Christopher D.
Jones
a,
Aled R.
Lewis
b,
Daniel. R.
Jones
b,
Christopher J.
Ottley
a,
Kaiqiang
Liu
c and
Jonathan W.
Steed
*a
aDurham University, South Road, Durham, DH1 3LE, UK. E-mail: jon.steed@durham.ac.uk
bSystems and Process Engineering Centre (SPEC), Energy Safety Research Institute (ESRI), College of Engineering, University of Swansea, Singleton Park, Swansea, SA2 8PP, UK
cKey Laboratory of Applied Surface and Colloid Chemistry (Ministry of Education), School of Chemistry and Chemical Engineering, Shaanxi Normal University, Xi'an, 710119, China
First published on 7th July 2020
Spatially resolved soft materials, such as vesicles and microgels, have shown promise as selective adsorbents and microscale reaction vessels. However, spatiotemporal control of aggregation can be difficult to achieve. In this study, nickel(II) chloride and a dipyridyl oligo(urea) ligand were combined in a vapour-diffusion setup to produce a localised spheroidal aggregate at the liquid–vapour interface. This aggregate forms via the self-assembly and fusion of monodisperse colloids and grows until its weight is no longer counterbalanced by surface tension. A simple physical model reveals that this process, termed lilypad aggregation, is possible only for surface energies that favour neither bulk aggregation nor the growth of an interfacial film. These surface energies dictate the final size and shape of the aggregate and may be estimated through visual monitoring of its changing morphology. Lilypad aggregates sequester metal from the surrounding sol and can be collected manually from the surface of the liquid.
Many studies of spatially controlled aggregation focus on the formation of a gel, a viscoelastic soft material that is almost fully resistant to flow.5 For gelation to occur in a non-uniform fashion, the rate of aggregation must exceed the rate of mixing.6 Furthermore, the gel must form under non-equilibrium conditions, such as a temperature or concentration gradient.7 Low-molecular-weight gelators (LMWGs) are of particular interest since they interact dynamically via supramolecular motifs, which rapidly form or disassemble as stimuli are applied.8–10 Localised gelation is commonly achieved by functionalising an LMWG with photoswitchable groups, such as stilbenes and azobenzenes, and partially illuminating the gelator sol.11 For pH-sensitive gelation processes, spatial resolution can also be controlled by varying the concentration of a proton source, such as glucono-δ-lactone,12 irradiating selected regions of a photoacid solution13–15 or carrying out electrochemical reactions on the surface of an electrode.16 Likewise, gels may be confined to the interface of two immiscible phases17 or shaped by the gradual mixing of co-gelator solutions.18 Inducing sol–gel transitions at specific locations can provide exquisite control over mass transport processes, including molecular diffusion, bacterial motility and enzyme activity.19
Spatiotemporal control of a sol–gel transition offers access to aggregates with uniform sizes and well-defined morphologies. Rapid gelation of liquid droplets may give rise to monodisperse spheroidal microgels, which can further self-assemble into a close-packed lattice.20 Microgels are useful as microscale reaction vessels,21 adsorbents,22 crystallisation media23 and drug delivery devices,24 and their assemblies can function as structural colourants for chemical sensors and electrochromic displays.25 It is interesting to note that many such technologies possess analogues in the natural world, wherein clusters of cells give rise to similar microstructures and rheological properties. Indeed, microgels have been studied as model cells for simulating the communication and differentiation pathways of biological tissues.26,27
The self-assembly of microgels and other colloidal aggregates is often controlled using physical stimuli, such as photoirradiation or magnetic fields.28 However, such processes may also occur passively in response to surface tension and density gradients, causing the suspended particles to separate spontaneously from a liquid medium.29 For example, gels incorporating water-immiscible solvents may be dispersed into an aqueous solution, allowed to bind and absorb a dissolved species, then manually retrieved from the liquid surface. This strategy has been used to sequester a range of common pollutants, including organic dyes and heavy metal ions.30–32
Self-assembly processes have been observed at a variety of solid–liquid, liquid–liquid and liquid–vapour interfaces.33–35 Colloidal particles can pack together to form symmetrical clusters and two-dimensional lattices,36 while polymers and hydrogen-bonding oligomers are often incorporated into fibrous membranes.37 Surface aggregates may prolong the lifetime of metastable phase mixtures, such as foams, emulsions and bicontinuous fluids.38–40 In addition, one self-assembled material can template the formation of another, such that successive stages of aggregation give rise to complex hierarchical structures.41 At the macroscopic scale, spatiotemporal control is typically achieved through layering, emulsification or dropwise addition of precursor solutions.41–45 Aggregation is thus localised to the initial liquid–liquid interfaces or droplet boundaries, where critical reagents such as solvents, gelators, acids and metal ions are most readily mixed.
An alternative method for generating interfacial aggregates is to identify a liquid in which the material is insoluble and diffuse the vapour of this antisolvent into the precursor sol. Vapour diffusion is a popular strategy for crystal growth and has also been used to control bulk gel formation and colloidal self-assembly.46,47 A benefit of this approach is that a concentration gradient is established uniformly and reproducibly, without the local variations produced by manual reagent additions.48 Furthermore, introducing the antisolvent passively minimises turbulent mixing and deformation of the interface. Thus, the formation of spatially resolved materials depends solely on interparticle interactions29 and the composition of the sol, ensuring that the key physical properties of the system can be reliably assessed.
In this investigation, a vapour diffusion method was used to generate interfacial aggregates of nickel(II) chloride and pyridyl-functionalised oligo(urea)s 1 and 2 (Fig. 1).49 To the best of our knowledge, this is the first demonstration of metal sequestration through in situ aggregation at a liquid–vapour interface. Aggregates were observed exclusively at the base of the solvent meniscus and formed in a hierarchical fashion, via the self-assembly and coalescence of monodisperse colloids. Compounds 1 and 2 were chosen for this study because they are simple oligomers which, like film-forming peptides,50 feature multiple binding sites for both self-association and metal complexation. Urea groups are known to form robust, linear arrays of hydrogen bonds known as α-tape motifs, making them well suited to the growth of fibrous aggregates.51 Moreover, the self-assembly of α-tapes may be promoted or disrupted by other supramolecular processes, allowing aggregation to be triggered or reversed by strongly binding guests.52
Fig. 1 (a) Structure of oligo(urea)s 1 and 2 and (b) schematic illustration of the vapour-diffusion strategy used to generate interfacial aggregates. |
It has previously been shown that compound 1 forms solid precipitates above its solubility limit, while 2 self-assembles into gel fibres with interesting braided and helical morphologies.49 Addition of nickel(II) chloride to the oligo(urea) solutions below their solubility limits does not result in visible precipitation. However, the use of a vapour-diffusion setup causes the metal to be sequestered into a localised spheroidal aggregate. We have termed this phenomenon lilypad aggregation, noting that the materials resemble the aquatic plants in that they grow at a liquid–air interface while supported by surface tension. It is proposed that lilypad aggregation represents an unusual but highly general pathway for non-equilibrium self-assembly, which arises when neither bulk aggregates nor interfacial films are energetically favoured. Because lilypad aggregates are isolated and well-defined, they serve as convenient packages for target solutes, which can be lifted manually from the liquid surface without filtering or evaporating the parent sol.
Compounds 1 and 2 do not differ significantly in their initial lilypad aggregation behaviour. However, given that 1 can ultimately give rise larger and more clearly defined aggregates, this system was selected for further investigation. Optical micrographs reveal three main stages in the aggregation process (Fig. 2b, c and S3, ESI†). Firstly, microparticles 5–10 μm in diameter self-assemble into a membrane-like particle raft that can be lifted intact from the liquid surface.60 The hexagonal close-packed structure of the membrane is attributed to the “Cheerios effect”, whereby particles on a liquid surface are drawn together by the tension of the meniscus between them.61 In the second stage, the particles gradually fuse62 into a continuous aggregate, which adopts a spherical cap morphology to minimise surface energy. Finally, the aggregate reaches a threshold volume, Vend, of 1.3–1.8 mm3 and, being denser than the underlying sol, sinks from the interface. Due to the small quantity of aggregate generated by this process and the variability introduced by drying, we were unable to obtain reliable rheological data for the material. Nonetheless, it should be noted that the aggregate sometimes remains connected to the surface by a 2–3 mm ligament, which persists for several minutes before yielding takes place. This stalling is evidence of viscoelasticity, as extension of the metastable ligament would continue if it were not opposed by a significant increase in elastic potential energy.63
The evolving shape and size of a lilypad aggregate may be rationalised via a simple geometric model of the aggregation process (Fig. 3 and S4, ESI†). The sol is represented as a homogeneous liquid of density ρsol, with a planar interface intersecting a uniform aggregate of density ρagg > ρsol. The aggregate represents a spherical cap with radius R and a centre located a perpendicular distance fR from the sol–vapour interface, where −1 < f < 1. At each stage of growth, the observed value of f corresponds to a local minimum in the potential energy, E. The assumption of a spherical cap geometry prevents consideration of elastic deformation processes, such as the formation of a ligament at the end of aggregate growth. Nonetheless, this simple model reproduces the key features of the aggregation process and avoids some of the difficulties of classical force-based models, which often require the geometry of the interfacial phase boundaries to be precisely known.60,65
Fig. 3 Geometric model of a lilypad aggregate, from which general expressions for the pattern of growth may be derived. |
It can be shown that:
(1) |
Eqn (1) is not applicable to aggregates that are close to sinking, as changes in elastic potential energy are not accounted for. Nonetheless, the model accurately describes materials in the earlier stages of growth, where a lack of deformation away from spherical cap morphologies suggests these elastic contributions are small. Indeed, for realistic values of ρ, γsol and γvap, the value of E in eqn (1) is overwhelmingly dependent on the surface energy terms (Fig. S5, ESI†). It may be deduced that the Bond number65 of the aggregate is significantly less than unity, causing it to adopt a spherical cap morphology to minimise its surface area. Closer examination of eqn (1) reveals that lilypad aggregation is made possible by the term in γsol, as this is the only contribution that passes through a local minimum with varying f at constant V. By contrast, the term in γvap remains small and relatively fixed as V increases. Aggregate-vapour interactions are therefore energetically important only in the initial stages of aggregate growth.
The lilypad aggregate becomes localised at the base of the liquid meniscus to minimise the gravitational potential energy, given by the first term in eqn (1). For each value of V, the material adopts a morphology with the most favourable balance of aggregate-sol and aggregate-vapour surface energies. This optimal value of f is determined by minimising E at constant V:
(2) |
Rearranging eqn (2) reveals a linear relationship between γsol and γvap:
(3) |
Physical values of V and f are only possible if γsol > 0 and |γvap| < γsol. In a plot of γsol against γvap, a combination of γsol and γvap is allowed if it lies in the quadrant bounded by γvap = γsol and γvap = −γsol and bisected by the positive x-axis (Fig. S6, ESI†). This observation offers a partial explanation for the rarity of lilypad aggregates: though the aggregate-vapour surface energy is only weakly constrained, interactions between the aggregate and sol must be unfavourable. Consequently, in addition to reducing the solubility of the precursors, an antisolvent may promote lilypad aggregation by destabilising aggregates within the bulk of the solvent.
Although the above results do not account for elasticity or local variations in ρ, γsol and γvap, they nonetheless provide realistic estimates for the physical properties of a lilypad aggregate. Eqn (3) may be used to predict the growth of a material with known surface characteristics or deduce the values of γsol and γvap from the aggregate geometry. In a plot of γvap against γsol, straight lines intersect combinations of surface energies which, for a given value of f, produce the same value of V (Fig. 4a). The gradient of the line connecting the origin and the point (γsol, γvap) is given by the ratio γvap/γsol and equal to the value of f at the beginning of aggregate growth. Likewise, the intercept of the line with gradient f = 1 that passes through (γsol, γvap) may be used to calculate V when the material is fully spherical. If γvap ≥ 0, this volume defines the maximum size the aggregate can reach before sinking. For aggregates with γvap < 0, however, there is a critical value of f, fend, beyond which the energy of the system, E, exhibits no local minimum (Fig. S7 and S8,† ESI). These aggregates become unstable and sink before attaining a spherical geometry.
In the case γvap < 0, an expression for fend is obtained by maximising V with respect to f:
(4) |
The maximum aggregate volume, Vend, may be calculated by substituting this expression into eqn (3). Coordinates of constant fend are specified by straight lines in the surface energy plot, while Vend and the corresponding radii Rend form curved contours (Fig. S9, ESI†). Marking the intersection of these lines may allow γsol and γvap to be estimated in situations where only the final aggregate geometry can be accurately observed. It should be noted that aggregates approaching fend may display significant extension under gravity, necessitating the consideration of elastic potential energy and prolate spheroid geometries. At the cost of increasing the complexity of the model, such corrections could aid our understanding of ligament formation and enable a more predictive description of aggregates in the final stages of growth.
To obtain absolute values for γvap and γsol, additional measurements are required. The value of ρ was estimated after eight hours of growth by weighing the aggregate and a 0.100 cm3 sample of sol from close to the interface. Averaging the results of three experiments reveals that ρsol = 0.82 ± 0.02 g cm−3 and ρagg = 1.00 ± 0.05 g cm−3, meaning that ρ = ρagg − ρsol = 0.18 ± 0.05 g cm−3. The error of approximately 30% in ρ affects the intercept in eqn (3) to produce errors of equal magnitude in the estimated surface energies. The irregular shape of the aggregate in the early stages of growth adds to these uncertainties by limiting the precision with which f and R can be monitored. Nonetheless, the intersection of three observations in a surface energy plot yields realistic surface energies of γsol = 0.25 ± 0.11 mJ m−2 and γvap = −0.20 ± 0.07 mJ m−2. Although the errors in these results are large, the signs of the surface energies and their orders of magnitude may be reliably determined.
It should be noted that γvap and γsol correspond to the changes in surface energy when solution at the interface is displaced by the lilypad aggregate. The measured energies are therefore much smaller than values typically reported for liquid–vapour and solid–liquid interfaces (15–30 mJ m−2 for organic solvents under air66 and 5–50 mJ m−2 for water on organic polymers67), which denote the absolute energies for surface formation. Indeed, comparable surface energies below 1.0 mJ m−2 have been measured at the interfaces of liquid phases.68 Interfacial aggregation occurs because aggregate-sol interactions are slightly weaker than interactions between the solvent molecules themselves (γsol > 0), whereas replacing solvent–vapour interactions with an aggregate-vapour interface is energetically favourable (γvap < 0).
Lilypad aggregation is an example of a Rayleigh–Taylor instability, in which layers of immiscible fluids interpenetrate due to the greater density of the upper phase.69 As the radius of a lilypad aggregate approaches Rend, sinking may occur spontaneously and unpredictably in response to mechanical disturbances or other minor physical inputs. For systems in which γvap < 0, this observation may be rationalised by the presence of a local maximum in E at a value of f > fend. The energy needed to deform beyond this value of f, Ea, decreases in magnitude as the aggregate grows larger. Thus, lilypad aggregates with f > 0 and γvap < 0 are metastable, meaning that they may relax into a spherical geometry and detach from the interface. For a typical aggregate in this study, Ea is approximately 50 pJ at f = −0.2, but decreases to 12 pJ at f = 0.0 and just 0.6 pJ at f = 0.2 (Fig. S12, ESI†). The sharp reduction in Ea on approaching fend = 0.50 ± 0.07 greatly increases the probability of sinking, as even small perturbations can provide the energy required.
Analysing the surface energies of a lilypad aggregate provides insight into the factors underlying this unusual physical phenomenon. The values of γsol and γvap are likely to depend on the ligand structure and aggregation conditions. It is possible that modifying these parameters will enable rational tuning of the aggregate process, to generate larger aggregates incorporating a higher percentage of the dissolved metal salt. A key aim of future investigations will be to use surface energy measurements to rationalise the design of lilypad aggregate systems and target more efficient and versatile metal-sequestration agents.
Interestingly, if DMF is replaced with dimethylacetamide or a different metal halide or antisolvent is used, lilypad aggregation does not occur (Table S1 and Fig. S13, ESI†). However, lilypad aggregates are formed if the concentration of 1 is reduced, albeit more slowly and with smaller final volumes (Fig. S14, ESI†). It may be concluded that the experimental outcome is highly sensitive to the choice of metal and solvent system but relatively weakly affected by changes in ligand concentration. To optimise the process for practical applications, it would be useful to measure the effect of these and other physical parameters, such as the temperature and container geometry, on the rate of aggregation.48 Obtaining such kinetic data will be a key objective of future investigations.
In the absence of metal ions, the oligo(urea)s used in this study display very different aggregation behaviours. Tris(urea) 1 gives rise to microcrystalline precipitates in DMF while pentakis(urea) 2 forms braided networks of helical gel fibres.49 In both cases, however, diffraction and modelling studies suggest that the aggregates consist of molecules in extended conformations. By contrast, many other linear oligo(urea)s have been shown to form foldamer structures through intramolecular hydrogen bonding70 or anion complexation.71 It is proposed that 1 and 2 do not exhibit interactions of this type due to the relative rigidity of the p-benzyl spacers between urea groups. When the compounds are incorporated into labile72 nickel(II) complexes, conformational strain could similarly favour coordination polymers over discrete chelate structures, generating a flexible cross-linked network as the thermodynamic product.
To assess the potential for folded conformations, we optimised the geometry of compound 1 in Gaussian 16 with the DFT functional B3LYP, using the 6-31+G* basis set for initial calculations and a range of larger basis sets for subsequent refinements. The tris(urea) was modelled both as an extended molecule and in a C-shaped geometry stabilised by intramolecular hydrogen bonds (Fig. S15, ESI†). The strength of this motif was calculated by subtracting the energy of the folded tris(urea) from that of the extended molecule. For comparison, we also calculated the interaction energy of mono(urea) 3, which can form intermolecular hydrogen bonds with little conformational strain (Fig. 5a). Finally, the activation energy for intramolecular hydrogen bonding was estimated by incrementally separating the terminal urea groups of the folded conformation of 1 and optimising the remainder of the structure after each scan step (Fig. 5b).
The results of our DFT analysis indicate that the backbone of 1 is moderately flexible (Table S2, ESI†). For urea–urea distances greater than 7.5 Å, the molecule can adopt a range of conformations with similar energies. However, intramolecular hydrogen bonding requires the molecule to access less favourable geometries, producing an activation barrier of approximately 6 kJ mol−1 (Fig. S16, ESI†). This conformational strain weakens the intramolecular hydrogen bond by 19–20 kJ mol−1 (58–62%) relative to the dimer of 3. The stability of the folded geometry may be further reduced by the loss of conformational freedom, which is not accounted for in DFT calculations.73 Indeed, studies of ring closure reactions suggest that the six fully rotatable bonds between terminal urea groups are significantly constrained, resulting in a total entropic cost of approximately 7 kJ mol−1 at 298 K.74
Despite the impact of conformational strain, folding of the tris(urea) molecule significantly lowers the energy of the system. Intramolecular hydrogen bonds may also be kinetically favoured due to the high effective molarities of the interacting groups.74 However, the greater stabilities of motifs involving extended tris(urea) molecules, represented by the hydrogen bonded dimer of mono(urea) 3, likely leads to a preference for these structures at thermodynamic equilibrium. The p-benzyl linkages of 1 and 2 may promote the self-assembly of soft materials because they are compatible with supramolecular polymerisation but sufficiently flexible to inhibit crystallisation pathways. More versatile lilypad aggregation processes could be developed by incorporating similar spacers into alternative bidentate ligands, using metrics such as interaction plots to achieve optimal metal binding in aqueous solvent systems.75–77
To assess the efficiency of metal sequestration, lilypad aggregates were prepared from DMF solutions of 1 (0.50% w/v, 9.8 mM) and nickel(II) chloride hexahydrate (0.24% w/v, 1 eq.) and analysed by a variety of techniques. The FT-IR spectrum of the wet aggregate indicates a high concentration of 1, but the characteristic signals of this species differ significantly from those of the pure compound (Fig. S17, ESI†). In particular, shifting of the amide I band from 1654 cm−1 to 1710 cm−1 suggests that the urea carbonyl interacts more weakly in the aggregate,79 and aggregation is likely driven by the formation of a nickel(II) complex. X-ray photoelectron spectroscopy (XPS) measurements on the surface of the vacuum-dried aggregate support this conclusion, as the presence of shake-up peaks in the Ni 2p region are diagnostic of a paramagnetic octahedral complex (Fig. 6 and S18 and S19,† ESI).54 Scanning electron microscopy (SEM) images of the coated aggregate reveal a lack of discernible microstructure at the surface of the material. However, fibrous structures are visible within desiccation cracks (Fig. 7a and S20, ESI†), suggesting that the nickel(II) complexes give rise to polymeric assemblies that densely entangle to form a homogeneous material.
The composition of the lilypad aggregate was analysed by inductively coupled plasma mass spectrometry (ICP-MS). After drying under vacuum, the material displayed a nickel content of 5.4 ± 0.1% w/w (Table S3, ESI†). This value closely matches the expected value of 5.1% w/w for a 1:2 nickel(II)–tris(urea) complex. It is concluded that the tris(urea) accounts for most of the organic component of the aggregate, and occupies four of the six coordination sites of the octahedral nickel(II) complexes.
Further compositional information was obtained through thermogravimetric analysis (TGA) of the wet, as-synthesised material under a nitrogen atmosphere (Fig. 7b). An initial mass loss of 41% below 150 °C is attributed to the removal of weakly bound solvent, while a subsequent decrease of 47% between 150 and 900 °C corresponds to decomposition of the metal complexes. The measured solvent content is at least one order of magnitude lower than in comparable soft materials, such as supramolecular gels, and consistent with the high aggregate density observed at the surface of the material. SEM images of the black pyrolysate powder display a uniform arrangement of pores, 100–300 nm in diameter (Fig. 7c and S21, ESI†), on the particle surfaces. It may be concluded that the organic solvent occupies narrow voids and is homogeneously distributed in the original material.
After a lilypad aggregate is pyrolysed, its metal content may be usefully recovered from the solid residue. Powder X-ray diffraction (PXRD) analysis of a pyrolysate sample reveals peaks at 2θ = 44.5 and 51.9°, matching the (111) and (200) reflections of elemental nickel (Fig. S22, ESI†).80 However, studies of pyrolysed elastomer-metal halide mixtures have shown that carbonaceous material often accounts for 20–40% of the residue mass.81 ICP-MS experiments reveal a total nickel concentration of 35.5 ± 0.3% w/w in the pyrolysed aggregate, suggesting that the material incorporates just 6% of the nickel in the parent sol (Table S3, ESI†). Nonetheless, formation of the aggregate may aid recovery of the metal, as the material exceeds the nickel concentration of the original DMF solution by 40–80 times (Table S4, ESI†) and may be harvested manually from the surface of the liquid. Future studies will focus on improving the efficiency and versatility of this metal sequestration and modifying the properties of aqueous film-forming ligands, such as peptides,50 to target similar lilypad aggregation pathways.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d0sc02190c. The underlying research data for this paper is available in accordance with RCUK open data policy from DOI: 10.15128/r1ms35t8650. |
This journal is © The Royal Society of Chemistry 2020 |