Mikhail Kim‡
a,
Coral Hillel‡b,
Kayrel Edwardsa,
William Pietroc,
Ozzy Mermutbc and
Christopher J. Barrett*ab
aDepartment of Chemistry, McGill University, Montreal, QC, Canada. E-mail: christopher.barrett@mcgill.ca
bDepartment of Physics and Astronomy, York University, Toronto, ON, Canada
cDepartment of Chemistry, York University, Toronto, ON, Canada
First published on 16th August 2024
Biopolymer composite materials were prepared by combining bio-sourced cationic water-soluble chitosan with bi-functional water-soluble anionic azo food dyes amaranth (AMA) or allura red (ALR) as ionic cross-linkers, mixing well in water, and then slow-drying in air. The electrostatically-assembled ionically-paired films showed good long-term stability to dissolution, with no re-solubility in water, and competitive mechanical properties as plastic materials. However, upon exposure of the bioplastics to low power light at sunlight wavelengths and intensities stirring in water, the stable materials photo-disassembled back to their water-soluble and low-toxicity (edible) constituent components, via structural photo-isomerization of the azo ionic crosslinkers. XRD, UV-vis, and IR spectroscopy confirmed that these assemblies are reversibly recoverable and so can in principle represent fully recyclable, environmentally degradable materials triggered by exposure to sunlight and water after use, with full recovery of starting components ready for re-use. A density functional theory treatment of the amaranth azo dye identified a tautomeric equilibrium favouring the hydrazone form and rationalized geometrical isomerization as a mechanism for photo-disassembly. The proof-of-principle suitability of films of these biomaterial composites as food industry packaging was assessed via measurement of mechanical, water and vapour barrier properties, and stability to solvent tests. Tensile strength of the composite materials was found to be 25–30 MPa, with elongation at break 3–5%, in a range acceptable as competitive for some applications to replace oil-based permanently insoluble non-recyclable artificial plastics, as fully recyclable, recoverable, and reusable low-toxicity green biomaterials in natural environmental conditions.
An alternative strategy involves designing polymer materials that can be assembled from non-toxic and readily recyclable components that are capable of reversible disassembly under specific conditions or stimuli. Such a disassembly process for example could selectively release soft bonds holding the constitutional parts together with desired properties for use, while keeping the covalent bonds intact. Some such previously reported polymer systems have been shown to disassemble on exposure to stimuli including temperature change,7 pH change,8 exposure to metal ions with chelating agents,9 mechanical force,10 or UV light.10–13 Exposure to these relatively harsh conditions however also risks the chemical or mechanical degradation of the polymer components covalently, potentially reducing recycling efficiency and re-use properties. Additionally, pH-responsive systems and those needing metal ions and chelating agents often require substantial water-based solutions of acids or bases, complicating the disassembly process and recovery of constitutional parts. Furthermore, these approaches often involve complex synthesis and result in materials mainly existing in the form of hydrogels, limiting their practical use, for instance as packaging materials.
In our work, visible light was chosen as a relatively clean, abundant, and gentle stimulus to induce photo-disassembly of water-insoluble bioplastic materials multi-layered on surfaces from water-soluble polyelectrolytes and water-soluble azo dyes as photo-reversible ionic crosslinkers, depicted schematically in Fig. 1.14 The azo dyes isomerize upon light irradiation, changing their geometry significantly from a planar trans isomer to a bent, metastable cis isomer.15 This isomerization affects structural interactions between polymer and azo dyes, gently ‘shaking’ the assemblies apart and recovering the water-solubility of each part.14 This concept of structural photo-release with azo dyes has been reported previously in some specialized systems, providing light-control over structure or solubility. For example, it was demonstrated that trans-to-cis isomerization enhanced the aqueous solubility of an anti-cancer candidate drug in aqueous environment,16 and a drug from an azobenzene–polyethylene glycol-nanoparticles micelle system was similarly shown to be released with light.17 In other previous work,18 co-crystallization of a fluorinated azobenzene with volatile components yielded halogen-bonded cocrystals that could photo-disassemble to be cut, carved, or engraved with low-power visible laser light.
We propose here to extend this light-triggered disassembly to low-toxicity and fully recyclable bioplastics, by combining relatively harmless anionic azo food dyes with positively charged biopolymers such as chitosan,19 now used widely in toxicity-sensitive materials applications, such as cosmetics, medicine, agriculture, and the food industry.20–22
Previously, our group has reported the aqueous assembly of materials from water solutions of polyelectrolytes such as sulphonated cellulose and sodium alginate, and water solutions of azo dyes, such as Bismarck brown Y, using a layer-by-layer (LbL) dipping technique to produce thin multilayer films on a glass substrate.14 The photo-induced disassembly of these photo-reversible materials in water was demonstrated under exposure to low-power visible light while being gently washed with water. However, the preparation of multilayer thin films by the LbL method was tedious. Moreover, they were not strong mechanically or free-standing, and they were limited to just 10s or a few 100s of nanometres in thickness, requiring a supporting substrate. In one report,19 an attempt to make thicker LbL films approaching micrometre thickness was attempted by spray-coating LbL multilayers. However, that technique was still very time-consuming and costly. For example, the production of an 80-bilayer film with a thickness of just 100 nm took 48 h. For mass production and use of such light-disassembly materials, larger-scale methods must be applied to produce the thick, strong, free-standing films many micrometres thick required for any real plastic applications, such as for packaging.
In our present work, an aqueous solvent-casting method was developed as a facile, low-cost, and high-throughput method to make thick, strong, and free-standing photo-reversible plastic films that could dis-assemble on visible light irradiation. Materials prepared by this method were water-resistant under ambient conditions and robust, yet demonstrated light-disassembly in water under gentle shaking during irradiation with blue or green light simulating sunlight. To assess the potential of these chitosan-azo dye films as replacements for petrol-based films in packaging applications, the mechanical and barrier properties of the films were measured and compared to polyethylene.
Advancing from previously-used fabric azo dyes such as Bismarck brown Y and Bismarck brown R, to low-toxicity azo food dyes such as allura red (ALR) and amaranth (AMA) has the advantage of offering systems with lower environmental impact, yet this now complicates experimental spectroscopic confirmation of the trans-to-cis isomerization, as these dyes now contain a hydroxy-group in ortho ring position with respect to the azo-group, leading to hydrazone character and associated exceedingly rapid thermal cis-to-trans back-isomerization times, too fast to observe directly with standard spectroscopies. So lastly, in place of experimental spectroscopic confirmation of photo-switching between geometric forms, a theoretical analysis of AMA was conducted via density functional theory (DFT) approaches, as a representative system for the general class of ortho-hydroxy azobenzene food dyes.
During the photo-disassembly process, the azo dye disassembled from the chitosan polymer and re-dissolved in the water solution. At time t = 0, no azo dye was present in the solution as validated with UV-vis spectroscopy. After slow re-dissolving of azo-dye and chitosan, the concentration of both components increases and can be readily detected and measured by tracking their UV-vis spectra. Since azo dyes have characteristic absorbance in the visible region, plots of dye absorbance maxima (ALR: λ = 510 nm and AMA: λ = 530 nm) versus time were used to quantify the rate of the film disassembly.
The FTIR spectra of pure CS films, powder azo dyes, and CS-azo dye films are shown in Fig. S2.† For CS and CS-azo dye films, the bands 3200–3500 cm−1 are assigned as overlapped stretching O–H on N–H bands. The band near 2867 cm−1 is assumed to be stretching (–CO) of the amide group CONHR of the CS.37 The presence of residual N-acetyl groups was confirmed by the bands near 1645 cm−1 (CO stretching of amide I) and 1320 cm−1 (C–N stretching of amide III), respectively. The band at 1550 cm−1 corresponds to N–H bending of amide II and is likely occluded by a band at 1548 cm−1 corresponding to the N–H bending of the primary amine.38 For both azo dyes, the two bands at 1038 and 1136 cm−1 are assigned to the coupling between naphthalene rings and stretching of SO3, and the bands at 1038, 1186, and 1200 cm−1 are linked to the SO3 stretching mode.39,40 Azo bond (NN) vibrations are identifiable between 1504 cm−1 and 1550 cm−1.41 The effect of the ionic interaction between polymer and azo dye was observed by the shifting peak of sulfonate groups of azo dyes from 1038 cm−1 to 1030 cm−1 and a shifting of the primary amine peak from 1548 cm−1 to 1559 cm−1.
Fig. S3† shows XRD patterns of CS–ALR and CS–AMA films. From the pattern, the two characteristic peaks are visible at 2θ = 11° and 2θ = 20°, corresponding to count indexes (142) and (191), respectively, assigned to the CS.42 The first peak at 11° becomes wider and smaller with decreasing content of CS in the film, while the second peak at 20° becomes sharper and stronger. The weak peaks in the XRD pattern of CS reflect great disarray in chain alignment of CS,42 and the broadening of the peak at 11° toward 9° and increase in the 20° peak confirm the presence of the azo dye. In this study, the crystalline index (CrI) was determined from eqn (1) by the fitting of the diffraction profiles in Fig. S3.† This formula uses the ratio between the intensity of a crystalline reflection and intensity of the minimum peak at 2θ = 16°, which describes the diffuse halo peak (Iam). Intensity of the reflection at 20° (I200) is determined for the most frequently used hydrated form of CS.43
CrI = (I200 − Iam)/I200 | (1) |
The lowest calculated CrI was 38% for CS films and this increased to a maximum of 56% for CS–ALR films with the mass ratio CS/ALR = 2.5/1. The higher degree of crystallinity of the CS-azo dye films can be explained by presence of crystalline azo dye and not necessarily by an increase of crystallinity of CS itself, which is also supported by the mechanical analysis of the CS-azo dye films. The higher crystallinity of CS itself would be expected to increase the tensile strength,44 however the measured strength of CS-azo dye films was in fact lower (Section 3.6).
We hypothesize that the disassembly of the CS-azo dye films occurs due to the geometric trans-to-cis isomerization of the azo dyes during irradiation with the blue or green light. Under ambient low intensity white light, both azo dyes ALR and AMA have a planar elongated orientation within the film. Fig. 1 shows structures of AMA and ALR which contain negatively charged sulfonate (SO3−) groups on opposite sides of the molecule. These negatively charged sulfonate groups can ionically bond strongly to the positively charged protonated amino groups located along the CS polymer chains. The azo dyes thus could effectively act as soft-bonded reversible crosslinkers holding the CS chains together, resulting in a water-resistant and robust material. During constant irradiation with blue or green light, the azo dyes undergo a geometric rearrangement from the trans isomer to the cis isomer, which changes the shape, orientation, and length of the crosslinker. This movement appears sufficient to disrupt or weaken the ionic bonding between the azo dyes and the CS polymer, which exposes the positive and negative ionic groups cross-linking the assemblies together, re-solubilizing each component in the moving water and, in principle, enabling them to be separated from each other. Indeed, a similar observation was made for the previously reported multilayer sulfonated cellulose-azo dye films prepared on supported substrates,14 in which the trans-to-cis geometric rearrangement of the azo dye was slow enough to be studied by pump-probe experiments and confirmed to be coincident with the photo-driven disassembly of the films.
Plotted in Fig. 5 is the absorbance vs. time profile for light-triggered disassembly of CS–AMA (Fig. 5A) and CS–ALR (Fig. 5B) compared to control vials where films were kept in the dark. In the dark, we observed the initial appearance of both the ALR and AMA dyes in the solution, which corresponded to a concentration of 0.015 mg mL−1 which, as discussed previously, we attributed to un-bound excess azo dye that wasn't completely released during the washing step. This initial concentration did not increase under dark conditions, except when exposed to light. Typical times for complete film disassembly were 150–250 h of irradiation at approximately 100 mW under both green and blue light. To determine the rate of increase in the absorbance of the azo dye released into solution (which equates to the rate of disassembly of the material), absorbance against time was further plotted (Fig. S5†) as both 1/absorbance vs. time and ln(absorbance) vs. time. We found that the R2 value was highest for the absorbance-time linear fit, which led us to conclude that the disassembly of the multilayer films can best be described by a zero-order process. The rates of disassembly in light were normalized to the weight of the samples and to the disassembly rate of samples kept in the dark. These normalized rates were termed ‘relative rates of disassembly’ (RRD) for comparison between samples and systems. For the CS–AMA material, the RRD ratio was determined to be 117:1, while the RRD for CS–ALR was determined to be 14:1. This indicated that for both CS-azo dye films, the films disassembled at a much faster rate when exposed to the light stimulus. The significantly higher RRD observed for the CS–AMA compared to CS–ALR may be rationalized by the differences in the structures of the two azo dyes with respect to charges (AMA is tri-functional while ALR is bi-functional), and perhaps also by ALR having a more favourable geometry for ionic bonding with the CS polymer which slows down the disassembly process.
CS–NH2(s) + CH3COOH(aq) → CS–NH3+CH3COO(aq)− | (2) |
3CS–NH3+CH3COO(aq)− + AMA(SO3−)3(aq) → 3CS–NH3+CH3COO(aq)−AMA(SO3−)3(s) | (3) |
CS–NH3+CH3COO−AMA(SO3−)3(s) → CS–NH3+AMA(SO3−)3(s) + CH3COO−(g) | (4) |
CS is produced commercially by deacetylation of chitin, and the proportion of amino groups in CS is related to the degree of deacetylation that describes the ratio of glucosamine to acetyl-glucosamine groups. For the CS used in the study the degree of deacetylation was 83%, determined by FTIR spectroscopy. Upon dissolving in acetic acid, the amino groups become protonated, and CS acetate solution can participate in the interaction with azo dyes. The molar ratio of amino groups of CS to sulfonate groups of azo dyes in the fabricated free-standing films varied from 15–315 to 1, i.e., from 5 to 105 times the stoichiometric amount according to eqn (3). Thus, the CS-azo dye films were made primarily of CS and contained protonated amino groups in a form of acetic acid salt. Upon drying and heating, the films release free acetic acid and some groups become unprotonated (eqn (4)), suggesting that to start the disassembly process, the washing solution should be more acidic. A series of experiments to determine the effect of water pH on the ability of the films to light-disassemble is shown in Fig. 6, where one can see that the disassembly of CS-azo dye films started only at pH conditions lower than the range of 5.5 and 6.8. Future studies will include chemical modifications of CS to increase the pH at which CS-azo dye films disassemble, such as depolymerization,45,46 or introducing into the CS structure acrylate groups.46
Fig. 6 Disassembly of thick film CS–AMA under 523 nm green light irradiation over 3 days duration shaking in water at different pH values. |
The degree of protonation of CS was regulated by conducting another series of experiments whereby 0.1 M NaOH was added to a completely dissolved (fully protonated) CS solution and the pH recorded. Using the Henderson–Hasselbalch equation (eqn (5)), we calculated the degree of protonation of CS, as eqn (5) relates the pH of the CS solution to the pKa of the weak acid CS–NH3+ (where CS- is the D-glucosamine plus N-acetyl-D-glucosamine residual of CS) with concentrations in equilibrium of deprotonated ([CS–NH2]) and protonated ([CS–NH3+]) form of CS. The acid–base equation for protonated CS is shown in eqn (6), with results provided in Table 1.
pH = pKa + log10([CS–NH2]/[CS–NH3+]) | (5) |
CS–NH3+ → CS–NH2 + H+ | (6) |
CS protonation, (%) | Mass ratio CS/AMA | Disassembly rate ratio |
---|---|---|
99.7 | 5 | 1.5 |
98.5 | 100 | 1 |
98.5 | 25 | 1 |
98.5 | 20 | 2.8 |
98.5 | 5 | 3.5 |
97.0 | 5 | 4.7 |
The ratio between disassembly rates in light and dark indicates how sufficiently the azo dye crosslinks CS and to what extent the film is light responsive. The higher the ratio, the more light was required in the disassembly process. Remarkably, this ratio was higher for films made with CS with partially unprotonated amino groups (4.7 for 97% protonated vs. 1.5 for 99.7% protonated, Table 1). The highly protonated CS films had almost equal solubility in the dark and when irradiated. With the mass ratio CS/azo dye approaching the stoichiometric ratio (for protonation degree 98.5% in Table 1), the ratio of disassembly rates increased. Thus, to achieve an optimal disassembly rate in light vs. dark, an optimal amount of protonated amino groups in the initial CS and in the film after cross-linking must be found.
Additionally, the mechanical tensile strength of CSH–ALR films (mass ratio 1:0.125) was tested in both dry and wet states (Fig. 7B). The tensile stress at break was much higher for dry films but elongation at break was 5 times larger for the wet films, indicating that properties of the films may be tunable according to the application, as for example, some biomedical applications prioritize higher elasticity of the films over toughness. The values measured for CSH–ALR (tensile strength 15–40 MPa with elongation at the break 400–600%, WVTR = 4.7–7.8 g day−1·m−2 and OTR = 2300–3100 cc (m−2 day−1)) compare reasonably well with those of standard plastic packaging.46
To demonstrate a proof-of-principle recycling ability for the CS-azo dye films, a basic recycling process was simulated as shown in Fig. 8. Pieces of the CS–AMA film were cut and mixed together with common domestic waste slurries including egg shell, organic waste, paper, polyethylene (PE), and aluminium foil. The mixture was submerged in water and under stirring was exposed to green light over 5 days. After 5 days, the water solution turned a deep red colour owing to the release of the azo dye into solution, and the liquid component was separated from the solid residue by filtration. To separate the dissolved CS from the azo dye, the pH of the solution was adjusted to basic (pH = 9) using NaOH. CS is insoluble at pH above 6.3, and therefore precipitated out of solution, leaving the azo dye dissolved. The precipitated CS was then separated by filtration and the residual azo dye removed using carbon black. The recollected CS was then redissolved in an acidic solution and solvent-cast to create a second-generation film. Fig. 9 shows the mechanical properties of the recovered CS film which had a lower tensile strength (10 MPa) and elongation at break (2%) compared to the fresh CS film (40 MPa) (12%), yet is encouraging for a first attempt. Further optimization of this recycling process is currently being studied to produce second-generation CS films with mechanical properties that are comparable to the first-generation products, such as developing a more selective filtering process (step 7) than the use of simple micro-filters used to separate CS solution from activated charcoal, which may have filtered out preferentially the high molecular weight polymer chains of CS and resulted in reduction of mechanical properties.
Fig. 9 Mechanical properties of CSH–AMA thick films in dry and wet states, and after recovery (2nd generation). |
Film, mass ratio | Density, g cm−3 | Density, g cm−3 |
---|---|---|
Volume by geometry | Volume by ethanol displacement | |
CS | 1.65 ± 0.10 | 1.54 ± 0.03 |
CS–AMA, 2.5/1 | 1.60 ± 0.12 | 1.43 ± 0.02 |
CS–AMA, 8/1 | 1.50 ± 0.08 | 1.38 ± 0.07 |
CS–ALR, 2.5/1 | 1.55 ± 0.09 | 1.34 ± 0.05 |
WVTR measures the amount in grams of water vapour transmitted through a fixed area of the film per day; the lower the WVTR, the less external moisture permeates through the film, and the higher the barrier properties of the material. As shown in Table 3, the WVTR increased upon the crosslinking of the CS polymer (WVTR of pure chitosan = 59 g day−1·m−2) with azo dyes AMA and ALR, WVTR = 143–176 g day−1·m−2 for CS–AMA and WVTR = 120–162 g day−1·m−2 for CS–ALR. Such changes might be explained by an increased porosity of the films, although as described by the comparison chart on Fig. S6,† such materials can still be used as a packaging materials for some types of food such as cheese and bakery products.8 The WVTR also can be decreased by additional crosslinking of CS or by using filling materials, or by decreasing porosity by applying hot-pressing after film fabrication.
Film | WVTR, g day−1 m−2 |
---|---|
CS | 59 ± 14 |
CS–AMA mass ratio 2.5/1 | 176 ± 22 |
CS–AMA mass ratio 5/1 | 143 ± 19 |
CS–AMA mass ratio 8/1 | 163 ± 15 |
CS–ALR mass ratio 2.5/1 | 146 ± 20 |
CS–ALR mass ratio 5/1 | 162 ± 17 |
CS–ALR mass ratio 8/1 | 120 ± 19 |
Contact angle (CA) measurements show that the CS and CS-azo dye films are strongly hydrophilic, with maximum contact angle 68±2° for CS film and lower CA for all films CS-azo dye with minimum 28±2° for CS–ALR films. To achieve more hydrophobic films, in the future a modification of CS with aldehydes (such as octanal47) might be possible.
We first investigated the enol–azo form of AMA, denoted structure 1. This structure of AMA where the azo double bond is intact is often reported in literature concerning the azo food dyes, as well as manufacturer data. All possible conformers a, b, c, and d of trans-AMA were investigated, corresponding to possible rotations of the naphthyl and naphthol rings about the azo bond as shown in Chart 1. Conformer 1c was the lowest-energy conformer, followed by 1d with a relative Gibb's free energy of 9.37 kJ mol−1 (Table 4). The orientation of the naphthol ring with respect to the azo group in both these structures enabled the formation of a pseudo-6-membered ring involving a hydrogen bond between the hydroxyl and nearby azo lone pair. In addition, this hydrogen bond tended to lock the azo group and the naphthol ring in the same plane, while the opposing naphthyl ring was twisted out of plane. This twist was commensurate with the relative energy of each conformer, where the value was 18.8° for 1c, 21.5° for 1d, and 35.5° for the highest-energy conformer 1b. In a DFT study of Ponceau 4R, an isomer of AMA, Bevziuk et al. also found that planar trans isomers were impossible, and that the lowest-energy conformer acquired a conformation which maximized the separation of the naphthyl and naphthol rings,53 as in trans-1c. Furthermore, all possible conformers of cis-1 were likewise investigated by rotating each conformer of 1 in Chart 1 about the azo bond, and then reoptimizing the resultant structure. Conformer 1a was lowest in energy, followed closely by 1c whose relative Gibb's free energy was 3.14 kJ mol−1 (Table 4). Like the trans isomer, the lowest-energy conformers were marked by the ability to form intramolecular hydrogen bonds (cis-1a), as well as the minimization of steric hindrance between the naphthol and naphthyl rings. Optimized structures of trans-1 and cis-1 minimum energy conformers are shown in Fig. 10A.
Chart 1 Possible conformers and isomeric forms of enol–azo AMA (1) and keto–hydrazone AMA (2) investigated at the B3LYP-D3BJ/ma-def2-TZVP level. Structures for cis-1 have been omitted for simplicity. |
Structure | ΔG, kJ mol−1 | |
---|---|---|
cis | trans | |
1a | 0.00 | 17.90 |
1b | 14.89 | 30.55 |
1c | 3.14 | 0.00 |
1d | 12.90 | 9.37 |
2a | 0.00 | 0.00 |
2b | 11.47 | 8.39 |
Fig. 10 DFT-optimized geometries of AMA, showing minimum-energy conformers for the trans and cis isomers of (A) the azo form 1 and (B) the hydrazone form 2. |
Azobenzenes ortho- or para-substituted with hydroxyl groups may exist in a tautomeric equilibrium of the enol–azo and keto–hydrazone forms. For the case of AMA, the proton is transferred from the hydroxyl group to the azo nitrogen adjacent to the naphthalene ring as shown in Scheme 1. The resonance stabilization in the hydrazone form that neutralizes the resulting charge separation concomitantly reduces the double-bond character of the azo linkage. Hydrazone dyes adopt a different naming convention from the azo dyes, where trans and cis now refer to configuration of functional groups about the CN imine bond. Moreover, literature conventions also classify the naphthol group in this case as the ‘rotor’, while the naphthyl group is the ‘stator’. For the hydrazone form, denoted structure 2, there are two possible conformers a and b for the trans and cis isomers corresponding to possible rotations of the naphthyl stator, as shown in Chart 1. In contrast to the azo form, the hydrazone form is locked into the lower energy, nearly planar cis isomer due to the formation of an intramolecular hydrogen bond between the tautomeric hydrogen and ketone oxygen in a 6-membered pseudo-ring. This hydrogen bond is lost in the trans isomer, accompanied by significant puckering of the naphthol ring due to steric interaction with the tautomeric hydrogen. Although other computational investigations tend to ignore possible rotations of the stator, our investigation explicitly verifies these assumptions, demonstrating that the rotation of the naphthyl stator in AMA is not significant. The conformers trans-2b and cis-2b presented Gibb's free energies of 8.39 kJ mol−1 and 11.47 kJ mol−1 relative to the lower energy conformers trans-2a and cis-2a, respectively (Table 4). Assuming a Boltzmann distribution of conformers, the contribution of trans-2b and cis-2b to the overall structures of the trans and cis hydrazone forms is negligible. Optimized structures of trans-2 and cis-2 minimum energy conformers are shown in Fig. 10B. Although the crystal structure of AMA has not been reported, we found agreement between our calculations and the crystallographic characterization of a similar dye, Pigment Red 40, where Heaney et al. isolated the planar cis hydrazone form and identified a covalent bond between the hydrogen atom and azo nitrogen.54 However, they found a hydrogen bond distance O⋯H of 1.85 Å, whereas our DFT calculations predicted 1.65 Å for cis-2a. This discrepancy may be attributed partly to structural differences between Pigment Red 40 and AMA, since sulphonation of the naphthol and naphthyl rings in AMA can affect the aromaticity and electronics of these rings, and hence the hydrogen bond.
Scheme 1 Equilibrium of enol–azo (NN) and keto–hydrazone (CN–N–H) tautomeric forms for azo dye AMA. Azo and hydrazone groups are highlighted in red. |
According to the relationship between the standard Gibb's free energy and the equilibrium constant for a reaction, it is possible to estimate the tautomeric equilibrium of the AMA azo and hydrazone forms. For a standard temperature and pressure of 298.15 K and 1 atm, and assuming a Boltzmann distribution of possible conformers, the standard Gibb's free energy for the unimolecular enol-to-keto (trans-1-to-cis-2) tautomerization is −6.89 kJ mol−1, corresponding to a 16:1 keto-to-enol ratio in the gas phase. Indeed, reports of other dyes based on a 1-arylhydrazone-2-naphthol skeleton have found that the keto form is more energetically favourable both in solution and the solid state, which points to an often erroneous classification of hydrazone dyes as azo dyes.54–56 In our previous work, DFT modelling of the azo food dye ALR identified an energetically favourable keto form where a nonplanar local minimum cis isomer about the azo bond could be optimized.52 Yet for AMA, this was not the case, as attempts to optimize the cis configuration about the azo bond yielded high-energy structures with respect to the hydrazone forms presented here, or simply recovered the planar hydrazone forms themselves. This observation leads us to consider possible mechanisms responsible for disassembly of the chitosan films. DFT predictions of the dipole moment show that isomerization from cis-2a to trans-2a is accompanied by a change from 8.77 D to 9.22 D, markedly less significant compared to the azo form, where isomerization from trans-1c to cis-1a results in a change from 9.89 D to 13.37 D. In addition, unlike azo isomerization where a nonplanar, disordered cis isomer is formed, hydrazone isomerization is uniquely described as an order-to-order transition57 where both isomers are nearly planar. Consequently, the typical changes associated with azobenzene isomerization—dipole moment and geometry—cannot be invoked to completely explain the mechanism of disassembly. Likely, the significant geometric change of the hydrazone isomerization process itself is responsible. This geometric change has the capability to modulate the bulk properties of soft-bonded supramolecular assemblies, which was recently shown for the chirality and polymer actuation of liquid crystals.58,59
Isomerization of the hydrazone form has been found to proceed by a keto-to-enol tautomerization which enables rotor rotation about a transient C–N bond.60–62 Other mechanisms have been reported, including in-plane inversion of the imine bond, or hybrid inversion-rotation mechanisms involving the entire hydrazone CN–N–H unit and a nonplanar intermediate.63–65 In addition, Deneva et al. used a simple explicit solvation scheme to demonstrate that water molecules present even as trace contaminants in organic solvent can catalyze the tautomerization process and lower the isomerization barrier.62 We have shown that AMA exists predominantly in the hydrazone form. Due to the loss of a favourable intramolecular hydrogen-bond upon isomerization from cis-2a to trans-2a, we suspect AMA may not exhibit bistability such as the hydrazones designed to kinetically lock supramolecular assemblies into different configurations for extended periods of time.58,59,66 Furthermore, since the presence of water in wet multilayer films can markedly accelerate the thermal isomerization process, we also suspect that, like ALR, AMA is a fast-isomerizing switch which is not easily amenable to spectroscopic verification. A thorough and comparative treatment of the azo food dyes including ALR, AMA, and others will be the subject of a forthcoming report. We plan to address the literature gap regarding the isomerization of the azo food dyes and employ multireference formulations of DFT for the correct treatment of rotational mechanisms.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra02211d |
‡ The first two authors MK and CH contributed equally and share first authorship. |
This journal is © The Royal Society of Chemistry 2024 |