NMR, Raman and DFT Study of Lyotropic Chromonic Liquid Crystals

5 hours ago - In this case, the system evolves via long lasting self-assembling from the supercooled state to the equilibrium. This process affects th...
1 downloads 10 Views 1MB Size
Subscriber access provided by UNIV OF SCIENCES PHILADELPHIA

Article

NMR, Raman and DFT Study of Lyotropic Chromonic Liquid Crystals of Biomedical Interest: Tautomeric Equilibrium and Slow Self-Assembling in Sunset Yellow Aqueous Solutions Kristina Kristinaityte, Arunas Marsalka, Laurynas Dagys, Kestutis Aidas, Iryna Doroshenko, Yevhenii Vaskivskyi, Yelyzaveta Chernolevska, Valeriy Pogorelov, Nomeda Rima Valeviciene, and Vytautas Balevicius J. Phys. Chem. B, Just Accepted Manuscript • DOI: 10.1021/acs.jpcb.8b00350 • Publication Date (Web): 28 Feb 2018 Downloaded from http://pubs.acs.org on February 28, 2018

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.

The Journal of Physical Chemistry B 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 32 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

The Journal of Physical Chemistry

NMR, Raman and DFT Study of Lyotropic Chromonic Liquid Crystals of Biomedical Interest: Tautomeric Equilibrium and Slow Self-Assembling in Sunset Yellow Aqueous Solutions

Kristina Kristinaitytė1, Arūnas Maršalka1, Laurynas Dagys1, Kęstutis Aidas1, Iryna Doroshenko2, Yevhenii Vaskivskyi2, Yelyzaveta Chernolevska2, Valeriy Pogorelov2, Nomeda Rima Valevičienė3, Vytautas Balevicius1*

1

2

Faculty of Physics, Vilnius University, Sauletekio al. 3, LT-10257 Vilnius, Lithuania

Faculty of Physics, Taras Shevchenko National University of Kyiv, Glushkova av. 4, 03022, Kyiv, Ukraine

3

Faculty of Medicine, Vilnius University, Santariskiu 2, Vilnius LT-08661 Vilnius, Lithuania

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry 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 32

ABSTRACT: Temperature and composition effects in Sunset Yellow FCF (SSY) aqueous solutions were studied by the 1H,

15

N NMR as well as Raman spectroscopy passing through all

phase transitions between isotropic phase (I) and chromonic phases - nematic (N) and columnar (M). It was shown that the tautomeric equilibrium in SSY is strongly shifted towards the hydrazone form. The corresponding equilibrium constant pKT = 2.5 was deduced using the DFT SMD model. The dominance of the hydrazone form was confirmed experimentally using the long-range 1H–15N correlation, widely known as HMBC. The peak found in 1H NMR spectra at ca 14.5 ppm can be attributed to the proton in the intramolecular N–H…O bond. The existence of this signal shows that (i) the growing of SSY aggregates is accompanied by the segregation of water in the intercolumnar areas with no access for exchange with the N–H protons in the internal layers of the columnar stacks and that (ii) the life time of those aggregates is ≥ 10–8 s or even longer. The temperature dependences of H2O chemical shift and Raman O–H stretching band shape show that water confined in the intercolumnar areas behaves as in the neat substance. When the sample is heated and the transition from M phase to N phase occurs, the molecular motion of water is seen to change in a manner similar to that when water is melting. The equilibration time for N+M → M is very long because of slow supramolecular restructuring, i.e. the growing of columnar stacks and building of hexagonal arrays. If the sample is cooled down to the temperature below N → M transition relatively fast, the structural changes are behind, and the system falls into supercooled state. In this case, the system evolves via long lasting selfassembling from the supercooled state to the equilibrium. This process affects the shape of the 1

H NMR signal and is easy to monitor.

ACS Paragon Plus Environment

2

Page 3 of 32 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

The Journal of Physical Chemistry

Introduction Lyotropic liquid crystals represent an intersection of several challenging fields of research – supramolecular self-assembling that very often takes place at mesoscopic scale and broad variety of processes in ordered complex fluids.1,2 The spontaneous aggregation due to a combination of wide set of molecular interactions, ranging from weak Van der Waals forces to strong hydrogen bonding, media and hydrophobic effects are very important phenomena in the field of both fundamental research and technological applications such as development of new materials or new ways to treat diseases.3,4 The importance of the balance between different types of intermolecular forces for designing processes at molecular scale was reported very recently,5 showing that the transfer of assembles of the lipid rafts into the surfactant-based lyotropic liquid crystals can be realized in reality. Lyotropic chromonic liquid crystals (LCLCs) are a novel and important yet relatively poorly studied class of soft matter. These materials have attracted considerable attention first of all in the areas of optics and innovative biomaterials.6–8 The combination of self-assembling, ease of alignment, sensitivity to changing conditions and additives, coupled with their optical properties, makes these systems unique and valuable creating micropatterned materials and constructing various sophisticated components and devices, including polarizers, optical compensators, lightharvesting equipment. The fact that LCLCs are water-based, suggests and promises a future role in biosensing for medical diagnostics. And indeed, there was noted a growing interest in using the liquid crystals in biological sensors as the medium that amplifies the reactions in molecularand meso(submicron)-scales rescaling those to the macro-scale, which is accessible for optical detection.9 However, the thermotropic liquid crystals are often toxic.10 The antigen-bearing

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry 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 32

agents (particles or microbes) and the corresponding antibodies are free to move in the cell filled with the water-based but non-surfactant and thus non-toxic lyotropic chromonic liquid crystal (LCLC). Small and isolated particles do not disrupt the uniform alignment of LCLC but the formation and growth of immune complexes trigger director distortions detectable by optical means.9 Chromonic phases are usually formed in water from disk-like or plate -like multi-ring aromatic compounds including drugs, dyes and nucleic acids. These phases occur as the result of the faceto-face aggregation of the molecules into columns forming stacks with the ionic solubilizing groups exposed at the aggregate-water interfaces.6–8,11 There are two principal chromonic phases – the N phase, which consists of a nematic array of columns and at higher concentrations - the M phase (also called columnar), in which the columns lie in a hexagonal array. The aggregation takes place at all concentrations (without a critical micelle concentration), with the distribution of aggregate sizes tending toward larger aggregates as the concentration increases.11 The number of molecules in an aggregate in the columnar phase tends to be unconstrained. In the nematic as well as in the isotropic phases the aggregates are of a finite size.4,12-16 As discussed in these papers, the reason for the finite length of the aggregate is of an entropic origin. The aggregate length decreases with decreasing pH when HCl is added to the solution.17 The most well-known LCLC materials are disodium cromoglycate (DSCG), also called cromolyn sodium, and 6-hydroxy-5-[(4-sulfophenyl)azo]-2-naphthalenesulfonic acid, also known as Sunset Yellow FCF (SSY) or Edicol. Disodium cromoglycate has been used since the mid-1970s for the prophylactic treatment of allergic diseases, such as asthma and allergic rhinitis.18,19 Recently, the phases and structure of mixtures composed of these two representative

ACS Paragon Plus Environment

4

Page 5 of 32 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

The Journal of Physical Chemistry

chromonic liquid crystal materials in water have been studied using polarized light microscopy, microspectroscopy, centrifugation and X-ray diffraction.20 A variety of combinations of isotropic (I), nematic (N) and columnar (M) phases are observed depending on their concentrations. It was noted that their miscibility in the M phase is so low that essentially complete phase separation occurs. This phase behavior of the mixture is due to both separation of the molecules in the formation of the aggregates and separation of aggregates of the same species by their slow diffusion. Our initial visual observations of the changes in the samples as well as some quick spectroscopic experiments have revealed that extremely slow processes can take place in SSY solutions in water. They can last over hours or even over tens of days. The samples of certain composition kept for a long time start to look as if they were dried, i.e. the water was evaporated or phase separation occurred. Therefore, the purpose of the present work was the precise study of temperature and composition effects on the phase behavior, slow self-assembling processes and tautomeric equilibrium in SSY aqueous solutions applying NMR and Raman spectroscopy. In order to make some predictions and to obtain supplementary insights into the experimental observations, the DFT calculations of the tautomeric equilibrium constant and the 1H, 15

13

C and

N magnetic shielding tensors of SSY have been performed.

Experimental Samples. The commercial Sunset yellow FCF (SSY) was used from abcr GmbH, Karlsruhe, Germany (www.abcr.de), of 90 % purity. The water used was freshly bidistilled. For concentration dependence measurements the samples were prepared using micropipette and high-accuracy weighting scales during the preparation to ensure the exact concentration. Then

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry 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 32

the samples were placed in standard 5 mm or 10 mm NMR tubes and in a rectangular quartz cuvette (Helma Analytics) for NMR and Raman experiments, respectively. The samples of 0.6 1 mL volume were tightly closed to avoid any changes of composition. No changes in the sample weights were observed after long lasting experiments at the varying temperature. NMR experiments were carried out on Bruker AVANCE III HD spectrometer operating at resonance frequency of 400 MHz and 40 MHz for 1H and 15N respectively (magnetic field of 9.4 T) using BBO probe head. The concentrations were chosen according to the phase diagram of SSY in order to observe all phase transitions in the temperature range from 25 oC up to 95 oC. The temperature in a probe was controlled with an accuracy of ± 0.2 deg. The signal of DSS in D2O solution in capillary insert was used as the reference. The D2O in the same capillary insert was used for locking.

15

N NMR spectra were referred respect to CH3NO2, also in the capillary

insert. The 90° pulse length was set 25 µs and up to 1024 scans were accumulated with a repetition delay of 2 s in all 1D NMR experiments. Unless stated otherwise, spectral width was 20 ppm with 96152 points as size of FID. The hetero nuclear 1H–15N correlation spectrum was acquired using the multiple bond J-coupling pulse sequence (HMBC experiment) with the additional pulses for the coherence selection. The experiment was acquired with 256 points in the channel F1 and 16000 points at the channel F2 with repetition delay of 2 s. 10 mm tube with a separate capillary insert for referencing and shimming the magnet was used in order to achieve a better signal-to-noise ratio. The number of scans NS = 80 was found to be appropriate to accumulate enough intense signals of nitromethane (reference) as well as SSY ≥ 0.7 mol/kg in water solutions. NMR spectra were processed using Topsin 3.2 software.

ACS Paragon Plus Environment

6

Page 7 of 32 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

The Journal of Physical Chemistry

Raman spectra were recorded in the wavenumber region of 100 – 4000 cm–1 using Fourier transform (FT) Raman spectrometer MultiRAM (Bruker Optik GmbH, Germany) coupled with an avalanche diode detector. A semiconductor laser with a wavelength of 785 nm was used as an excitation source. Laser power was set to 500 mW. A gold-plated 90 degree angle mirror objective (focal length of 33 mm) was used for sample irradiation and collecting of the light scattered by the sample. The diameter of the laser spot at the focal point was 100 µm. To obtain the spectra 256 interferograms were averaged and the resulting interferogram was Fourier transformed by applying Blackman-Harris 3-Term apodization function and zero filling factor of 2. The spectra were recorded using 4 cm–1 spectral resolution. The temperature of the samples were varied using Linkam LNP95 cooling-heating system coupled with the temperature controlled stage FTIR600 (Linkam Scientific, Tadworth, United Kingdom). Some additional processing of NMR as well as Raman spectra was carried out using Microcal Origin 9 package.

Computational details All electronic structure calculations were performed using Gaussian 09 suite of programs.21 To predict isotropic NMR shielding constants, we have first optimized the geometries of isolated species using the B3LYP exchange correlation functional22 along with the Pople type 6-311G* basis set.23,24 The NMR shielding constants were computed using the PBE0 exchange correlation functional25 and 6-311++G(2d,2p) basis set,23,24 an approach, which has proven to provide reliable results for NMR shieldings.26–28 All calculations here were performed for isolated molecules. The pruned (99,590) grid implemented in Gaussian 09 for the evaluation of the two-

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry 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 32

electron integrals was utilized in both the geometry optimization and calculation of the NMR shielding constants. The tautomeric equilibrium constant, pKT, is associated with the Gibbs free energy of the tautomerization reaction in aqueous solution, ∆Gaq, according to

୼ீೌ೜

‫= ்ܭ݌‬ . ோ்୪୬ଵ଴

(1)

Here R is the universal gas constant and T is the temperature. To compute the free energy change we rely on the standard thermodynamic cycle approach where ∆Gaq is evaluated as

Δ‫ܩ‬௔௤ = Δ‫ܩ‬௚௔௦ + ΔΔ‫ܩ‬௦௢௟௩ .

(2)

Here ∆Ggas is the gas-phase free energy of tautomerization reaction and ∆∆Gsolv is the difference of the solvation free energies of both tautomers. The former term is computed using standard rigid rotator harmonic oscillator approach in the standard state of 1 atm. The solvation free energies in the standard state of 1 mol/l were evaluated using the SMD model,29 which belongs to the class of continuum solvation models. The M05-2X exchange correlation functional30 and the 6-31G* basis set31,32 along with the pruned (99,590) grid were utilized in the geometry optimization and harmonic frequency calculations of tautomeric species in both gas phase and aqueous solution. The tautomeric equilibrium constant was evaluated at the temperature of 298.15 K and at the pressure of 1 atm.

ACS Paragon Plus Environment

8

Page 9 of 32 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

The Journal of Physical Chemistry

Table 1. Calculated and Experimental Chemical Shifts of dianionic SSY Nuclei Closely Involved in Tautomeric Equilibrium Process. atom

azo

hydrazone

experimental

H1

15.4a

16.8

14.2 – 14.7

N2

58.4

– 191.5

– 178

N3

122.3

– 8.9



C4

156.9

184.3

177.9 - 178.7b

a

all calculated chemical shifts are given in the δ-scale as the difference between the isotropic part of magnetic shielding tensor (σiso) and that of TMS 31.6 ppm (1H) and 187.9 ppm (13C), respectively, and nitromethane (CH3NO2) for the nitrogen shifts (– 144.7 ppm); b taken from Ref. 33.

Figure 1. The calculated molecular structures of azo- and hydrazine tautomers of dianionic SSY at B3LYP/6-311G* level and the labeling of atoms used in discussion.

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry 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 32

Results and discussion The phase diagram of SSY aqueous solution is known. Isotropic (I), nematic (N) and columnar (M) phases are observed depending on the composition of the mixture and temperature.4,8,33,34 The solutions in the concentration range of 0.7 - 1.4 mol/kg of SSY were prepared according to the phase diagram. This allowed us to pass through all possible phase transitions I ↔ N ↔ M by simply changing the temperature from 25 to 90 oC. The structural and dynamic features of these phases were reviewed in Ref. 6. Nematic (N) phase is characterized by the orientational long-range order of columns only, and no positional order is present. The rotational disorder of columns around the column axes with some rotational disorder of molecules within columns are present in this phase. Columnar (M) phase is an array of parallel columns with space- and time-averaged hexagonal symmetry, however, it exhibits regions of short range structures with orientational and positional long range order of column axes and restricted rotation of columns. NMR and Raman spectroscopy techniques have already proven to be a powerful tools in LCLCs studies.18,19,33–35 In this work too, already during our first experiments we have noticed that NMR technique is sensitive to many processes in the samples of SSY. It was very clearly observed in the 1H NMR spectra that after the initial mixing in the isotropic phase we observe different kinds of phase behavior and slow self-assembling depending on the concentration, age and thermal history of the mixtures (Fig. 2).

ACS Paragon Plus Environment

10

Page 11 of 32 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

The Journal of Physical Chemistry

Figure 2. Temperature evolution of the 1H NMR spectra in 1.2 mol/kg SSY aqueous mixture at phase transitions and slow self-assembling depending on the degree of mixing and thermal history. The detailed evolution of signal shapes and chemical shifts is shown in Figs. 4 - 6. See text for more comments.

Tautomeric equilibrium and aggregation The greater stability of the N–H…O form as compared to that of O–H…N in the series of molecular systems with intramolecular H-bonds was noted and discussed in terms of the greater proton affinity of nitrogen with respect to that of oxygen and the resonance-assisted H-bonding.36 The DFT calculations carried out in the present work also predict that the tautomeric equilibrium in SSY should be strongly shifted towards the hydrazone form. It is characterized by the equilibrium constant pKT = 2.5 as obtained at the DFT SMD level of theory. However, our DFT

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry 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 32

based estimate of the tautomeric equilibrium constant is regarded as a lower bound of the equilibrium. The fact that the hydazone form is dominant for SSY was deduced also in other works using

13

C NMR shifts33 and Raman spectroscopy.34 We decided to revisit this problem,

which is inspired by the fact that the experimentally observed 13C chemical shift of C4 (Fig. 1) of 177 ppm33 falls in the range between presently calculated values of 156.9 ppm and 184.3 ppm for C4 in azo and hydrazone tautomers, respectively, using DFT (Table 1). Moreover, the DFT predicted difference of chemical shifts of N1 nitrogen in those tautomeric forms is very huge – up to 260 ppm (!) and therefore

15

N NMR experiments look to be extremely promising for

detection of possible tautomeric forms. The long-range 1H–15N correlation, widely known as HMBC, was applied for this purpose (Fig. 3). The 2D spectrum is smeared out due to the strong solvent signal at 4.7 ppm (H2O). Nevertheless the peaks at (6.9 ppm; – 178 ppm) and (4.4 ppm; 0 ppm) are clearly seen. They correspond to the 1H and

15

N correlations in SSY and in nitromethane (used for reference),

respectively. The value observed chemical shift δ(15N) = – 178 ppm is in fairly good agreement with the DFT calculated one at – 191.5 ppm for N2–H nitrogen in hydrazone tautomer of SSY (Fig. 1), whereas this peak in azo tautomer should be observed at + 58.4 ppm (see Table 1). This means that

15

N NMR data also confirm the dominance of hydrazone form of SSY at present

conditions (0.7 M of SSY in water at 293 K). Unfortunately, the

15

N NMR signal of N3 that

should be present at – 8.9 ppm or + 122 ppm for hydrazone- and azo tautomers, respectively, was not found experimentally. Most probably, this is due to the too long correlation with protons as it is evident from the molecular structure of SSY.

ACS Paragon Plus Environment

12

Page 13 of 32 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

The Journal of Physical Chemistry

Figure 3. 1H–15N correlation trough long-range couplings in 0.7 mol/kg SSY aqueous solution at 25 oC. The experimental details are given in the text.

The chemical shift of the N–H…O bridge proton (H1, see Fig. 1) is expected to be another very sensitive NMR spectral feature because it can provide the direct information on the tautomeric dynamics and the strength of H-bond in SSY. The chosen DFT level using the PBE0 exchange correlation functional has proven to be one of most successful choices for NMR shielding calculations, particularly for the nuclei involved in strong H-bonds.27,28 A perfect agreement between the calculated and experimentally observed 1H chemical shift was obtained in many cases, e. g. picolinic acid N-oxide (PANO) dissolved in various solvents.28 PANO and SSY can be considered as closely related systems because strong intramolecular H-bonds are formed in both of them. The success of DFT application for the first system makes it promising for SSY as well. The calculations predict that the chemical shifts of H1 proton of SSY are 15.4

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry 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 32

and 16.8 ppm, depending on the tautomeric form (Table 1). Note the shift values covering the range from 14.1 ppm to 18.4 ppm were determined experimentally for series of N–H…O bonds as the N…O distance shortens from 2.51 Å to 2.56 Å.36 A problem of experimental observation of this signal in SSY aqueous solution may arise because of the solvent, which is water. It is usually expected that protons involved in H-bonding should undergo fast exchange with water. Thus, their NMR signals are consequently broadened and not observed. A nice example of such behavior was observed in PANO: the peak of O–H…O proton at 17 - 19 ppm is clearly observed in a series of solutions in traditional solvents and even in some ionic liquids, but not in aqueous solution.28 Such a priori knowledge leads to certain pessimism, so that the researchers even do not struggle much to find this peak. However, in the case of chromonic liquid crystals we meet one peculiar feature of these systems – a strong self-assembling. It is known that chromonics are lyotropic liquid crystals, but in general they do not possess a critical micelle concentration,4,11,17 as it is usually in many cases of amphiphilic molecules forming the lyotropic liquid crystalline phases and gels, e.g. Refs 27, 37 and Refs therein. Hence, unlike many lyotropic liquid crystals, self-assembly in chromonics occurs at very low concentrations.38 The self-assembling leads to the formation of rod-like aggregates, also more complex geometries with 'stacking faults'.8,39 These structures are stabilized by the π-π stacking interactions between the aromatic rings.4,11,38 Thus one can expect that water molecules can interact and thus get into exchange with H1 protons of SSY placed on the aggregate surfaces only. But water does not penetrate into the bulk of the aggregate. It could mean that the N–H…O bridges are confined inside the aggregates and if the life time of those aggregates is ≥ 10–7 - 10–8 s, which is typical NMR time scale or is even longer, the corresponding 1H NMR signal can be expected to appear in the spectrum. And indeed, the peak

ACS Paragon Plus Environment

14

Page 15 of 32 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

The Journal of Physical Chemistry

at 14.2 - 14.7 ppm was found at all compositions studied in the present work and in all phases, except at three lowest temperatures 25, 30 and 35 oC at the cooling cycle (Fig. 4A). After the system has evolved via long lasting self- assembling to the equilibrium this peak was seen even at those temperatures (Figure 4B). First of all, these values of chemical shifts are slightly below DFT predicted ones (Table 1). For one thing, our DFT calculations of NMR shielding have been carried out for isolated molecules, and solvent effects remain unaccounted for. On the other hand, this could indicate that some H1 protons of SSY, most probably, on the aggregate surfaces, exchange with water indeed. The observed slope ∆δ/∆Τ ∼ – 0.022 ppm/deg (Fig. 4C) is significantly higher than could be expected in the case of solely H-bond effect. ∆δ/∆Τ values usually observed for various Hbonded systems are spread from ca – 0.006 ppm/deg (e.g. pyridine N-oxide/acids complexes and PANO) to – 0.012 ppm/deg (the neat water).28,40 The steep increase of chemical shift δ(N–H) with decreasing temperature in M phase (Fig. 4C) can be explained to be due to the growth of SSY aggregates (columnar stacks) accompanied by the segregation of water in the intercolumnar regions and thus drastically reducing the exchange between water and N–H protons. The broadening of H1 signal at decreasing of temperature is caused by several overlapping factors. The main contribution to the signal broadening comes from the slow-down of an overall rotational motion with increasing size of the SSY aggregates. NOESY experiments35 have shown that the SSY molecules are in the intermediate tumbling regime with the rotational correlation time of ∼ 300 ps in the isotropic phase. However, this system reaches at higher concentrations the slow tumbling or spin-diffusion limit, characteristic for large molecules with long rotational correlation times.35 As the peak at 14.2 - 14.7 ppm was not found at lowest temperatures at the

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry 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 32

cooling cycle (Figure 4A) it can be stated that the non-equilibrium fluctuations in the local environment can also contribute to the signal broadening. All other 1H NMR peaks of SSY protons disappeared at the lowest studied temperatures for the same reasons too (Fig. 5), but not the peak of water (Fig. 6). Moreover, the signal shape and the chemical shift δ(H2O) depend not only on concentration and temperature, but, – what is the most important, – on the thermal history (Fig. 2). Thus, it can be effectively used to study the slow self-assembling when the borders between N, N+M and M phases are crossed.

ACS Paragon Plus Environment

16

Page 17 of 32 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

The Journal of Physical Chemistry

Figure 4. Evolution of 1H NMR signal shape at cooling (A) and heating (B) cycles and the dependencies of chemical shift of N–H…O proton on temperature in 1.2 mol/kg SSY aqueous solution (C).

ACS Paragon Plus Environment

17

The Journal of Physical Chemistry 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 32

Figure 5. Temperature evolution of 1H NMR spectra of SSY protons in 1.2 mol/kg SSY aqueous solution at cooling (A)and heating (B) cycles.

ACS Paragon Plus Environment

18

Page 19 of 32 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

The Journal of Physical Chemistry

Figure 6. Evolution of 1H NMR signal shape of H2O protons at cooling (A) and heating (B) cycles and the dependency of H2O chemical shift on temperature (C). The spectra in the cooling cycle were normalized for better visualization to maximal peak height. The points corresponding to the chemical shifts of 'broad' and 'narrow' signal components are marked by blue triangles and red circles, respectively.

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry 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 20 of 32

Slow self-assembling after N+M → M phase transition Going through the experimental details published in literature, it becomes evident that the temperature variation rates traditionally used in the experiments where phase behavior in LCLC systems is studied are in the range of 0.2 - 0.4 deg/min.4,8 In some cases the heating/cooling cycles were carried out using a much higher rate of 5 deg/min, although it used to be reduced to 1 deg/min near the phase transitions.33 The results presented in Figs 4 and 6 show that such rates indeed look to be properly working within I, N and the coexisting I+N phases. The measured dependencies of chemical shifts δ(N–H) and δ(H2O) on temperature are practically identical, as for the cooling as well as for the heating cycles. However, it is obvious that such rates at the cooling and crossing N+M to M phase border are too high. The equilibration time for N+M → M process is very long, most probably because of supramolecular restructuring, i.e. the growing of aggregates (columnar stacks) in M phase. Hence, if the sample is cooled down relatively fast to the temperature below N → M transition, the structural changes are behind, and the system falls to the supercooled state. In that case, the system evolves via slow self-assembling from this state to the equilibrium. This process is very clearly reflected in the 1H NMR spectra via appearance of the wide ‘podos’ around the peak of H2O at 4.1 - 4.8 ppm (Figs 2 and 6). The signal can be decomposed into 'narrow' and 'broad' components. If the sample is kept at the constant temperature in M phase (+25 oC) for a day or tens of days in some cases the complex signal shape collapses to the single contour having a moderate width. Such evolution of the 1H NMR signal shape at slow self-assembling is shown in Fig. 2. Visually it looks like an extremely slow phase separation (see the insert photo in Fig. 2) that can last over a day or tens of days. The samples start to look as if the bulk water was gone (evaporated) and samples were dried. It is

ACS Paragon Plus Environment

20

Page 21 of 32 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

The Journal of Physical Chemistry

possible that the observed phenomenon can be related to a certain stage of the drying of liquid crystal droplets that was very recently discovered dealing also with SSY aqueous solutions.41 However note that the process observed in the present work is reversible – the spectra return to the initial look after the sample is heated. It means no water was gone from the samples and this was confirmed repeatedly weighting the samples. A quite logical question arises – what are the state and properties of water after the slow self-assembling is completed. It is obvious to assume that the growing of columnar stacks is accompanied by the segregation of water molecules in the intercolumnar areas. The state of confined water was studied by analyzing the temperature dependencies of the 1H NMR shift of H2O and the Raman band shape. The Raman spectra exhibit fewer changes (Fig. 7) and thus appear to be less sensitive and less informative than NMR. Weak gradual changes in the band intensities can only be seen in the range of internal vibrations of SSY molecules covering 300 1800 cm–1 (Fig. 7). Similar observation were recorded when SSY aqueous solutions were studied by means of polarized Raman spectroscopy.34 The spectral range of 2800 - 3800 cm–1, where the bands of O–H stretching vibrations of water molecules are located, is expected to be more informative. However, the intensities of these bands are significantly lower than those of SSY molecules, and in order to analyze them additional digital processing (base line correction, noise smoothing) was necessary. And indeed, it is interesting to note that in the vicinity of the N → N+M phase transition the observed changes of Raman band shape in the range of 2800 - 3800 cm–1 are mimicking the νO–H...O band intensity evolution in the neat water at the melting point.42 Namely, increase in temperature and the crossing TNM ∼ 55 oC induces the redistribution of band intensities at ∼ 3185 cm–1 and ∼ 3397 cm–1 (Fig. 7) in very similarly manner as was observed in the neat water (3138 cm–1 and 3385 cm–1 at 0 oC).42 Furthermore, after the slow self-

ACS Paragon Plus Environment

21

The Journal of Physical Chemistry 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 32

assembling was completed the slope of the dependency of the 1H chemical shift of water on temperature during the heating cycle was deduced to be – 0.013 ppm/deg, i.e. practically the same as for the neat water (– 0.012 pm/deg, Ref 40 and Refs cited therein). These findings allow concluding that (i) water in the intercolumnar areas can be considered as neat and that (ii) the degrees of freedom of molecular motion in confined water are changing at the border to M phase in similar manner as at the melting of ice. However, the former process takes place at TNM ∼ 55 o

C, i.e. at significantly higher temperature than the melting point of water. On the other hand it is

well known that the geometric confinement can cause the shifts of the melting point, e. g. water in mesoporous carbon materials.43

Figure 7. Temperature evolution of Raman spectra of 1.2 mol/kg SSY aqueous solution by the heating from 25 to 85 °С (25, 35, 45, 55, 65, 75, 85 oC) covering N → N+M phase transition at TNM ∼ 55 oC in the whole range and around the stretching O–H vibrations of water molecules (zoom sector).

ACS Paragon Plus Environment

22

Page 23 of 32 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

The Journal of Physical Chemistry

Concluding remarks The joint application of NMR, Raman spectroscopy and DFT methods allowed to reveal fine details in temperature and composition effects passing through various phases of SSY aqueous solutions from the isotropic (I) to the principal chromonic phases – nematic (N) and columnar (M) ones. The phase transitions I ↔ I+N ↔ N look to be 'soft' enough from the point of view that no drastic discontinuous changes have been noticed in both Raman as well as in NMR spectra. The changes of spectral parameters covering these phase transitions are reversible in both direction (heating and cooling) at the traditionally used temperature variation rates. The equilibration time for N+M → M transition is very long, most probably because of slow supramolecular restructuring, i.e. the growing of hexagonal stack arrays in M phase. The growing of SSY aggregates is accompanied by the segregation of water in the intercolumnar areas. This restricts the exchange between H2O and N–H protons in the internal layers of the columnar stacks. The N–H signal in 1H NMR spectra observed in the present work for the first time is supports this statement. The joint analysis of temperature dependences of H2O chemical shift and Raman νO–H band shape has shown that water confined in intercolumnar areas behaves as in a neat substance. Water molecules in the intercolumnar space are confined possibly in a form of droplets or 'water pockets', as it was found recently in some ionic liquid/water solutions.44 Since LCLC molecules and ionic liquids have ionic groups at the periphery, such from the first sight far analogy between them can have a sense and can be a challenging goal for future experiments.

ACS Paragon Plus Environment

23

The Journal of Physical Chemistry 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 32

The tautomeric equilibrium in SSY is strongly shifted towards the hydrazone form – less than one azo tautomer is present among ∼ 100 hyrazone tautomers. The dominance of hydrazone form was confirmed experimentally using the long-range 1H–15N correlation (HMBC) and by the DFT SMD calculation. The LCLCs can appear to be very intriguing systems to study H-bonding and proton transfer (PT) phenomena. Discussions and recent works show persisting or even boosting interest in these phenomena (Ref. 45 and Refs cited therein). The hydrazone form of SSY, according to the equilibrium constant pKT = 2.5, can be considered as the protonated H-bond system N2–H1...O (Fig. 1), where 1H proton is most of the time located in the energy minimum near the N2 nitrogen, and jumps to the O–C4 oxygen are rare. However, there are many other LCLCs with the intramolecular H-bonds (e. g. Acid Red 27, RU31156, etc., Ref. 7) in which the low-barrier H-bond could be formed and thus the bridge proton motion of much larger amplitude could be expected. The reversing of the system to the N...H–O bond could be realized then by various external stimuli (solvent, added ions, temperature, etc). Crucial role of medium in PT processes is stressed almost in all studies on this topic. Therefore in the most optimistic scenario the chromonic phases could provide a unique possibility to study the effect of crossover from the solvent reaction field, which acts in highly diluted LCLC, to the crystal force field in large molecular stacks in the M-phase on PT processes.

AUTHOR INFORMATION Corresponding Author E-mail address: [email protected] (V. Balevicius).

ACS Paragon Plus Environment

24

Page 25 of 32 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

The Journal of Physical Chemistry

Notes The authors declare no competing financial interest.

ACKNOWLEDGMENT The work was carried out in the frame of Lithuanian - Ukrainian bilateral agreement supported by Ministries of Education and Science of Lithuania and Ukraine (grants TAP-LU-15-017 and M/49-2016). The authors acknowledge Center of Spectroscopic Characterization of Materials and

Electronic/Molecular

Processes

(Scientific

infrastructure

"Spectroversum"

www.spectroversum.ff.vu.lt) at Lithuanian National Center for Physical Sciences and Technology for use of spectroscopic equipment. Computations were performed on resources at the High Performance Computing Center”HPC Sauletekis” of Vilnius University.

REFERENCES (1) Liquid Crystals beyond Displays: Chemistry, Physics, and Applications; Li, Q., Ed.; Wiley: Hoboken, 2012, 600 p. (2) Lee, Y. S. Self-Assembly and Nanotechnology: A Force Balance Approach; Wiley: Hoboken, 2008, 344 p. (3) Park, H. S.; Lavrentovich, O. D. In Liquid Crystals Beyond Displays: Chemistry, Physics, and Applications; Li, Q., Ed.; Wiley: Hoboken, 2012; pp 449–484.

ACS Paragon Plus Environment

25

The Journal of Physical Chemistry 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 32

(4) Horowitz, V. R.; Janowitz, L. A.; Modic, A. L.; Heiney, P. A.; Collings, P. J. Aggregation Behavior and Chromonic Liquid Crystal Properties of an Anionic Monoazo Dye. Phys. Rev. E 2005, 72, 041710. (5) Lee, Y. S. Assembly of Lyotropic Liquid Crystals with Solid Crystal’s Structural Order Translated from the Lipid Rafts in Cell Membranes. J. Am. Chem. Soc. 2017, DOI: 10.1021/jacs.7b06720 (6) Lydon, J. Chromonic Review. J. Mater. Chem. 2010, 20, 10071–10099. (7) Lydon, J. Chromonic Liquid Crystalline Phases. Liquid Crystals 2011, 38, 1663–1681. (8) Park, H. S.; Kang, S. W.; Tortora, L.; Nastishin, Y.; Finotello, D.; Kumar, S.; Lavrentovich, O. D. Self-Assembly of Lyotropic Chromonic Liquid Crystal Sunset Yellow and Effects of Ionic Additives. J. Phys. Chem. B 2008, 112, 16307–16319. (9) Shiyanovskii, S. V.; Lavrentovich, O. D.; Schneider, T.; Ishikawa, T.; Smalyukh, I. I.; Woolverton, C. J.; Niehaus, G. D.; Doane, K. J. Lyotropic Chromonic Liquid Crystals for Biological Sensing Applications. Mol. Cryst. Liq. Cryst. 2005, 434(1), 259/[587])–270/[598]. (10) Luk, Y.-Y.; Campbell, S. F.; Abbott, N. L.; Murphy, C. J. Non-toxic Thermotropic Liquid Crystals for Use with Mammalian Cells. Liquid Crystals 2004, 31, 611–621. (11) Chami, F.; Wilson, M. R. Molecular Order in a Chromonic Liquid Crystal: A Molecular Simulation Study of the Anionic Azo Dye Sunset Yellow. J. Am. Chem. Soc. 2010, 132, 7794– 7802.

ACS Paragon Plus Environment

26

Page 27 of 32 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

The Journal of Physical Chemistry

(12) Nastishin, Y. A.; Liu, H.; Shiyanovskii, S. V.; Lavrentovich, O. D.; Kostko, A. F.; Anisimov, M. A. Pretransitional Fluctuations in the Isotropic Phase of a Lyotropic Chromonic Liquid Crystal, Phys. Rev. E 2004, 70, 051706. (13) Zhou, S.; Nastishin, Y. A.; Omelchenko, M. M.; Tortora, L.; Nazarenko, V. G.; Boiko, O. P.; Ostapenko, T.; Hu, T.; Almasan, C. C.; Sprunt, S. N.; et al. Elasticity of Lyotropic Chromonic Liquid Crystal Probed by Director Reorientation in a Magnetic Field, Phys. Rev. Lett. 2012, 109, 037801. (14) Ostapenko T.; Nastishin, Y. A.; Collings, P. J.; Sprunt, S. N.; Lavrentovich, O. D.; Gleeson, J. T. Aggregation, Pretransitional Behavior, and Optical Properties in the Isotropic Phase of Lyotropic Chromonic Liquid Crystals Studied in High Magnetic Fields, Soft Matter 2013, 9, 9487–9498. (15) Zhou, S.; Cervenka, A. J.; Lavrentovich, O. D. Ionic-Content Dependence of Viscoelasticity of the Lyotropic Chromonic Liquid Crystal Sunset Yellow, Phys. Rev. E 2014, 90, 042505. (16) Collings, P. J.; Goldstein, J. N.; Hamilton, E. J.; Mercado, B. R.; Nieser, K. J.; Regan, M. H. The Nature of the Assembly Process in Chromonic Liquid Crystals, Liquid Crystals Reviews 2015, 3, 1–27. (17) Joshi, L.; Kang, S. W.; Agra-Kooijman, D. M.; Kumar, S. Concentration, Temperature, and pH Dependence of Sunset-Yellow Aggregates in Aqueous Solutions: An X-Ray Investigation. Phys. Rev. E 2009, 80, 041703.

ACS Paragon Plus Environment

27

The Journal of Physical Chemistry 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 28 of 32

(18) Ding, X.; Stringfellow, T. C.; Robinson, J. R. Self-Association of Cromolyn Sodium in Aqueous Solution Characterized by Nuclear Magnetic Resonance Spectroscopy. J. Pharm. Sci. 2004, 93, 1351–1358. (19) Troche-Pesqueira, E.; Cid, M. M.; Navarro-Vázquez, A. Disodium Cromoglycate: Exploiting its Properties as a NMR Weak-aligning Medium for Small Organic Molecules. Org. Biomol. Chem. 2014, 12, 1957–1965. (20) Yamaguchi, A.; Smith, G. P.; Yi, Y.; Xu, C.; Biffi, S.; Serra, F.; Bellini, T.; Zhu, C.; Clark, N. A. Phases and Structures of Sunset Yellow and Disodium Cromoglycate Mixtures in Water. Phys. Rev. E 2016, 93, 012704. (21) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; et al. Gaussian 09, Revision D.01; Gaussian, Inc.: Wallingford, CT, 2010. (22) Becke, A. D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. (23) Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. Self-Consistent Molecular Orbital Methods. XX. A Basis Set for Correlated Wave Functions. J. Chem. Phys. 1980, 72, 650–654. (24) McLean, A. D.; Chandler, G. S. Contracted Gaussian Basis Sets for Molecular Calculations. I. Second Row Atoms, Z=11-18. J. Chem. Phys. 1980, 72, 5639–5648. (25) Adamo, C.; Barone, V. Toward Reliable Density Functional Methods without Adjustable Parameters: The PBE0 Method. J. Chem. Phys., 1999, 110, 6158–6170.

ACS Paragon Plus Environment

28

Page 29 of 32 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

The Journal of Physical Chemistry

(26) Aidas, K.; Møgelhøj, A.; Kjær, H.; Nielsen, C. B.; Mikkelsen, K. V.; Ruud, K.; Christiansen, O.; Kongsted, J. Solvent Effects on NMR Isotropic Shielding Constants. A Comparison between Explicit Polarizable Discrete and Continuum Approaches. J. Phys. Chem. A 2007, 111, 4199−4210. (27) Klimavicius, V.; Gdaniec, Z.; Kausteklis, J.; Aleksa, V.; Aidas, K.; Balevicius, V. NMR and Raman Spectroscopy Monitoring of Proton/Deuteron Exchange in Aqueous Solutions of Ionic Liquids Forming Hydrogen Bond: A Role of Anions, Self-Aggregation, and Mesophase Formation. J. Phys. Chem. B 2013, 117, 10211–10220. (28) Balevicius, V.; Gdaniec, Z.; Aidas, K. NMR and DFT Study on Media Effects on Proton Transfer in Hydrogen Bonding: Concept of Molecular Probe with an Application to Ionic and Super-polar Liquids. Phys. Chem. Chem. Phys. 2009, 11, 8592–8600. (29) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. (30) Zhao, Y.; Schultz, N. E.; Truhlar, D. G. Design of Density Functionals by Combining the Method of Constraint Satisfaction with Parameterization for Thermochemistry, Thermochemical Kinetics, and Noncovalent Interactions. J. Chem. Theory Comput. 2006, 2, 364–382. (31) Hehre, W. J.; Ditchfield, R. J.; Pople, J. A. Self-Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian-Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules. J. Chem. Phys. 1972, 56, 2257–2261.

ACS Paragon Plus Environment

29

The Journal of Physical Chemistry 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 30 of 32

(32) Hariharan, P. C.; Pople, J. A. The Influence of Polarization Functions on Molecular Orbital Hydrogenation Energies. Theor. Chim. Acta 1973, 28, 213–222. (33) Edwards, D. J.; Jones, J. W.; Lozman, O.; Ormerod, A. P.; Sintyureva, M.; Tiddy, G. J. T. Chromonic Liquid Crystal Formation by Edicol Sunset Yellow. J. Phys. Chem. B 2008, 112, 14628–14636. (34) Yao, X.; Nayani, K.; Park, J. O.; Srinivasarao, M. Orientational Order of a Lyotropic Chromonic Liquid Crystal Measured by Polarized Raman Spectroscopy. J. Phys. Chem. B 2016, 120, 4508–4512. (35) Renshaw, M. P.; Day, I. J. NMR Characterization of the Aggregation State of the Azo Dye Sunset Yellow in the Isotropic Phase. J Phys Chem. 2010, 114, 10032–10038. (36) Gilli, P.; Bertolasi, V.; Ferretti, V.; Gilli, G. Evidence for Intramolecular N–H...O Resonance-Assisted Hydrogen Bonding in β-Enaminones and Related Heterodienes. A Combined Crystal-Structural, IR and NMR Spectroscopic, and Quantum-Mechanical Investigation. J. Am. Chem. Soc. 2000, 122, 10405–10417. (37) Klimavicius, V.; Bacevicius, V.; Gdaniec, Z.; Balevicius, V. Pulsed-field Gradient 1H NMR Study of Diffusion and Self-aggregation of Long-chain Imidazolium-based Ionic Liquids. J. Mol. Liq. 2015, 210, 223–226. (38) Wu, L.; Lal, J.; Simon, K. A.; Burton, E. A.; Luk, Y. Y. Nonamphiphilic Assembly in Water: Polymorphic Nature, Thread Structure, and Thermodynamic Incompatibility. J. Am. Chem. Soc. 2009, 131, 7430–7443.

ACS Paragon Plus Environment

30

Page 31 of 32 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

The Journal of Physical Chemistry

(39) Zhou, S.; Neupane, K.; Nastishin, Y. A.; Baldwin, A. R.; Shiyanovskii, S. V.; Lavrentovich, O. D.; Sprunt, S. Elasticity, Viscosity, and Orientational Fluctuations of a Lyotropic Chromonic Nematic Liquid Crystal Disodium Cromoglycate. Soft Matter 2014, 10, 6571–6581. (40) Balevicius, V.; Aidas, K. Temperature Dependence of 1H and 17O NMR Shifts of Water: Entropy Effect. Appl. Magn. Reson. 2007, 32, 363–376. (41) Davidson, Z. S.; Huang, Y.; Gross, A.; Martinez, A.; Still, T.; Zhou, C.; Collings, P. J.; Kamien, R. D.; Yodh, A. G. Deposition and Drying Dynamics of Liquid Crystal Droplets. Nat. Commun. 2017, 8:15642, DOI: 10.1038/ncomms15642 (42) Duričkovic, I.; Claverie, R.; Bourson, P.; Marchetti, M.; Chassot, J. M.; Fontana, M. D. Water – Ice Phase Transition Probed by Raman Spectroscopy. J. Raman Spectrosc. 2011, 42, 1408–1412. (43) Xu, Y.; Watermann, T.; Limbach, H. H.; Gutmann, T.; Sebastiani, D.; Buntkowsky, G. Water and Small Organic Molecules as Probes for Geometric Confinement in Well-ordered Mesoporous Carbon Materials. Phys. Chem. Chem. Phys. 2014, 16, 9327–9336. (44) Saihara, K.; Yoshimura, Y.; Ohta, S.; Shimizu, A. Properties of Water Confined in Ionic Liquids. Sci. Rep. 2015, 5,10619; doi: 10.1038/srep10619. (45) Lu, J.; Hung, I.; Brinkmann, A.; Gan, Z.; Kong, X.; Wu, G. Solid-State 17O NMR Reveals Hydrogen-Bonding Energetics: Not All Low-Barrier Hydrogen Bonds Are Strong. Angew. Chem. Int. Ed. 2017, 56, 6166–6170.

ACS Paragon Plus Environment

31

The Journal of Physical Chemistry 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 32 of 32

TOC Graphic

ACS Paragon Plus Environment

32