Bettina
Baumgartner
a,
Ken
Ikigaki
a,
Kenji
Okada
*ab and
Masahide
Takahashi
*a
aDepartment of Materials Science, Graduate School of Engineering, Osaka Prefecture University, Sakai, Osaka, 599-8531, Japan
bJST, PRESTO, 4-1-8 Honcho, Kawaguchi, Saitama 332-0012, Japan. E-mail: k_okada@mtr.osakafu-u.ac.jp; masa@mtr.osakafu-u.ac.jp
First published on 18th June 2021
Pore alignment and linker orientation influence diffusion and guest molecule interactions in metal–organic frameworks (MOFs) and play a pivotal role for successful utilization of MOFs. The crystallographic orientation and the degree of orientation of MOF films are generally determined using X-ray diffraction. However, diffraction methods reach their limit when it comes to very thin films, identification of chemical connectivity or the orientation of organic functional groups in MOFs. Cu-based 2D MOF and 3D MOF films prepared via layer-by-layer method and from aligned Cu(OH)2 substrates were studied with polarization-dependent Fourier-transform infrared (FTIR) spectroscopy in transmission and attenuated total reflection configuration. Thereby, the degrees for in-plane and out-of-plane orientation, the aromatic linker orientation and the initial alignment during layer-by-layer MOF growth, which is impossible to investigate by laboratory XRD equipment, was determined. Experimental IR spectra correlate with theoretical explanations, paving the way to expand the principle of IR crystallography to oriented, organic–inorganic hybrid films beyond MOFs.
Typically, oriented MOF thin films are synthesized by the layer-by-layer (LbL) approach on substrates modified with self-assembled monolayers (SAM).15 Thereby, MOF films composed of 2D sheets (2D MOF) and coordinatively connected 2D sheets by pillar molecules (3D MOFs) with a multitude of different organic and metallic building blocks have been obtained.16–18 Although, this method allows to control the orientation of the lattice plane parallel to the substrate (out-of-plane orientation) by changing the terminated functional group of the SAM, these MOF films lack in-plane orientation.19 Recently, our group reported a synthetic method to obtain three-dimensionally oriented 2D and 3D Cu-based MOF films by epitaxial growth on copper hydroxide,20,21 yielding large scale films with controlled pore alignment.
The standard method to determine the crystallographic orientation, and thus the pore alignment, of MOF films is X-ray diffraction (XRD) in out-of-plane (OOP) and in-plane (IP) configuration, i.e. perpendicular and horizontal relative to the substrate. The degree of in-plane orientation of films can be quantified using azimuthal angle dependent intensity profiles (φ scan) in the in-plane XRD configuration, i.e. rotating the sample around the axis perpendicular to the substrate center (φ axis) at constant X-ray source and detector position.20–22 Besides dedicated diffractometer equipment and rather long measurement times of several hours, this method requires for considerably thick MOF films (typically >40 layers of LbL MOF films) to achieve sufficient signal-to-noise ratios.
With respect to high surface sensitivity, infrared reflection absorption spectroscopy (IRRAS) has proven as powerful tool to determine the structure and orientation of molecular layers on planar surfaces,23,24 to study the initial growth of MOF films,25 or determine structural defects within MOF films.26,27 Polarized IR radiation is reflected from a thin film deposited on metallic surfaces under a grazing angle of incidence. In this configuration, parallel-polarized light is solely absorbed by molecules with chemical bands having their transition dipole moment perpendicular to the substrate's surface.28,29 Thus, IRRAS provides structural, chemical information on the axis perpendicular to the surface. This fact has been exploited by Terfort and co-works to differentiate between surface-attached MOF (SURMOF) films synthesized in different crystallographic orientations.19,30 The carboxylate bands originating from the coordination of the organic linker with the metal nodes are ideally suited to study the MOF orientation: four carboxylate bonds in their bridging configuration are oriented in a Cu-paddle wheel that directs the structure of all MOFs with bimetal building units (e.g. MOFs based on M = Cu, Cr, Mo, etc., compare Fig. 1).31 The perpendicular orientation of the carboxylate bonds to each other and also to the surface allowed Terfort and co-works to determine the out-of-plane orientation of MOF films with different preferred orientation synthesized by LbL synthesis on gold surfaces functionalized with SAMs.19,30 Although IRRAS is well suited to study MOF films prepared from SAMs, the technique is restricted to oriented films deposited on metal surfaces. Furthermore, information on the orientation of chemical bonds apart from perpendicular orientation, i.e. out-of-plane, is not accessible.
Polarization-dependent IR spectroscopy in transmission and attenuated total reflection (ATR) geometry is highly sensitive to structural changes and atomic interactions in x, y and x, y, z-direction (i.e. in-plane and out-of-plane, compare Fig. 1), respectively. Researchers mainly in the fields of macromolecular chemistry and structural biology have taken advantage of these sensitivity to study the degree of crystallinity and the orientation in (semi-)crystalline polymers32–34 or the conformational changes in biomolecules.35–37 In this context, the term infrared crystallography is commonly used, even though the method only provides information on the orientation of molecules and functional groups but not the crystal's periodicity.36
In this contribution, as outlined in Fig. 1, we employed polarization-dependent IR spectroscopy to study the in-plane and out-of-plane orientation of Cu-based 2D and 3D MOF films on Si substrates. Besides confirming the film orientation and comparing favorably with results from XRD measurements just obtained at shorter time scales, additional structural information was retrieved: the orientation of the aromatic linker in the 3D MOF Cu2(BDC)2DABCO (BDC: 1,4-benzenedicarboxylate, DABCO: 1,4-diazabicyclo[2.2.2]octane), to date inaccessible with conventional techniques such as XRD and IRRAS, and highly essential for the accessibility of the pores, was determined to be parallel to the 2D MOF sheets and perpendicular to the bridging carboxylate plane. Furthermore, the initial orientation of MOF films in the LbL synthesis, otherwise only feasible with synchrotron techniques due to the low amount of material, could be investigated.
From these profiles, the degree of in-plane orientation was determined using the full width at half maximum (FWHM) of the peaks and eqn (1):40
(1) |
FTIR spectra recorded with s- and p-polarized light of the same set of films with different degrees of orientation were acquired in transmission using blank silicon as background and are given in Fig. 2 The bands at 1570 cm−1 and 1616 cm−1 stem from the asymmetric carboxylate vibration, and the bands at 1390 cm−1 and 1423 cm−1 are assigned to the symmetric carboxylate vibration, which is consistent with previous reports.39,41 The bands at 1465 and 1370 cm−1 in the 3D MOF stem from C–H vibrations of DABCO with a transition dipole moment perpendicular to the N–N axis.42 No spectral differences between the polarized IR spectra were observed for the MOF films obtained from random Cu(OH)2 nanobelts, while a strong polarization-dependence is visible for three-dimensionally oriented MOF films.
The reason for this becomes clear if we set the orientation of the Cu-paddle wheel of CuBDC and Cu2(1,4-NDC)2DABCO in relation to the Si substrate as given in Fig. 2A: the transition dipole of the νsymm(COO−) band (yellow arrow) is oriented in x-direction, thus, this band can only interact with p-polarized light. Vice versa holds true for the asymmetric vibration that can only interact with s-polarized light. This polarization dependence is visible in the spectra of both aligned films. Furthermore, the transition dipole moments of the C–H vibrations of DABCO present in the 3D MOFs are oriented in x-direction and hence show the same polarization-dependence as the equally oriented dipole of the νsymm(COO−) band. Note that, even if DABCO can rotate freely around the N–N axis, this would not change the fact that it C–H groups only interacts with p-polarized light. For a perfectly aligned film, no bands for the symmetric and asymmetric vibration would be visible for p- and s-polarized light, respectively. However, the investigated MOF films show a degree of in-plane orientation of 0.80 and 0.82 determined from XRD data, respectively, hence, misaligned crystallites are present. This misalignment leads to vibrations of νsymm(COO−) in the s-polarized spectrum and νasymm(COO−) in the p-polarized spectrum. This fact allows to use the linear dichroism, i.e. the ratio of absorbance for s- and p-polarization, of each COO− vibration to calculate the degree of in-plane orientation FIR from IR transmission spectra. The respective band areas A were integrated, and FIR was retrieved using the following equations:
(2) |
(3) |
Thereby, FIR(νasymm(COO−)) values of 0.80 and 0.80 were obtained from the spectra of both aligned MOF films in Fig. 2, which is in excellent agreement with the degree of orientation obtained from XRD experiment. FIR(νsymm(COO−)) = 0.68 for CuBDC, which has less correlation with the corresponding FXRD value. We attribute this deviation of FIR to FXRD to crystallite tilting and a different structure, as we will discuss in the last section.
To demonstrate that this method is valid over the entire range of in-plane orientation values, we prepared CuBDC and Cu2(1,4-NDC)2DABCO films from Cu(OH)2 films with different degrees of orientation. The in-plane orientation of all films was determined with φ-scans in XRD and FTIR transmission spectroscopy. The results are compared in Fig. 3 and show similar values for FIR values >0.5. A lower correlation was found for the νsymm(COO−) band for CuBDC. Note that FIR = 0.5 corresponds to completely unoriented films, while FIR = 0 results only from films perfectly aligned perpendicular to films with FIR = 1 (see Fig. S6 in ESI† for graphical illustration). In contrast, FXRD is an arbitrarily defined ratio that allows comparison between oriented films.40 In addition, peak fitting and determining the FWHM becomes increasingly prone to errors for FXRD < 0.5 and yields −∞ for films without preferential orientation and flat φ-scans. For these reasons, a correlation between FIR and FXRD is physically not valid and shall only be used to compare films during material development for F > 0.5. Nevertheless, these findings show that the degree of in-plane orientation of 2D and 3D MOF films in the given boarders can be derived from simple FTIR transmission spectra on Si substrates, which allows for fast screening without the need of dedicated diffraction instrumentation or substrates with metallic coatings.
Fig. 3 Correlation of degree of orientation F obtained from XRD and FTIR spectroscopy in transmission for CuBDC and Cu2(1,4-NDC)2DABCO films for νsymm(COO−) and νasymm(COO−) bands. |
(4) |
(5) |
(6) |
A⊥ ∝ (μyEy)2 = (μ0sinαsinβEy)2 | (7a) |
A∥ ∝ (μxEx)2 + (μzEz)2 = (μ0sinαcosβEx)2 + (μ0cosαEz)2 | (7b) |
The simplified eqn (7a) and (7b) hold true for MOFs with the molecule axis aligned along the coordination system of the ATR crystal, as in the case for the studied MOF films (see ref. 44 and 45 for more complex systems), and only for aligned dipoles, not for dipole distributions. The νsymm(COO−) and νasymm(COO−) bands in the films investigated have dipole moments with a and β being either 0° or 90°, which simplifies interpretation of polarization dependent spectra. However, the trigonometric relationship allows also to study bands under a specific angle.
The ratio of absorbances A‖ and Anormal measured for p-polarized and s-polarized light, respectively, is called the dichroic ratio R:
(8) |
Experimentally, R can be determined by comparing the absorbance of a specific band in both polarizations. R = 2 for an angle of incidence of 45° (as used in this study) and isotropic bulk samples such as water (see Fig. S9† for spectra). Note that R is unity for isotropic media investigated in transmission. The calculated field amplitudes for varying refractive indices of the sample are depicted in Fig. 4 (see ESI† for further information). The fraction of Ey remains constant, while Ex and Ez are affected by the refractive index of the sample (see ESI† for further information on the determination of the refractive index of the films). With the relative intensities of the evanescent wave in each direction of Ex2 = 0.30, Ey2 = 0.33, Ez2 = 0.37 at hand, the orientation of MOF films can be determined. 3D MOFs with different crystallographic orientation were prepared on Si ATR crystals and the SEM images of the films and corresponding ATR-IR spectra of Cu2(BDC)2DABCO and Cu2(1,4-NDC)2DABCO films are given in Fig. 5. Their crystallographic orientations were confirmed by XRD (see Fig. S10 in ESI†) and the orientation of the MOFs in x, y and z-direction is shown schematically on the left in Fig. 5.
The Cu–Cu axis of Cu-paddle wheel unit and the N–N axis of the DABCO pillar are oriented perpendicular to the substrate in 3D MOF films with [001] preferred OOP orientation (compare Fig. 5D). Both MOF films with [001] preferred OOP orientation show no in-plane orientation (Fig. 5A). This gives rise to transition dipole moments of the νasymm(COO−) bands oriented in z-direction, while the dipole moments of the νsymm(COO−) band are aligned in the x–y plane, which is consistent with the recorded IR spectra: νasymm(COO−) is only visible in the p-polarized spectrum and only contributions of MOF crystallites deviating from the z-axis alignment contribute to this band in the s-polarized spectrum (compare Table 1 for R values). As MOF films with [001] preferred OOP orientation rotate freely around the z-axis and due to the fact that both νsymm(COO−) and νsymm(C–H) are oriented in the x–y plane, these bands show the same absorbance for both polarizations. Two types of MOFs films with [010] preferred OOP orientation (compare Fig. 5E) were prepared from random and aligned Cu(OH)2 nanobelts, yielding MOFs films with and without in-plane orientation (Fig. 5B and C). For 3D MOFs prepared from randomly oriented Cu(OH)2, the transition dipole moment of the νasymm(COO−) rotates freely around the z-axis, characterized by identical absorbance in p- and s-polarized spectra. For the transition dipole moment of the νsymm(COO−) and of ν(C–H) bands determining the dichroic ratio is more complex as contributions in all three axes have to be considered (see ESI† for full explanation).
R (νasymm) predicted | R (νasymm) founda | R (νsymm) predicted | R (νasymm) founda | |
---|---|---|---|---|
a Top and bottom values correspond to Cu2(BDC)2DABCO and Cu2(1,4-NDC)2DABCO, respectively. | ||||
x: random | ∞ | 5.0 | 0.9 | 1.1 |
y: random | ||||
3.7 | 1.1 | |||
z‖[001] | ||||
x: random | 0.9 | 1.0 | 2 | 2.0 |
y: random | ||||
0.9 | 2.0 | |||
z‖[010] | ||||
x‖[100] | 0 | 0.2 | ∞ | 5.5 |
y‖[001] | ||||
0.3 | 5.0 | |||
z‖[010] |
A dichroic ratio of 2 was derived for this case as found in the spectra of both 3D MOFs. Lastly, 3D MOF films prepared from oriented Cu(OH)2 nanobelts show a high degree of IP and OOP orientation. The transition dipole moment of the νasymm(COO−) band is oriented in y-direction, thus yielding absorbance in the s-polarized spectra. The dipole moments of νsymm(COO−) and νsymm(C–H) bands are oriented in x- and z-direction and therefore interact with p-polarized light. Analogous to the previous section, φ scans were recorded and FXRD was determined to be 0.83 and 0.82 for Cu2(BDC)2DABCO and Cu2(1,4-NDC)2DABCO, respectively. FIR was determined from the ATR spectra (using eqn (2) and (3)) to be 0.84 and 0.83 for νsymm(COO−) band. For aligned films, FIR(νasymm(COO−)) = FIR,OOP and was determined to be 0.83 and 0.82 for the Cu2(BDC)2DABCO film and Cu2(1,4-NDC)2DABCO film, respectively.
Note that also transmission IR spectra show distinct differences that can be used to determine the MOF orientation (compare ESI† for explanation). This set of 3D MOF films demonstrates that different crystallographic orientations can be easily distinguished and the degree of in- and out-of-plane orientation can be determined using ATR-IR spectroscopy.
Polarization-dependent ATR spectra of CuBDC, Cu2(BDC)2DABCO and Cu2(1,4-NDC)2DABCO with 3 and 10 layers, respectively, are given in Fig. 6. While due to the low film thickness of the films with three layers, no reflections in the XRD patterns were detected using standard integration times, the polarization dependence of the νasymm(COO−) is clearly visible for the 3D MOFs films. Spectra of CuBDC after 3 and 10 layer deposition are identical with the spectra of CuBDC films from random Cu(OH)2. Cu2(BDC)2DABCO fits the structure of the film with [001] preferred OOP orientation without in-plane orientation: the νasymm(COO−) band shows strong polarization dependence, as its transition dipole moment is aligned in z-direction, thus, this band is only visible in the p-polarized spectrum. Free rotation around the z-direction yield identical absorbance values for the νsymm(COO−) band. The νasymm(COO−) bands allow to determine the degree of out-of-plane orientation (FIR(νasymm(COO−)) = FIR,OOP) according to eqn (2) to be 0.95 and 0.97 for 3 and 10 layers, respectively. A similar behavior was found for Cu2(1,4-NDC)2DABCO for three deposited layers. However, the polarization-dependent spectra of the film with 10 layers show an increased absorbance for the p-polarized spectrum for all bands (FIR,OOP= 0.70 and 0.57 for 3 and 10 layers, respectively). This indicates that the preferential order in z-direction of the Cu2(1,4-NDC)2DABCO MOF decreases with increasing the number of deposition cycles. Although being less ordered, a dichroic ratio of R = 2 (for isotropic materials) is not observed. We attribute the deviation from R = 2 to the low film thickness that is known to decrease the field amplitude of Ez and thus the dichroic ratio for thin films converges to 1 for a film thickness of 0.48 Given these results, polarization-dependent ATR spectroscopy allows for quick inspection and in-depth investigations of the orientation of MOF ultrathin films that are otherwise only accessible with long integration times or in synchrotron facilities.
This finding was further verified in polarization dependent IR transmission spectra of the out-of-plane vibration at 798 cm−1, which has its dipole moment oriented in y-direction and perpendicular to the band at 1506 cm−1. Strong absorbance of the Si-ATR crystal at wavenumber <1000 cm−1 prevents measurements in ATR configuration. With a single path through the substrate in transmission configuration, however, the vibrational modes around 800 cm−1 can be analyzed (see Fig. S14 in ESI† for spectra). The band at 798 cm−1 shows high absorbance in the s-polarized spectrum, which is in line with the dipole assignment for an aromatic ring oriented perpendicular to the carboxylate plane.
Here, IR spectroscopy provides for the first time information on the linker orientation in Cu2(BDC)2DABCO, which is not accessible with XRD measurements as the rotation of the aromatic linker causes no change in the MOF patterns (compare calculated patterns for crystal structures with different linker orientation in Fig. S16 in ESI†).
In contrast to Cu2(BDC)2DABCO, the aromatic linker vibration at 1515 cm−1 of Cu2(1,4-NDC)2DABCO shows no polarization dependence in the ATR spectra (compare Fig. 5). However, a strong polarization dependence, identical to Cu2(BDC)2DABCO, is found for the out-of-plane vibrations in the region around of 800 cm−1 for the aligned MOF film (compare spectra in Fig. S17 in ESI†). Although this confirms a preferential orientation of the aromatic linkers with similar linker orientation as found in the Cu2(BDC)2DABCO, the polarization dependence is less pronounced. This is most likely due to the fact that the 1,4-NDC linker can either rotate freely as it has been suggested for coordination polymers,50 or because the aromatic ring is not oriented parallel or perpendicular to the Cu-paddle wheel plane and is oriented under an angle of 48° as found in the reported crystal structure.51
We considered these structures and tested them against the IR spectra: for linkers parallel to the surface (γ = 0°), both carboxylate vibrations would show no contribution in z-direction. Thus, the transmission spectra and the ATR spectra would be identical. However, the band ratio νsymm(COO−)/νasymm(COO−) of 0.77 for transmission spectra of random CuBDC compared to 0.92 for the same film in ATR configuration suggests that the νsymm(COO−) band has a component in z-direction, which is not accessible in transmission. From this ratio, a percentage of 19% of linkers aligned perpendicular to the surface was derived. The fraction of CuBDC with linkers oriented perpendicular to the surface causes a 100 reflection in the out-of-plane XRD pattern at 2θ = ∼8.2°, while the parallel oriented fraction leads to the same reflection in the in-plane pattern. Typically, the occurrence of this reflection in both XRD configurations is used to confirm the presence of the Cu paddle wheel subunit connected by BDC linkers, and thus, the 2D MOF structure. Since XRD patterns obtained from different configurations cannot be compared quantitatively, we cannot assign the 100 reflection in the OOP pattern unambiguously to either the fraction of 19% of misaligned crystallites or the presence of a 2D MOF structure. However, the usually low porosity observed for CuBDC39 and the position of the COO− bands in the IR spectrum further point towards a more densely packed layered structure. While in 3D MOFs the νasymm(COO−) band is located between 1610–1640 cm−1,30 this band is located at 1570 cm−1 for CuBDC, which is the same position as reported for layered structures.25 Therefore, a different copper carboxylate connectivity, as reported for the layered structure, can be assumed in the CuBDC structure. Note that the methodology to calculate the degree of in-plane orientation FIR based on IR spectroscopy in transmission, as presented in the previous section, is still applicable for this structure: FIR derived from νasymm(COO−) is not affected by the tilting as its direction of the transition dipole moment is still in x–y plane. In summary, polarization-dependent IR spectroscopy provided information of the types of coordination bonds and their three-dimensional orientations in the CuBDC structure, which are difficult to reveal by XRD. It can be expected that the combination of XRD and polarization-dependent IR spectroscopy will contribute significantly to the complete clarification of the real structure of other coordination polymers and MOFs with unknown structures.
Future work will combine the experimental setup with adsorption studies of probe molecules, similar to ref. 56 and 57. These would not just allow to investigate the orientation of the MOF itself but the alignment of guest molecules during the adsorption process. Insights into the orientation and adsorption site of the probe molecules will deepen the understanding of interactions between adsorbent and adsorbate, which is key to exploit the full potential of MOF films.
The presented principles are not limited to MOF films but can be easily translated to other thin films, e.g. inorganic, organic and inorganic–organic hybrid thin films. For instance, the orientation of nanocrystal films based on heavy atoms can be easily studied by X-ray and electron diffraction. However, these nanocrystals are typically stabilized with organic ligands, which significantly influence the nanocrystal's properties. For organic ligands and the orientation of functional groups, diffraction techniques lack in sensitivity. In general, inorganic–organic composite materials account for the majority of nanomaterials with highly promising properties currently under intense investigations. Consequently, the potential of IR crystallography has yet to be fully exploited, allowing to fill the information gap left by diffraction techniques and giving access to key information of hybrid material structures.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d1sc02370e |
This journal is © The Royal Society of Chemistry 2021 |