Andrew W.
Knight
a,
Poorandokht
Ilani-Kashkouli
b,
Jacob A.
Harvey
c,
Jeffery A.
Greathouse
c,
Tuan A.
Ho
c,
Nadine
Kabengi
bd and
Anastasia G.
Ilgen
*c
aStorage and Transportation Technology Department, Sandia National Laboratories, Albuquerque, NM 87185-0754, USA
bDepartment of Geosciences, Georgia State University, Atlanta, 30303 Georgia
cGeochemistry Department, Sandia National Laboratories, Albuquerque, NM 87185-0754, USA. E-mail: agilgen@sandia.gov; Tel: +1 (505) 284 1393
dDepartment of Chemistry, Georgia State University, Atlanta, 30303 Georgia
First published on 31st October 2019
Spatial confinement is prevalent in sedimentary rocks and can lead to changes in the chemical behavior at mineral–water interfaces. This includes both deviations in the physico-chemical properties of confined water, when compared to the bulk liquid phase, and subsequent alterations in adsorption chemistry of ions inside nano-scale mineral pores. Here we document contrasting adsorption mechanisms and differences in local coordination environments of copper (Cu2+), depending on whether the ion is adsorbed on non-porous silica surfaces, versus inside 8 nm and 4 nm pores in silica. X-ray absorption spectroscopy, flow micro-calorimetry, and batch adsorption methods together with molecular modeling are used to thoroughly describe the dependence of these adsorption processes on the pore size. We show that confinement within silica pores promotes Cu–Cu interactions and increases the formation of Cu–Cu polynuclear surface complexes. We also demonstrate that the mechanism of Cu2+ adsorption on non-porous versus porous silica is vastly different. The adsorption of Cu2+ on non-porous silica is an endothermic process, whereby Cu2+ undergoes dehydration prior to surface complexation. In contrast, adsorption within nano-scale pores is preceded by only partial dehydration, and significant formation of Cu–Cu polynuclear complexes, which leads to an overall exothermic signal. Interfacial confinement leads to dramatic changes in the adsorption mechanism and speciation, which in turn controls the fate and transport of chemical species in natural environments.
Environmental significanceReactions at mineral–water interfaces control the fate and transport of elements in aqueous and soil environments, as well as in sedimentary rocks. In tightly compacted soils and rocks, mineral–water interfaces are spatially confined, with the confined reactive domains having nm-scale dimensions. The reactivity of mineral–water interfaces is affected by this spatial confinement. Namely, the dielectric constant and density of water decrease when the domain size approaches less than 5 nm. These changes in the properties of solvent (water) lead to the changes in hydration and coordination chemistry of ions, and, as a result, changes in the complexation of ions with the mineral surfaces. We discovered that adsorption of copper Cu2+ aqueous species onto silica surface depends on spatial confinement. Here we show that adsorption mechanisms of Cu2+ onto confined (8 nm and 4 nm pores) vs. non-confined silica surfaces are vastly different. The confinement within silica pores increases the formation of Cu–Cu polynuclear surface complexes, and changes the sign of the overall heat of adsorption. Identifying the reactivity changes and chemical mechanisms active under nano-scale confinement further expands our understanding of interfacial chemistry in environmentally-relevant systems. |
Nano-scale pores are ubiquitous in nature and have large surface area to volume ratios, thus the physical and chemical processes occurring within this reactive porous media can control the macroscopic behavior of water and ions, including solute hydration, chemical and physical sorption processes, stable isotope fractionation, and nucleation.3,12,13 Currently, a fundamental understanding of nano-scale confinement effects on interfacial processes is inadequate, and therefore cannot be incorporated into continuum models for predicting macroscopic chemical behavior. Recent studies have used systematic approaches to quantify chemical changes in incrementally smaller pores.4,6,9 These studies are critical for the development of fundamental frameworks for predicting geochemical reactions in the subsurface.
Nano-scale confinement induces shifts in water's physical properties, which drives enhanced cation adsorption and electron transfer reactions on mineral surfaces.5–9,14–16 Nano-scale confinement alters intermolecular water–water interactions, leading to changes in the dielectric constant, density, viscosity, and surface tension of water in pores with diameter less than 5 nm.8,14,15 Driven by changes in water properties, cation hydration energies decrease, shifting the equilibrium of the solvation and dehydration reactions, the structure and influence of the electrical double layer, and changes in the interfacial energies between the surface and solution.3,4,9 Previously, we demonstrated an increased adsorption maximum and faster reaction kinetics when Cu2+ is adsorbed on mesoporous silica (SBA-15) with a 4 nm pore diameter, compared to SBA-15 materials with 6 and 8 nm pore diameters.4 Similarly, Kim et al. reported enhanced adsorption of arsenic (As) on ordered mesoporous alumina when compared to its non-porous counterpart.17 Additionally, Mantha et al. 2019 demonstrated that Cu-bearing nano-particles can form in confined pore spaces present in organic matter of soils, and that these Cu-bearing nano-particles play an important role in the transport of metals within the soil column.2
Congruently, calorimetry studies presented thermodynamic evidence for enhanced inner-sphere adsorption inside nano-scale pores.18 Sodium, for example, forms inner-sphere adsorption complexes when confined inside zeolite pores (0.51 nm pore diameter), while mostly forming outer-sphere complexes on unconfined surfaces.19 Anomalous adsorption behavior was observed for sodium, nickel, potassium and calcium in nano-scale pores of zeolite.18,20 These studies reported that adsorption of cations is enhanced in incrementally smaller pores in zeolite (all examined pores were <1 nm diameter). The proposed explanation for the enhanced adsorption is that in the pores between 0.3–0.5 nm (where the hydrated ion diameter is larger than the pore diameter) ions have to dehydrate to enter the pore, and these dehydrated ions form inner-sphere complexes.18,20
While efforts have been made to understand thermodynamic changes of nano-scale confined systems, few datasets quantify the microstructure (i.e. local coordination environment) around ions adsorbed on confined surfaces. Recently, Nelson et al. investigated coordination environment of zinc (Zn2+) confined inside controlled pore glasses (CPGs), where Zn2+ adsorbed predominantly as a tetrahedral complex inside 10 nm pores (the smallest pore evaluated), while inside larger pores (up to 330 nm pore diameters) the coordination environment varied between tetrahedral and octahedral.9 Their study provides valuable analyses to improve our understanding of Zn2+ adsorption on siliceous materials, however it is critical to interrogate the confinement effects in pores with <10 nm diameter, since significant changes in the solvent (water) properties were only observed for spatial confinement of 10 nm and less. While few studies indicate confinement effects inside pores that are >10 nm in diameter,9,21 the research overwhelmingly suggests that these effects are most pronounced in pores that are <10 nm.5,22–24 Therefore, it is imperative to investigate the coordination chemistry of adsorbed ions confined within materials with pore sizes less than 10 nm, as these nano-scale domains are ubiquitous in the environment.3,4,11,12
In this study we address the scientific gap in understanding changes in the coordination environment and reaction pathway for Cu2+ by using X-ray absorption fine structure spectroscopy (XAFS), flow micro-calorimetry, and molecular dynamics (MD) simulations to decipher the systematic changes as a function of decreasing pore size of adsorbing silica. We interrogate the impact of nano-scale confinement on the coordination chemistry and heat of adsorption during Cu2+ partitioning onto non-porous silica and within nano-scale pores in mesoporous (SBA-15) silica with 8 and 4 nm pores. The sum of our findings provides a complete understanding of the pathways, products and thermodynamics of Cu2+ adsorption reaction onto silica surface under nano-scale confinement.
The surface area was quantified using a Brunauer–Emmett–Teller (BET) surface area analyzer, as reported previously.4 Briefly, approximately 200 mg of dried fumed silica was transferred to a tarred BET tube equipped with an airtight cap and loaded onto an ASAP 2020 to be degassed for 4 hours at 300 °C and backfilled with inert He gas. Following sample degas, any mass change resulting from the degas step was noted and adjusted. To start the analysis, a dewar was filled with liquid N2 and the BET tube was equipped with a thermal jacket and placed in the sample holder. The BET surface area and the non-local density functional theory (NLDFT) method was used to determine the average pore diameter and pore volume of fumed silica.26,27
XAFS data were collected at the undulator beamline sector 20-ID (XOR) at the Advanced Photon Source (APS) at Argonne National Laboratory. As described previously, the uncollimated beam was used with a Si(111) monochromator that was detuned by 15% for harmonic rejection.13 The monochromator was tuned to the Cu K-edge at approximately 9200 eV via a Cu metal foil, similar to Cheah et al.28,29 The monochromator step size was 10 eV in the pre-edge, with a counting time of 0.5 seconds per point, 0.3 eV step size in the near edge region with 0.5 second counting time, and 0.05 eV in the XAFS region with a counting time of 0.5 seconds per point. The fluorescence points for Cu were collected on a Vortex Si Drift solid state 4 element detector. During the analysis, the sample chamber was kept at 22 K using a Displex liquid helium cryostat to avoid beam damage. For each sample, a minimum of 3 XAFS scans were collected to achieve an appropriate signal to noise ratio. The Cu standards used to determine the amplitude reduction factor (S0) were copper(II) acetate, copper(II) hydroxide, copper(II) oxide, and copper(I) oxide. The S0 value was determined to be 0.89 by fitting Cu XAFS standards. This S0 value was used for fitting all Cu XAFS data collected during this analytical run.
The Cu K-edge XAFS data was processed and analyzed using Athena and Artemis30 interface to the IFEFFIT software.31 The background subtraction, normalization, and conversion into k-space using Athena are described in detail elsewhere.31,32 A Fourier transform (Hanning Window, dk = 2) was applied to Cu K-edge XAFS data over the k-range from approximately 3–12 (Table 1). The Fourier transformed data was imported to Artemis to be analyzed by IFEFFIT30 with a shell-by-shell fitting of theoretical single scattering (SS) paths approach.33 For analysis and calculation of theoretical phase and amplitude functions for Cu–O, Cu–Cu, and Cu–Si, paths were generated using FEFF6 code33 from the crystal structure of dioptase reported in Ribbe et al.34
Sample | Surface loading (μmol m−2) | k-Range | R-Range (Å) | Shell | CN | R (Å) | σ 2 (Å2) | ΔE0 eV | R-Factor | Red χ2 | Ind. Pts. |
---|---|---|---|---|---|---|---|---|---|---|---|
Notes:a Usable k-range.b Coordination number.c Bond length.d Debye–Waller factors: mean-square amplitude reduction factor, including thermal and static disorder components.e Energy shift between the theoretical and measured spectrum.f R-Factor (mean square misfit) .g Reduced chi-square h Independent points (number of data points minus number of variable parameters) Nidp = Npts − Nvar. | |||||||||||
Fumed silica-hCu | 0.18 (3) | 3.0–11.83 | 1.0–5.0 | Cu–Oeq | 4.3 (3) | 1.94 (7) | 0.007 | 2.6 (9) | 0.020 | 31.40 | 22.05 |
Cu–Oax | 0.6 (1) | 2.29 (1) | 0.002 | ||||||||
Cu–Cu | 0.9 (6) | 2.94 (2) | 0.009 | ||||||||
Cu–Si | 0.1 (3) | 3.52 (3) | 0.009 | ||||||||
SBA-15-8-hCu | 0.09 (1) | 2.6–10.7 | 1.0–5.0 | Cu–Oeq | 3.7 (7) | 1.94 (6) | 0.003 | −1.0 (7) | 0.026 | 48.07 | 22.02 |
Cu–Oax | 1.2 (6) | 2.84 (3) | 0.007 | ||||||||
Cu–Cu | 1.7 (2) | 2.96 (6) | 0.005 | ||||||||
Cu–Si | 0.3 (3) | 2.99 (6) | 0.008 | ||||||||
SBA-15-8-mCu | 0.06 (1) | 3.0–11.85 | 1.0–5.0 | Cu–Oeq | 3.4 (3) | 1.94 (7) | 0.002 | −1 (1) | 0.021 | 28.60 | 22.05 |
Cu–Oax | 0.8 (4) | 2.82 (2) | 0.005 | ||||||||
Cu–Cu | 1.7 (2) | 2.96 (7) | 0.008 | ||||||||
Cu–Si | 0.2 (3) | 3.5 (1) | 0.004 | ||||||||
SBA-15-4-hCu | 0.09 (2) | 3.0–11.86 | 1.0–5.0 | Cu–Oeq | 3.9 (2) | 1.94 (2) | 0.003 | 1 (3) | 0.009 | 34.14 | 22.04 |
Cu–Oax | 0.7 (5) | 2.90 (3) | 0.003 | ||||||||
Cu–Cu | 1.6 (5) | 2.96 (9) | 0.008 | ||||||||
Cu–Si | 0.9 (2) | 3.03 (2) | 0.009 | ||||||||
SBA-15-4-mCu | 0.06 (1) | 3.0–11.84 | 1.0–5.0 | Cu–Oeq | 3.5 (5) | 1.95 (1) | 0.002 | 1.8 (9) | 0.024 | 59.04 | 22.05 |
Cu–Oax | 1.2 (5) | 2.86 (2) | 0.005 | ||||||||
Cu–Cu | 1.9 (4) | 2.97 (5) | 0.009 | ||||||||
Cu–Si | 0.6 (4) | 3.53 (7) | 0.003 |
Material | BET surface area (m2 g−1) | Mesopore diameter (nm) | OH− density (OH−/nm2) | Mass (mg) | Normalized energy (mJ m−2) | Normalized adsorption (kJ mol−1) | Normalized adsorption (μmol m−2) |
---|---|---|---|---|---|---|---|
a This study. b From Knight et al. (2018).4 | |||||||
Fumed silica | 192 ± 7a | N/A | 2.15a | 20 ± 0.5 | 0.429 | 1.165 | 0.369 |
SBA-15-8 | 661 ± 5b | 7.0 ± 0.3b | 1.8 ± 0.2b | 20 ± 0.5 | −0.128 | −1.349 | 0.095 |
SBA-15-4 | 580 ± 13b | 4.4 ± 0.1b | 2.3 ± 0.2b | 20 ± 0.5 | −0.253 | −2.168 | 0.117 |
The heats of reactions (Qads in mJ m−2) and molar enthalpies (ΔH in kJ mol−1) were calculated by integrating the calorimetric peaks and were converted to energy units (Joules) by calibration with calorimetric peaks of known energy inputs generated from a calibrating resistor. The solution pH was adjusted daily using dropwise addition of 0.1 M HNO3 and 0.1 M NH4OH to achieve a final pH of 6.0 ± 0.1. Changes in total concentration and ionic strength resulting from pH adjustments were determined to be negligible.
(1) |
Details for the construction of the silica surface model, composition of pore fluids, and pre-equilibrium simulation methodology are described in detail in the ESI.† MD trajectories for analysis were generated from a 10 ns production simulation. Number densities were calculated by determining the distance from each Cu atom to the nearest pore oxygen divided by the available volume at that distance determined from a Monte-Carlo insertion code. Insertions were accepted if the distance from the attempted insert to any other atom was greater than 65% of the interatomic LJ sigma value. The 0.65 factor was chosen to accept all distances that were observed in the MD simulations.
Previous XAFS data analyses indicates that Cu2+ experiences strong Jahn–Teller effects due to its d9 election configuration, and therefore forms a distorted (elongated) octahedron; where the Cu–Oeq (bonds with four equatorial oxygen neighbors) are shorter than the Cu–Oax bonds (axial direction).28,29,60–63 Current experimental evidence is inconclusive and has led to a debate regarding the distorted octahedron as the preferred aqueous species: four-, five-, and six-coordinate complexes have been hypothesized,62 and the aqueous complex is likely to be a dynamic distribution of five- and six-coordinated complexes.60–62 For our XAFS analysis we are aware of this challenge and have maintained a consistent and rigorous fitting protocol in an attempt to isolate differences between samples, rather than proposing a specific coordination geometry. Therefore, in our model we simply considered the lowest energy conformation, with a coordination number (CN) of 6, consisting of four equidistance equatorial O atoms and two elongated axial O atoms. The second shell in our model consisted of Cu–Si and Cu–Cu backscattering paths. The average CN and bond distance (R) for each coordination shell are summarized in Table 1.
The XAFS data of adsorbed Cu2+ was fit using the paths for a distorted octahedral coordination environment, as described in the previous XAFS studies and simulations.28,29,61,63 The average measured CN for Cu–Oeq was ∼ 3.76 ± 0.36 (Table 1); while the CN for Cu–Oax was ∼0.9 ± 0.28 for non-porous and porous silica samples. Based upon these CN numbers, the local coordination around Cu is likely to contain 5 or 6 oxygen atoms. Likewise, the bond distances for Cu–Oeq were determined to be nearly identical for all silica samples (Table 1). The Cu–Oax bond distances are similar for all porous silica samples (2.82 ± 0.02–2.90 ± 0.03 Å); however, the Cu–Oax distance for non-porous silica is significantly shorter (2.29 ± 0.01 Å). The Cu–O distances for non-porous silica agreed well with the model and other previously reported values for a distorted octahedral complex, where Cu–Oeq = 1.94–1.96 Å, and Cu–Oax = 2.15–2.65 Å.34,61–63 However, not all previous research fit the Cu–Oax shell due to either large σ2 values indicating that no reasonable CN or bond distance could be obtained or the data were fit with a different coordination environment.28,29,62 Our σ2 for Cu–O shells did not exceed 0.008, and therefore the bond distances are quantified accurately for all samples, with an error of 0.1 Å.
The second coordination shell involves interactions between Cu with nearest Cu atoms (i.e., Cu–O–Cu) as well as coordination with Si (i.e., Cu–O–Si) at the surface. We observed the presence of Cu–Cu backscattering in all porous and non-porous silica samples. This observation agrees with previous studies, which report Cu2+ forming dimeric or polymeric complexes on the surface of non-porous silica, while maintaining a monomeric coordination when adsorbed on alumina.28,29 However, in our work we observed more dimeric complexes in porous silica, compared to non-porous silica, as evidenced by the amplitude of the Cu–Cu backscattering feature. When Cu2+ was adsorbed to non-porous silica, the CNCu–Cu = 0.9 ± 0.6, and RCu–Cu = 2.94 ± 0.02 Å (Table 1). Whereas the CNCu–Cu for the porous SBA-15 silicas was 1.7 ± 0.2 for both high- and medium-surface coverage inside the 8 nm pores, and 1.6 ± 0.5 and 1.9 ± 0.4 for the high- and medium-surface coverage inside the 4 nm pore, respectively. The CNCu–Cu on the SBA-15 samples did not significantly differ between 4 nm and 8 nm pore diameters and did not depend on the surface coverage of adsorbed Cu2+. The intensity of the Cu–Cu backscattering feature for SBA-15 samples was significantly higher than for non-porous silica (Fig. 1), and further suggests that, in confinement, Cu2+ has a greater tendency to form dimeric to multimeric complexes. The determined Cu–Cu distances in all samples were approximately the same in the mesoporous silica samples, where RCu–Cu ranged from 2.96–2.97 Å. These RCu–Cu and CNCu–Cu values correspond to the formation of a Cu2(OH)2-like adsorption complex on the surface (where one Cu is in inner-sphere surface coordination), as the CNCu–Cu was determined to be 1.9 ± 0.2 and RCu–Cu = 2.96 ± 0.002 Å, which agrees with the previous studies reporting Cu polymerization on non-porous silica.28,29 However, our complexes on mesoporous silica appear to differ from those described by Cheah et al. for amorphous silica, in which the authors observed an adsorption complex that was distinctly different from Cu(OH)2, with a lower CN and shorter bond length (CN = 0.5–0.9 and RCu–Cu = 2.58–2.59 Å).
The Cu–Si backscattering feature of adsorbed Cu2+ was fit with low (<1) CN numbers in all analyzed samples. Due to silicon's low mass, the backscattering signal from the Cu–Si shell is weaker, compared to Cu–Cu backscattering signal. Since Cu–Si and Cu–Cu shells overlap, the Cu–Si becomes nearly invisible. We fit all data including and excluding Cu–Si shell, and a statistically significant improvement of the fit was only seen for the SBA-15-4-hCu sample (ESI†). Because the XAFS signal is an average of all the Cu2+ species in the system, we conclude that copper forms both inner- and outer-sphere complexes on silica surface. Additionally, the presence of Cu-oxide nano-particles cannot be ruled out for the samples containing oligomeric Cu2+ species within confined silica pores. Based on the qualitative trends in the Cu–Si CNs, there may be a greater probability of Cu2+ to form inner-sphere complexes with Si in smaller pores, and that proportion of these complexes may decrease as the pore sizes increases. The Cu–Si bond length for the porous silicas with surface loadings of approximately 0.05 μmol m−2 were about 3.5 Å, while for the silicas with a higher Cu2+ surface loading, approximately 0.1 μmol m−2 for porous silica and 0.2 μmol m−2 for nonporous silica, the Cu–Si distance was determined to be approximately 3.0–3.5 Å (Table 1, Fig. 2).
Fig. 2 The distribution of average bond distances and coordination numbers of each shell for fumed silica, SBA-15-8, and SBA-15-4 for high Cu2+ surface loading. |
This effect of confinement on Cu2+ adsorption was explored further by examining the results of MD simulations of Cu-hydroxide complexes in 2, 4, and 8 nm amorphous silica slit pores (Fig. 3, see ESI† for full computational details). Cu2+ is least likely to be found at the pore wall in the 8 nm pore and that probability increases monotonically to 2 nm as evidenced by the increasing intensity of the <2 Å peak (Fig. 3a). By integrating the area under the curve in this region we find that 69% of the Cu2+ in the simulation is found at the pore wall for the 2 nm pore, compared to just 41% in the 8 nm pore, indicating increased adsorption with increasing degree of spatial confinement. Moreover, the CNCu–Si, calculated from the simulated radial distribution functions (RDFs), increases from 0.41 to 0.69 going from 8 nm to 2 nm pores, qualitatively agreeing with the experimentally observed trend of enhanced Cu–Si backscattering feature in XAFS (Table 3, Fig. 3).
Sample | Shell | CN | R (Å) |
---|---|---|---|
8 nm pore | Cu–Odep | 0.42 | 1.75 |
Cu–O | 1.88 | 1.78 | |
Cu–Ow | 2.74 | 1.98 | |
Cu–Cu | 0.59 | 2.65 | |
Cu–Si | 0.41 | 3.15 | |
4 nm pore | Cu–Odep | 0.58 | 1.75 |
Cu–O | 1.17 | 1.75 | |
Cu–Ow | 3.36 | 1.98 | |
Cu–Cu | 0.26 | 2.68 | |
Cu–Si | 0.58 | 3.15 | |
2 nm pore | Cu–Odep | 0.69 | 1.78 |
Cu–O | 1.05 | 1.78 | |
Cu–Ow | 3.58 | 1.98 | |
Cu–Cu | 0.35 | 2.65 | |
Cu–Si | 0.69 | 3.12 |
Fig. 4 Comparison of calorimetric heats measured during Cu2+ flow through experiment (adsorption) for fumed silica, SBA-15-8, and SBA-15-4 versus time. |
Based upon our measurements, the average Cu2+ coordination on the surface could either be considered a distorted octahedron or distorted square pyramid. In either case, the average local coordination environment of Cu2+ adsorbed onto the porous SBA-15 samples is more distorted compared to non-porous silica, as evidenced by the Cu–Oax and Cu–Oeq bond distances. Longer Cu–Oax bond distances, less similar to Cu–Oeq distances, represent a more disordered complex, where the Cu–Oax were 0.5–0.6 Å longer when confined within pores than adsorbed to non-porous silica. Additionally, within the confined porous SBA-15 samples, Cu–Oax distances for 4 nm pores were longer than those for 8 nm pores, suggesting that as confinement increases, distortion of the first coordination shell of Cu2+ also increases. Additionally, this may suggest that Cu–Oax elongation occurs in confinement due to Cu–O bond weakening as a result of increased oligomerization. We note that calculated Cu–Owater bonds (1.98 Å) are longer than both the Cu–OH and Cu–O− bonds (1.78 Å) in the simulations, indicating that both increased adsorption and dimerization lead to some Cu–O bonds that are shorter than others. In the experiments we also observed that these effects are sensitive to the density of adsorbed Cu2+ on the surface. While the overall surface coverage remains low (i.e. most OH available on the SiO2 surface are free of Cu2+), the Cu–Oax distances became longer with increased surface density of adsorbed Cu2+. With increasing surface coverage of Cu2+ on the surfaces in silica pores, the effects of spatial confinement become more pronounced. In both mesoporous silicas with 4 nm and 8 nm pores, increased surface Cu2+ coverage led to an increase in the average Cu–Oax distance of 0.2–0.4 Å. Therefore, we observe that the least distorted Cu2+ complexes exist on non-porous silica, and as surface coverage increases inside the pores of mesoporous silicas, disorder increases. Unlike some studies, where changes in the CN were observed,9 we did not see any significant changes in CNCu–O in the XAFS data, at least within the uncertainty of ∼10–20% typical for CNs derived from XAFS data. The subtle differences between Cu–O distances suggest that nano-scale confinement distorts the coordination geometry of Cu2+. This phenomenon could be a result of steric effects of a hydrated Cu2+ complex entering an incrementally smaller pore. The ion–water interactions inside nano-scale pore are weaker, compared to the non-confined phase, due to water's dielectric constant decreasing with increasing degree of confinement8,14 and, therefore, the hydration sphere of Cu2+ may become distorted inside the pores.
In addition to the elongation of the Cu–Oax bonds, confinement increased the propensity for Cu2+ to undergo dimerization or oligomerization. While bond distances in all samples remained nearly the same, the amplitude of Cu–Cu backscattering (CNCu–Cu) was nearly double for mesoporous samples, compared to non-porous silica. As simulation snapshots demonstrate, Cu–Cu species all exist in a distribution, however when confined, the average CNCu–Cu is greater than in the non-confined system. This finding is confirmed by preliminary MD simulation results on a non-porous silica surface where a vacuum gap is applied above the liquid layer (data not shown). In this system no Cu2+ dimers are observed, in stark contrast to the results observed in the confined simulations. Previous studies investigating Cu2+ adsorption on amorphous silica demonstrated that Cu2+ forms dimers on silica surfaces.28 Our findings on non-porous silica confirm these earlier reports, while we also show that in confinement Cu–Cu backscattering is more pronounced. This polymerization reaction explains the difference in calorimetric trends between porous and non-porous silicas. Adsorption of Cu2+ in confinement was exothermic, while it was endothermic for non-porous silica. The endothermic response for the adsorption of Cu2+ on non-porous silica is likely a result of dehydration of the Cu2+ ion prior to adsorption.18–20 This suggests that when Cu2+ displaces NH4+ at the silica adsorption sites, the Cu2+ ions dehydrate and form inner-sphere adsorption complexes on the surface, which has been shown to be an endothermic reaction.19,64 Similar results were previously noted for Na+ and Ca+ on zeolite surfaces. When Na+ replaced Ca2+ at an adsorption site, there was a small exothermic peak followed by large endothermic peak. This process was described as follows; Na+ enters the pore, displacing Ca2+, and Ca2+ rehydrates (exothermic). Then, in order to adsorb to the surface as an inner-sphere complex, Na+ dehydrates (endothermic).19
Our XAFS and calorimetry data indicate that Cu2+ forms inner-sphere adsorption complexes on non-porous silica (CNCu–Si = 0.1 and RCu–Si = 3.52 Å). Previous work has illustrated Cu–Si complexes and provided estimated Cu–Si bond distances reflecting different surface adsorption complexes. It was determined that, Cu2+ can form a corner-sharing mono-dentate (RCu–Si = 2.6–3.5 Å), edge-sharing bi-dentate (RCu–Si = 2.2–2.4 Å), or corner sharing bi-dentate complex (RCu–Si = 3.0–3.5 Å).28,29 From these values, we cannot definitively say which complex is present in our systems, however it is unlikely that the edge-sharing bi-dentate complex is occurring as the expected bond distances are too short. The formation of an inner-sphere mono-dentate complex is consistent with other reports of Cu2+ adsorbed onto amorphous non-porous silica.28 Based upon the assumption that the Si–O bond length (∼1.61 Å)65,66 remains unchanged in confinement, we can assess the likelihood of Cu2+ adsorption through Cu–Oeq or Cu–Oax based upon estimated bond angles.
Conversely, the adsorption of Cu2+ on the SBA-15 materials responded with a weak endothermic feature, followed by a strong exothermic peak. The magnitude of both the endothermic and exothermic signals were stronger as the pore size decreased from 8 nm to 4 nm in diameter. This observed magnitude change could be a result of known changes to water properties.67 Because the H-bonding network and dielectric constant of water are altered when confined inside mesoporous silica, it is likely that the hydration energy of Cu2+ decreases with decreasing pore diameter – thus changing the magnitude of the calorimetric signal. The small endothermic peak is likely attributed to dehydration of Cu2+, similar to the adsorption process on non-porous silica. However, as the Cu2+ surface loading increases over time, and additional adsorption occurs, dimerization is promoted. The dimerization reaction is an exothermic process, and thus would result in an exothermic signal. A possible mechanism for the dehydration and dimerization (assuming 6-coordiantated) process for Cu2+ inside the mesopores of the SBA-15 materials is:
(2) |
(3) |
The weak endothermic signal, coupled with XAFS results, indicates that Cu2+ forms an inner-sphere complex with silica in the pores of mesoporous silicas which is confirmed by MD simulations. Interestingly, the Cu–Si distance (∼3.0 Å) is shorter for porous silicas with high surface loadings, relative to non-porous silica and porous silicas with medium surface loadings (∼3.5 Å). In porous samples, it seems most likely that Cu forms a surface complex with Si through the Oeq bond, as the bond angles for Cu–Oax–Si are ∼100°. However, for non-porous silica, Cu coordination through either Oax or Oeq results in a Cu–O–Si bond angle that is >130 °. Furthermore, there seems to be a systematic increase in the CNCu–Si as a function of decreasing pore diameter and increasing surface loading (Fig. 1, Table 1). While we do observe systematic changes, a Cu–Si distance of 3.0 Å falls within the hypothesized range for a corner sharing mono-dentate surface complex (2.6–3.5 Å), and the CNCu–Si never exceeds 1.0. Therefore, the XAFS data are insufficient to identify the exact geometry (mono- or bi-dentate) of Cu–Si coordination on the examined silica surfaces.
These results provide the first assessment into the impact of nano-scale confinement on adsorption process of Cu2+ on silica with controlled pore sizes. The adsorption process, including adsorption pathway and surface speciation, is pore-size dependent, which has significant implications on ion adsorption, and, therefore, fate and transport in geologically relevant mesopores. The reported findings will aid in the development of more realistic understanding of fate and transport of chemical species in tight, porous geologic media, such as sedimentary rocks.
Footnote |
† Electronic supplementary information (ESI) available: Supporting information contains Fig. S1 showing the surface loading of Cu for the XAFS samples; Table S1 presents the hydration energy, ionic and hydrated radii, as well as hydration number of species involved in calorimetric studies; Fig. S2 showing the adsorption and desorption profiles of the samples used in the calorimetric analysis; details on the preparation of the nano-pore silica models for MD simulation along with Fig. S3 showing the density of silica surface as a function of position, Fig. S4 showing the Cu number density from MD simulations with initial configurations of all monomers or all dimers in solution, Table S2 showing the CN and bond distances from MD simulations comparing Cu coordination chemistry with initial configurations of all monomers or all dimers in solution, Table S3 presenting the fluid composition and pore diameters, and lastly the interaction parameters used in the MD simulation. See DOI: 10.1039/c9en00855a |
This journal is © The Royal Society of Chemistry 2020 |