuir

67. M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery, Jr., J.E. Peralta, F. Ogliaro, M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar, J. Tomasi, M. Cossi, N. Rega, N.J. Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma, V.G. Zakrzewski, G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, Ö. Farkas, J.B. Foresman, J.V. Ortiz, J. Cioslowski, D.J. Fox, 2009. Gaussian, Inc., Wallingford CT, Gaussian 09, Revision D.01.

Table of contents:

Photo-responsive behavior of wormlike micelles formed in 12-3-12·2Br-/transOHCA, 12-3-12·2Br-/trans-OMCA and 12-3-12·2Br-/trans-CA mixed systems.

ACS Paragon Plus Environment