Characteristic Features of CO2 and CO Adsorptions to Paddle-Wheel

Aug 8, 2017 - (58) A small model SM1 [Cu2(O2CH)4] was employed to evaluate the interaction of gas molecule with the Cu at the high-quality level, as s...
0 downloads 6 Views 1MB Size
Subscriber access provided by University of Florida | Smathers Libraries

Article 2

Characteristic Features of CO and CO Adsorptions to Paddle-Wheel-Type Porous Coordination Polymer Jia-Jia Zheng, Shinpei Kusaka, Ryotaro Matsuda, Susumu Kitagawa, and Shigeyoshi Sakaki J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.7b02707 • Publication Date (Web): 08 Aug 2017 Downloaded from http://pubs.acs.org on August 9, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Characteristic Features of CO2 and CO Adsorptions to PaddleWheel-Type Porous Coordination Polymer Jia-Jia Zheng,†,‡ Shinpei Kusaka,† Ryotaro Matsuda,†,§ Susumu Kitagawa,*,†,∥ and Shigeyoshi Sakaki*,‡ †

Institute for Integrated Cell-Material Sciences (WPI-iCeMS), Kyoto University, Katsura, Nishikyo-ku, Kyoto 615-8510, Japan



Fukui Institute for Fundamental Chemistry, Kyoto University, Nishi-hiraki cho, Takano, Sakyo-ku, Kyoto 606-8103, Japan

§

Department of Applied Chemistry, Graduate School of Engineering, Nagoya University, Chikusa-ku, Nagoya 464-8603, Japan ∥Department

of Synthetic Chemistry and Biological Chemistry, Graduate School of

Engineering, Kyoto University, Katsura, Nishikyo-ku, Kyoto 615-8510, Japan

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ABSTRACT: Adsorptions of CO, N2, NO, and CO2 in a paddle-wheel-type porous coordination polymer (PCP) [Cu(aip)]n (aip = 5-azidoisophthalate) were investigated with ONIOM[MP4(SDQ):ωB97XD] method using a model system consisting of two [Cu2(O2CC6H4-R)4] units (R = H and Me) and one [Cu2(O2CC6H4-R)4] unit; namely dimer and monomer models. The experimental CO adsorption position was reproduced well by the present calculation with the dimer model. For adsorptions of CO, N2, NO, and CO2 in the dimer model, the position of gas molecule deviates from the normal one that is found in the monomer model and becomes more distant from the surrounding phenyl group(s) of the neighbour [Cu(aip)] unit. For all these gas molecules, the calculated binding energy (BE) at the deviating adsorption position is larger than that at the normal one against our expectation that the normal position is the best for the gas adsorption. The deviation of gas adsorption position arises from the interaction between the organic linker (O2CC6H4-R moiety) and gas molecule. For all cases, the exchange repulsion with the organic linker decreases to a larger extent than the attractive electrostatic and dispersion interactions decrease, as going from the normal position to the deviating one. To enhance the binding energy of gas molecule, introduction of electron-donating substituent on phenyl moiety is computationally recommended for this PCP.

2 ACS Paragon Plus Environment

Page 2 of 38

Page 3 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

INTRODUCTION Porous coordination polymers (PCPs) or metal organic frameworks (MOFs)1,2 have attracted recent attentions as potential materials for gas storage,3-6 gas separation,7-11 and catalysis.12-16 Because a variety of metal ions and organic linkers are available for synthesis, geometry and property of PCP can be well tuned for achieving the purpose.17 Actually, many PCPs with excellent adsorption ability for target gas molecule have been designed and synthesized by tuning geometries and properties of PCPs.7-11 To solve the CO2 problem, it is now important to adsorb CO2 molecule efficiently either from air or from large point sources such as factory and power plant.18 Thus, a lot of efforts have been made to synthesize PCPs excellent for CO2 adsorption.19-23 For those PCPs, the open metal site (OMS) has been often utilized for the interaction with CO2. Recently, one experimental work elucidated that CO2 molecule interacts with OMS through O atom in an η1-O coordination form and the M-O-C angle is 120º to 130º, depending on the M, unlike CO (the M-C-O angle is 180º), interestingly.24 CO2 adsorption to PCPs has been investigated by many computational studies such as atomistic and continuum modelling,25 classical molecular dynamics and Monte Carlo simulations,26-29 and the first principle and quantum mechanics calculations.30-33 However, the determining factor(s) for the CO2 adsorption and even the reason of the coordination structure of CO2 with OMS have been unclear. Another important purpose is to separate target gas molecule from mixture containing several kinds of gas molecules with similar physicochemical properties such as CO/N2 mixture. Flexible PCPs, which undergo structural transformation upon external stimuli such as adsorption/desorption of some particular gas molecule,34,35 have been recognized as a promising candidate for the separation of target gas molecule from mixture. The structural

3 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

response of such flexible PCP to target gas molecule is an origin of the gate-opening gas adsorption which leads to significant increase in adsorption amount of target gas molecule. For instance, the gate-opening gas adsorption was recently applied to the separation of CO from N2 using a paddle-wheel-type PCP [Cu(aip)]n (aip = 5-azidoisophthalate, n = large number nearly infinite);36 this is named Cu(aip)-PCP hereafter. This PCP undergoes structural transformation upon CO adsorption/desorption, while the gate-opening does not occur in the N2 case. The difference in structural response between CO and N2 adsorptions was discussed in terms of the CO and N2 binding energies with the Cu-OMS, as follows: Though one Cu atom is coordinated with one oxygen atom of carboxylate of the neighbour paddle-wheel unit in addition to four oxygen atoms of carboxylates bridging two Cu atoms in the case of guest-free structure (Scheme 1a), CO molecule coordinates with the Cu-OMS to break this Cu-O bond (Scheme 1b) to induce significant change of structure. On the other hand, N2 does not strongly interact with the Cu-OMS to induce no structural transformation via the Cu-O bond breaking. One can expect that such gate-opening mechanism could work in CO2 separation from CO2/N2 mixture, if the CO2 adsorption with the Cu-OMS is stronger than the N2 adsorption and similar to the CO adsorption; it should be remembered that the binding energy is one factor for the gate-opening and other factors such as the morphology, surface termination, etc. are also important for the gate-opening. Therefore, the evaluation of binding energy of gas molecule with the Cu-OMS is important as one factor for understanding and predicting whether the gate-opening mechanism may work well in gas adsorption or not. One more interesting feature observed in the CO-adsorbed Cu(aip)-PCP (Scheme 1b) is that CO molecule interacting with the Cu-OMS is sandwiched by two phenyl moieties of a neighbour paddle-wheel unit and its adsorption position deviates from the normal

4 ACS Paragon Plus Environment

Page 4 of 38

Page 5 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

adsorption one that is observed in the CO interaction with one paddle-wheel unit [Cu2(O2CC6H4-R)4], as shown in Schemes 1c and 1d; hereafter, the gas adsorption position in one [Cu2(O2CC6H4-R)4] unit is named “normal position”. It is likely that the deviating adsorption position in the Cu(aip)-PCP is induced by steric repulsion of CO with phenyl moieties of the neighbour paddle-wheel unit. If so, the binding energy of CO must become smaller in the Cu(aip)-PCP than that with one paddle-wheel unit. This suggests that the presence of organic linker is not favourable for CO adsorption. Thus, it is necessary to evaluate how much the binding energy decreases at such deviating adsorption position and to know what type of linker is favourable for gas adsorption. Computational approach at various levels of theory30-33,37-51 has been extensively carried out to investigate interactions between gas molecule and PCPs. The density functional theory (DFT) combined with dispersion-corrected functional52 has been often employed in those studies. However, the use of more accurate post-Hartree-Fock method such as MP2– MP4 and CCSD(T) is highly desirable, because the binding energy of gas molecule with PCP is determined by delicate balance among various interactions such as electrostatic (ES), charge transfer (CT), exchange repulsion (EXR), and dispersion (DIS) interactions. Because PCP is a very large system, one of the reasonable computational methods is the ONIOM procedure,53,54 in which whole system is calculated with such a low-cost method as DFT with dispersion-corrected functional and an important moiety is calculated with the post-Hartree-Fock method. We recently applied the ONIOM[MP2.5:M06-2X] method to the adsorptions of CO2 and CS2 into the Hofmann-type PCP {Fe(Pz)[Pt(CN)4]}n and successfully evaluated adsorption positions and binding energies of CO2 and CS2.47 Snurr and co-workers also employed the ONIOM procedure to evaluate CO2 binding energy with CPO-27.32 These successful examples suggest that the ONIOM method is useful in

5 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

investigating the interaction of gas molecule with PCPs. In this work, we theoretically investigated interactions of such gas molecules as CO, N2, NO, and CO2 with a paddle-wheel type PCP [Cu(aip)]n using model systems consisting of one and two Cu paddle-wheel units [Cu2(O2CC6H4-R)4]2 (R = H and Me). Our purposes here are to investigate the adsorption structures of these gas molecules in the [Cu(aip)]n, to evaluate their binding energies, and to clarify the roles of Cu-OMS and organic linker in determining the adsorption structure and adsorption energy. It is also our intention to elucidate the reasons why the adsorption position in the [Cu(aip)]n deviates from the normal one but the binding energy is enhanced at such deviating adsorption position, which will be seen below.

COMPUTATIONAL DETAILS AND MODLES Though the real Cu(aip)-PCP is an infinite system (Schemes 1a and 1b; see Figure S1 in Supporting Information (SI) for the structure from another viewpoint), cluster model systems were employed here to apply post-Hartree-Fock method to the Cu(aip)-PCP. As shown in Scheme 1c, we employed a model system consisting of two Cu paddle-wheel units [Cu2(O2CC6H4-R)4]2 (R = H and Me), where the azido group in Cu(aip)-PCP was replaced by R for simplicity; this is named dimer model, Dm-R (R = H or Me), hereafter. The Cuα and Cuβ represent the outside Cu atom and the inside one of [Cu2(O2CC6H4-R)4]2, respectively; see Scheme 1c. Because the gas molecule at the β site is surrounded by two phenyl moieties in the Dm-R like that in the real Cu(aip)-PCP (Scheme 1c), the structural situation of the gas molecule in the Dm-R resembles that of gas molecule in the real Cu(aip)-PCP. Gas molecule at the α site is not surrounded at all by phenyl moiety, which is similar to that in a monomeric [Cu2(O2CC6H4-R)4] unit. For the comparison, we

6 ACS Paragon Plus Environment

Page 6 of 38

Page 7 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

investigated the interaction of gas molecule with a monomeric [Cu2(O2CC6H4-R)4] model, which is named monomer model, Mm-R (Scheme 1d). These dimer and monomer models with gas molecule (L) are named Dm-R-L and Mm-R-L hereafter. In the geometry optimization of Dm-R and Dm-R-L, the position of Cuα was fixed to be the same as the experimental one but the other moiety was fully optimized. Because each Cu2+ (d9) in the Cu paddle-wheel unit has one unpaired electron, there are two possible spin states, open-shell singlet and triplet states in one Cu paddle-wheel unit. Previous theoretical studies show that the Cu paddle-wheel system has an open-shell singlet ground state, which is slightly more stable than triplet state.46,49 However, the binding energy of gas molecule with the Cu paddle-wheel unit is calculated to be almost the same between the triplet and open-shell singlet states.46,49 Thus, the calculation was carried out here at the high spin state for both monomer (S = 1) and dimer (S = 2) models. All geometry optimizations were performed by the density functional theory (DFT) with the ωB97XD functional.55 The 6-31G(d) basis sets were used for carboxylate group, where a set of diffuse functions was added to O atom. The 6-31G basis sets were used for phenyl and substituent groups (R= H and Me). For all gas molecules, the 6-31G(d) basis sets were used. For Cu, the LANL2DZ basis set was used to represent the valence electrons and the effective core potentials (ECPs) were used for replacing core electrons.56 Subsequent frequency calculations were carried out at the same level of theory.57 The ONIOM method53,54 was used to evaluate the binding energy of gas molecule with this PCP, because it is particularly reliable for incorporating dispersion interaction.58 A small model SM1 [Cu2(O2CH)4] was employed to evaluate the interaction of gas molecule with the Cu at the high quality level, as shown in Scheme 2. Another small model SM2 consisting of [Cu(O2C-Ph)2] (Scheme 2) was used to evaluate the interaction of gas

7 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

molecule with two phenyl moieties at the high quality level, where one Cu(II) was involved to compensate negative charges of [(O2C-Ph)2]2-; see also Figure S2 in SI for SM1 and SM2. The binding energy (BEβ) of gas molecule adsorbed at the β site was calculated using Dm-R at a low level and SM1 and SM2 at a high level. The binding energy (BEα) of gas molecule at the α site was calculated using Dm-R at a low level and SM1 at a high-level; the superscripts “α” and “β” are used hereafter to represent the coordination sites; for instance, Lα means gas molecule L at the α site. The evaluation method of the binding energy is presented in SI; see page S3. In the ONIOM calculation, better basis sets were employed. The (311111/22111/411/11) basis set including two f polarization functions was employed for Cu, where its core electrons were replaced by Stuttgart-Dresden-Born (SDB) ECPs.59,60 The 6-311G(2d) basis sets were used for gas molecules. For other atoms, the 6-311G(d) basis sets were used, where a set of diffuse functions was added to O atoms of carboxylate group. The basis set superposition error (BSSE) was corrected by the counterpoise method.61 Natural bond orbital (NBO) analysis was made using the DFT method with the ωB97XD functional, where the better basis sets were employed. These calculations were performed with the Gaussian 09 program package.62 Localized molecular orbital (LMO) energy decomposition analysis63 was carried out with the GAMESS program,64 using SM1 and SM2 models.

RESULTS AND DISCUSSION Geometry and Binding Energy of CO2 with Mm-R Prior to the discussion of the interaction of gas molecule with Dm-R, we will briefly

8 ACS Paragon Plus Environment

Page 8 of 38

Page 9 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

discuss first the interaction of CO2 molecule with Mm-R to understand the normal coordination structure of CO2 without interference by other moiety such as phenyl group in PCPs, because the CO2 interaction structure with OMS has not been discussed well compared to the CO and N2 coordinations.49 In Mm-H-CO and Mm-H-N2, CO and N2 take a position on the Cu-Cu axis with a linear η1-end-on coordination form, as shown in Figures 1a and 1b; see Table S1 in SI for details in optimized geometrical parameters. NO and CO2 also take an η1-end-on coordination form, but their coordination positions deviate from the Cu-Cu axis unlike CO and N2 (Figures 1c and 1d). Also, NO and CO2 approach one O atom of carboxylate (donated as OC), where the OC-Cu-N-O and OC-Cu-O-C dihedral angles are nearly 0°. These adsorption positions are named “normal position”, as defined above. Important geometrical parameters such as Cu-L distance and Cu-Cu-L angle differ little between R = H and Me, indicating that the R substituent influences little the adsorption structure of gas molecule. This is reasonable because the R substituent is distant from the adsorption site. Consistent with the small difference in geometry, the charge distribution in Mm-R-L is almost the same between R = H and Me; see Table S2 in SI. In the normal CO2 adsorption structure, the Cu-O-C angle is not linear but 117° (Figure 1d). This coordination structure agrees with the experimental structure of CO2-adsorbed CPO-27-M (M = Mg, Mn, Fe, Co, Ni, Cu, or Zn).24 It is likely that the CO2 adsorption position is determined by both of HOMO-LUMO interaction and charge distribution of CO2. As well known, the HOMO of CO2 is non-bonding π, which expands perpendicularly to the O-C-O axis unlike the HOMO (lone pair) of CO, as shown in Scheme 3. As a result, the O atom of CO2 wants to interact with the Cu atom and the Cu-O-C angle tends to take around 90° to provide a good orbital overlap between the HOMO of CO2 and the empty Cu 4pz orbital. Simultaneously, the positively charged C atom of CO2 is withdrawn towards the

9 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

negatively charged OC atom of the carboxylate ligand (Scheme 3). As a result, the Cuβ-O-C angle becomes 117° and the Cuα-Cuβ-O angle is 167º in the normal position. Though NO coordination structure is different from the CO coordination one (Figure 1c), we omitted discussion because the NO interaction with transition metal complex has been discussed well.65,66 The ONIOM[MP4(SDQ):ωB97XD]-calculated binding energy (BE) of Mm-R-L decreases in the order CO > CO2 > NO > N2, as shown in Table 1; the R substituent influences little their binding energies. In comparison with previous experimental and computational studies on the HKUST-1,30,46,48,67 the binding energy is underestimated here but the decreasing trend of the binding energy (CO > CO2 > NO > N2) is the same as that of the previous works, suggesting that the method employed here describes well the relative binding energy of gas molecule with Cu-OMS. It is likely suggested that the gate-opening mechanism could be applied to CO2 adsorption because the CO2 binding energy is not different so much from the CO binding energy and much larger than that of N2.

Geometries and Binding Energies of CO, N2, NO, and CO2 Molecules with Dm-R Optimized geometries of Dm-R with such gas molecules as L = CO, N2, NO, and CO2, are shown in Figure 2; see Table S3 in SI for geometrical parameters. The CO molecule at the α site takes a linear η1-end-on coordination structure (Figure 2a), in which the carbon atom of COα takes a position on the Cuα-Cuβ axis. This is a typical adsorption structure observed in CO interaction with OMS of PCP.46,48,49 For the CO adsorption at the β site, on the other hand, the carbon atom takes a position deviating from the Cuα-Cuβ axis; the CuαCuβ-C and Cuβ-C-O angles are 169° and 170°, respectively, both of which differ from those of the normal adsorption structure. These angles agree with the experimental values in the

10 ACS Paragon Plus Environment

Page 10 of 38

Page 11 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

CO-adsorbed Cu(aip)-PCP (168° and 175° for Cuα-Cuβ-C and Cuβ-C-O angles, respectively).36 Also, the Cuα-Cuβ, Cuβ-Cuβ, and the average Cu-O distances of Dm-H-CO agree with the experimental values, as shown in Table S3; see page S8 for the brief discussion of changes in the Cuα-Cuβ, Cuβ-Cuβ, and the average Cu-O distances by CO adsorption. For the N2 adsorption (Figure 2b), N2α takes a position on the Cuα-Cuβ axis with a linear η1-end-on coordination form but N2β takes a position deviating from the Cuα-Cuβ axis like the COβ adsorption, where the Cuα-Cuβ-N and Cuβ-N-N angles are 167° and 165°, respectively. NO adsorption position at the α site moderately deviates from the Cuα-Cuβ axis (the Cuβ-Cuα-Nα angle = 175°) like that in the Mm-H-NO (Figures 1c and 2c). At the β site, the NO adsorption position deviates from the Cuα-Cuβ axis (the Cuα-Cuβ-Nβ angle = 171°) more than does the NOα position (Figure 2d). Also, the dihedral angle OCβ-Cuβ-Nβ-Oβ changes to 42° from 0° in the normal one, where the OCβ represents the O atom of carboxylate which is close to the β site; see Figures 2a to 2d for OCβ. For the CO2 adsorption at the α site, the O atom takes a position deviating from the Cuα-Cuβ axis like that in Mm-H-CO2 (Figures 1d and 2d); however, the Cuβ-Cuα-Oα and Cuα-Oα-Cα angles of Dm-H-CO2 are moderately smaller than those of the Mm-H-CO2, suggesting that the CO2 position and orientation are flexible, as found in previous studies.48,68 At the β site, the CuαCuβ-Oβ and Cuβ-Oβ-Cβ angles (169° and 119°, respectively) of Dm-H-CO2 are slightly larger than those of the Mm-H-CO2 and the OCβ-Cuβ-Oβ-Cβ dihedral angle increases to 13° from 3°, indicating that CO2 position at the β site is distant from the phenyl moiety of the neighbour [Cu2(O2CC6H4-R)4] unit.69 The above results lead us to the reasonable explanation described below: If gas molecule at the β site existed on the Cuα-Cuβ axis like that in the normal position, the position of the gas molecule was close to the phenyl moiety of the neighbour

11 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

[Cu2(O2CC6H4-R)4] unit to give rise to significant steric repulsion and destabilize the gas adsorption. To reduce such destabilization, the gas molecule has to take the deviating position from the Cuα-Cuβ axis that is distant from the phenyl moiety. The similar deviating position of gas adsorption is found in Dm-Me, too. Because the difference is not large and the same trend is observed between R = H and Me, we wish to skip the discussion of the difference between them. The ONIOM[MP4(SDQ):ωB97XD]-calculated binding energies (BEs) of these gas molecules with Dm-R are summarized in Table 1. The binding energy (BEα) at the α site is almost the same as that of L with Mm-R, as expected. However, the BEβ is unexpectedly larger than BEα for all these gas molecules. The larger BEβ value than BEα indicates that the gas adsorption is enhanced by the organic linker compared to that of one paddle-wheel unit. These results are surprising because the normal position is considered the best for the interaction of gas molecule with the Cu-OMS and the β-adsorption position deviates from the normal one probably due to the steric repulsion with the organic ligand. The BEβ decreases in the order CO2 ≈ CO > NO > N2 for Dm-H and CO > CO2 > NO > N2 for DmMe. Also, the BEβ value is larger for R = Me than for R = H, while the BEα is moderately smaller for R = Me than for R = H (Table 1). These results suggest that the Me substituent enhances the BEβ value by providing perturbation to the electronic structure of the phenyl moieties surrounding the gas molecule. Another interesting result is that the BEβ of CO2 is similar to that of CO, suggesting that the gate-opening mechanism could work in the CO2 adsorption to Cu(aip)-PCP.

Analysis of Binding Energies of Gas Molecules with Cu Center It is of considerable interest to elucidate what roles the Cu-OMS and the organic linker

12 ACS Paragon Plus Environment

Page 12 of 38

Page 13 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

play in determining the gas adsorption position and the binding energy at the β site. As shown in Figure 3, the sum of MP2-calculated CO binding energies with SM1 and SM2 shows the energy minimum similar to that of the realistic model (Dm-H-CO); see Figures S3 and S4 in SI for ωB97XD-calculated potential energy surfaces and those of CO2 adsorption. These results indicate that these SM1 and SM2 models are useful for elucidating the reasons why the deviation of gas adsorption position occurs and why the BE becomes larger at the deviating position than at the normal one. The binding energies of gas molecules with SM1 were analysed at both the normal structure and the deviating one (LN and Lβ, respectively) in Table 2, using energy decomposition analysis with localized molecular orbital at the Hartree-Fock (HF) level.63 Geometrical parameters of the deviating structures were taken to be the same as those in the optimized structures of Dm-R-L. In this analysis, the binding energy BEHF at the HF level is decomposed into electrostatic interaction (ES), exchange repulsion (EXR), and the sum of charge transfer (CT), polarization (Pol), and other mixing terms (Mix); see Ref. 63 for the explanations of these terms. The dispersion term (DIS) is defined as the difference in binding energy between at the MP4(SDQ) and the HF levels. MP4(SDQ) For all gas molecules, the binding energy BE SM1 at the MP4(SDQ) level is

moderately smaller at the deviating position than at the normal one, as shown in Table 2; we define that the negative value represents the stabilization energy and vice versa in Tables 2 and 3. For the CO adsorption, the dispersion interaction differs little between the normal and deviating structures, indicating that the change in binding energy occurs at the Hartree-Fock level. The ES term becomes more negative (more attractive) but the EXR term becomes more positive (more repulsive) in the deviating structure than in the normal one, while the CT+POL+Mix and DIS terms become moderately more attractive.70 These

13 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

results indicate that the CO binding energy is smaller at the deviating position than at the normal one because the EXR term becomes more repulsive than the other terms become more attractive. The EXR term arises from the overlap of occupied orbitals between gas molecule and Cu(aip)-PCP. As shown in Figure 4, the doubly occupied dπ and dπ* MOs of Cu(aip)-PCP expand on the xz and yz planes. The deviation of CO position from the normal one increases orbital overlap of the σ lone-pair orbital (HOMO) of CO with these dπ and dπ* MOs of Cu(aip)-PCP, and thereby, the EXR destabilization increases at the deviating structure. Because of such change in the EXR term, the interaction between the CO and the Cu site becomes weak at the deviating structure. For the N2 and NO adsorptions, on the other hand, the EXR becomes less repulsive at the deviating position than at the normal one, because the Cu-N distance is longer at the deviating position than at the normal one (Figures 2c and 2d).71 Because of the elongation of the Cu-N distance, the ES and CT+Pol+Mix terms become less attractive. For the NO case, the DIS also becomes less attractive. Because the sum of those attractive interactions MP4(SDQ) decreases more than the EXR decreases, the BE SM1 is smaller at the deviating position

than at the normal one. MP4(SDQ) For the CO2 interaction with SM1, the BE SM1 differs little between the normal

(CO2N) and deviating (CO2β) adsorption positions, suggesting that the CO2 adsorption structure is flexible, as mentioned above. The ES slightly decreases (less attractive), the EXR repulsion moderately decreases (less repulsive) and the CT+Pol+Mix slightly increases (more attractive), but the DIS changes little, as going from the CO2N to the CO2β, MP4(SDQ) leading to almost the same BE SM1 at both the normal and deviating positions. The

larger ES of CO2N arises from the electrostatic interaction between the positively charged C of CO2 and the negatively charge OC of the carboxylate shown in Scheme 3. For the CO2β,

14 ACS Paragon Plus Environment

Page 14 of 38

Page 15 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

the Cuβ-O-C angle increases to 119° and the OCβ-Cuβ-O-C dihedral angle increases to 13°, which elongates the distance between the C of CO2 and the OC to decrease the ES stabilization. Because the Cuα-Cuβ-O angle increases at the deviating position, the O of CO2 approaches the Cuα-Cuβ axis. As a result, the orbital overlap of the non-bonding π (HOMO) of CO2 with the Cu 4pz orbital becomes larger and that with the 3dxz/3dyz orbitals become smaller at the deviating position, leading to the larger CT+Pol+MIX and smaller EXR terms.72 MP4(SDQ) In conclusion, the BE SM1 with the Cu center is smaller at the deviating position than MP4(SDQ) at the normal one except for CO2 in which the BE SM1 differs little between CO2N and

CO2β. This indicates that the larger BEβ value arises from some attractive interaction between gas molecule and organic linker.

Analysis of Binding Energies of Gas Molecules with Phenyl Moiety We made energy decomposition analysis of the interaction of gas molecule with SM2, MP4(SDQ) as shown in Table 3. The BE SM2 is positive at the normal position but decreases very

much to negative value at the deviating position for all the gas molecules examined here, indicating that the interaction with SM2 is the origin of the larger BE at the deviating MP4(SDQ) position than at the normal one. For instance, the BE SM2 of CO with SM2 is 5.76 kcal

mol-1 larger at the deviating position than at the normal one but that with SM1 is 0.35 kcal mol-1 smaller at the deviating position (Table 2). It is clearly concluded that the interaction with SM2 is responsible for the deviating adsorption position. For the CO adsorption, the EXR becomes much smaller (less repulsive) by 15.79 kcal mol-1 than the ES term decreases by 7.74 kcal mol-1, as going from CON to COβ. The CT+Pol+Mix term moderately changes. Based on these results, it appears that the small

15 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

EXR term with the phenyl moieties is a main factor for the deviating adsorption position of CO. But, the change in the EXR is not enough to understand the reason why the CO binding energy is larger at the deviating position than at the normal one (Table 1); if only the EXR term between the CO and the phenyl moieties is the reason for taking the deviating position, the CO binding energy at the β site must be smaller than at the α site because CO molecule at the α site does not suffer at all from the EXR with the phenyl moieties. The other reason must exist for stabilizing the deviating position. It is noted that HF MP4(SDQ) the BE SM2 with SM2 is positive but the BE SM2 is negative at the deviating position,

indicating that the DIS term is important. Actually, the ES, EXR, and CT+Pol+Mix terms decrease by about 70% but the DIS decreases by about 30%, as going from CON to COβ (see values in parentheses in Table 3). In other words, the EXR with phenyl moieties is sensitive to the position change and decreases very much as going from the normal position to the deviating one, and the ES and CT+Pol+Mix decrease to a similar extent to the EXR but the DIS is not sensitive and decreases much less than the EXR. Thus, the CO binding energy is larger at the deviating position than at the normal one. The similar explanation can be presented for CO2, N2, and NO, in which the DIS term decreases much less than the ES, EXR, and CT+Pol+Mix terms decrease as going from the normal positon to the deviating one; see Table 3. These results suggest that the DIS term plays important role(s) in stabilizing gas adsorption in this PCP. Notably, the dispersion interaction plays crucial role(s) in many host-guest interactions of PCPs, as reported for H2 and benzene adsorptions in MOF-5.73,74 The above conclusion suggests that appropriate modification of phenyl moieties enhances gas adsorption to paddle-wheel PCPs. We investigated CO binding energy with two Cu paddle-wheel units [Cu2(O2CC6H4-R)4]2 employing several substituents R = CH3,

16 ACS Paragon Plus Environment

Page 16 of 38

Page 17 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

CF3, OCH3, and tBu. The ONIOM[MP4(SDQ):ωB97XD]-calculated BE at the β site increases in the order CF3 (–7.19) < OCH3 (–7.26) ~ CH3 (–7.28) < tBu (–7.47), where in parentheses are BE (in kcal mol-1). This increasing order is almost parallel to the DIS term MP4(SDQ) with SM2 and the binding energy ( BE SM2 ) with SM2; see Table S6 for DIS and MP4(SDQ) terms. These results suggest that the use of electron-releasing substituent BE SM2

enhances the dispersion interaction and the CO binding energy.

CONCLUSIONS We theoretically investigated the interactions of CO, N2, NO, and CO2 with a porous coordination polymer Cu(aip)-PCP consisting of Cu paddle-wheel units, using ONIOM[MP4(SDQ):ωB97XD] method. The optimized adsorption structure of COβ in the realistic model Dm-R is in good agreement with the experimental structure of CO-adsorbed Cu(aip)-PCP, where the CO adsorption position deviates from the normal linear position. The similar deviation of adsorption position at the β site was found for N2, NO, and CO2 in Dm-R. For all these gas adsorptions, the gas molecule at the β site tends to move to the deviating position distant from the phenyl group of the neighbouring [Cu2(O2CC6H4-R)4] paddle-wheel unit. Interestingly, the ONIOM[MP4(SDQ):ωB97XD]-calculated binding energy (BEβ) of gas molecule at the β site is larger than those at the α site and the normal adsorption position; the BEβ decreases in the order CO2 ≈ CO > NO > N2 for R = H and CO > CO2 > NO > N2 for R = CH3. The similar BEβs of CO and CO2 suggest that the gate-opening adsorption of CO2 may occur in Cu(aip)-PCP, if the other factors such as morphology, crystal edge structure, etc. are favourable for the gate-opening. The energy decomposition analysis at the Hartree-Fock level revealed that the phenyl

17 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

moiety of the organic linker plays an important role in deviating the adsorption position and enhancing the binding energy of gas molecule with this PCP. For all cases, the ES, EXR, and CT+Pol+Mix terms with the phenyl moiety become smaller at the deviating structure than at the normal one. However, the DIS term decreases much less than the ES, EXR, and CT+Pol+Mix terms. This means that the EXR is more sensitive to the position change of gas molecule than the DIS and decreases more than does the DIS as going from the normal position to the deviating one and thereby the binding energy becomes larger (more negative) at the deviating adsorption structure. This explanation suggests that the phenyl moiety enhances gas adsorption through the DIS term (i.e. dispersion interaction). Introduction of electron-donating substituents on the phenyl moiety is computationally predicted to enhance the binding energy, because such substituent induces large dispersion interaction with gas molecule. It is concluded that the presence of organic linkers surrounding an adsorbed gas molecule at the open-metal site leads to the deviation of adsorption position but increases the binding energy. These theoretical findings reveal the important roles of organic linkers on the adsorption of gas molecule into PCPs.

ASSOCIATED CONTENT Supporting Information Detailed equations in the ONIOM scheme, NBO charges of atoms and atomic groups in gas-adsorbed Dm-R and Mm-R, detailed structural parameters, additional presentations of real and computational models, results of LMO energy decomposition analysis for Lβ in Dm-Me, plots of ωB97XD-calculated binding energies of CO against deviation angles in SM1 and SM2 models, complete references of 54, 62, and 64, Cartesian coordinates and

18 ACS Paragon Plus Environment

Page 18 of 38

Page 19 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

total energies of Mm-R and Dm-R with and without gas molecules. This material is free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Authors [email protected] [email protected]

Notes The authors declare no completing financial interest.

ACKNOWLEDGMENTS This work is supported by the PRESTO and ACCEL project of the Japan Science and Technology Agency (JST), and JSPS KAKENHI Grant-in-Aid for Young Scientists (B) (Grant No. 25870360), for Challenging Exploratory Research (Grant No. 25620187) and for Specially Promoted Research (Grant No. 25000007). iCeMS is supported by the World Premier International Research Initiative (WPI) of the Ministry of Education, Culture, Sports, Science and Technology, Japan (MEXT). We are thankful for Institute for Molecular Science (Okazaki, Japan) to perform some of calculations with the computer systems.

REFERENCES (1) Kondo, M.; Yoshitomi, T.; Seki, K.; Matsuzaka, H.; Kitagawa, S. Three-Dimensional Framework

with

Channeling

Cavities

for

Small

19 ACS Paragon Plus Environment

Molecules:

{[M2(4,

4′-

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

bpy)3(NO3)4]·xH2O}n (M = Co, Ni, Zn). Angew. Chem. Int. Ed. 1997, 36, 1725-1727. (2) Li, H.; Eddaoudi, M.; Groy, T. L.; Yaghi, O. M. Establishing Microporosity in Open Metal-Organic Frameworks:  Gas Sorption Isotherms for Zn(BDC) (BDC = 1,4Benzenedicarboxylate). J. Am. Chem. Soc. 1998, 120, 8571-8572. (3) Morris, R. E.; Wheatley, P. S. Gas Storage in Nanoporous Materials. Angew. Chem. Int. Ed. 2008, 47, 4966-4981. (4) Murray, L. J.; Dinca, M.; Long, J. R. Hydrogen Storage in Metal-Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1294-1314. (5) Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; Bloch, E. D.; Herm, Z. R.; Bae, T.-H.; Long, J. R. Carbon Dioxide Capture in Metal-Organic Frameworks. Chem. Rev. 2012, 112, 724-781. (6) Suh, M. P.; Park, H. J.; Prasad, T. K.; Lim, D.-W. Hydrogen Storage in Metal-Organic Frameworks. Chem. Rev. 2012, 112, 782-835. (7) Kitagawa, S.; Matsuda, R. Chemistry of Coordination Space of Porous Coordination Polymers. Coord. Chem. Rev. 2007, 251, 2490-2509. (8) Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Selective Gas Adsorption and Separation in MetalOrganic Frameworks. Chem. Soc. Rev. 2009, 38, 1477-1504. (9) Czaja, A. U.; Trukhan, N.; Mueller, U. Industrial Applications of Metal-Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1284-1293. (10) Li, J.-R.; Sculley, J.; Zhou, H.-C. Metal-Organic Frameworks for Separations. Chem. Rev. 2012, 112, 869-932. (11) Wu, H.; Gong, Q.; Olson, D. H.; Li, J. Commensurate Adsorption of Hydrocarbons and Alcohols in Microporous Metal-Organic Frameworks. Chem. Rev. 2012, 112, 836-868. (12) Ma, L.; Abney, C.; Lin, W. Enantioselective Catalysis with Homochiral Metal-Organic

20 ACS Paragon Plus Environment

Page 20 of 38

Page 21 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Frameworks. Chem. Soc. Rev. 2009, 38, 1248-1256. (13) Farrusseng, D.; Aguado, S.; Pinel, C. Metal-Organic Frameworks: Opportunities for Catalysis. Angew. Chem. Int. Ed. 2009, 48, 7502-7513. (14) Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. MetalOrganic Framework Materials as Catalysts. Chem. Soc. Rev. 2009, 38, 1450-1459. (15) Corma, A.; Garcia, H.; Llabres i Xamena, F. X. L. I. Engineering Metal-Organic Frameworks for Heterogeneous Catalysis. Chem. Rev. 2010, 110, 4606-4655. (16) Yoon, M.; Srirambalaji, R.; Kim, K. Homochiral Metal-Organic Frameworks for Asymmetric Heterogeneous Catalysis. Chem. Rev. 2012, 112, 1196-1231. (17) Kitagawa, S.; Kitaura, R.; Noro, S. Functional Porous Coordination Polymers. Angew. Chem. Int. Ed. 2004, 43, 2334-2375. (18) Sanz-Perez, S. Z.; Murdock, C. R.; Didas, S. A.; Jones, C.W. Direct Capture of CO2 from Ambient Air. Chem. Rev. 2016, 116, 11840-11876. (19) Liu, J.; Thallapally, P. K.; McGrail, B. P.; Brown, D. R.; Liu, J. Progress in Adsorption-Based CO2 Capture by Metal-Organic Frameworks. Chem. Soc. Rev. 2012, 41, 2308-2322. (20) Li, J. R.; Ma, Y.; McCarthy, M. C.; Sculley, J.; Yu, J.; Jeong, H. K.; Balbuena, P. B.; Zhou, H. C. Carbon Dioxide Capture-Related Gas Adsorption and Separation in MetalOrganic Frameworks. Coord. Chem. Rev. 2011, 255, 1791-1823. (21) Lin, Z. J.; Lu, J.; Hong, M.; Cao, R. Metal-Organic Frameworks Based on Flexible ligands (FL-MOFs): Structures and Applications. Chem. Soc. Rev. 2014, 43, 5867-5895. (22) Seoane, B.; Coronas, J.; Gascon, I.; Benavides, M. E.; Karvan, O.; Caro, J.; Gascon, J. Metal-Organic Framework Based Mixed Matrix Membranes: A Solution for Highly Efficient CO2 Capture? Chem. Soc. Rev. 2014, 44, 2421-2454.

21 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(23) Darabi, A.; Jessop, P. G.; Cunningham, M. F. CO2-Responsive Polymeric Materials: Synthesis, Self-Assembly, and Functional Applications. Chem. Soc. Rev. 2016, 45, 43914436. (24) Queen, W. L.; Hudson, M. R.; Bloch, E. D.; Mason, J. A.; Gonzalez, M. I.; Lee, J. S.; Gygi, D.; Howe, J. D.; Lee, K.; Darwish, T. A. Comprehensive Study of Carbon Dioxide Adsorption in the Metal-Organic Frameworks M2(dobdc) (M = Mg, Mn, Fe, Co, Ni, Cu, Zn). Chem. Sci. 2014, 5, 4569-4581. (25) Keskin, S.; Sholl, D. S. Selecting Metal-Organic Frameworks as Enabling Materials in Mixed Matrix Membranes for High Efficiency Natural Gas Purification. Energy Environ. Sci. 2010, 3, 343-351; Assessment of a Metal-Organic Framework Membrane for Gas Separations Using Atomically Detailed Calculations: CO2, CH4, N2, H2 Mixtures in MOF5. Ind. Eng. Chem. Res. 2009, 48, 914-922. (26) Haldoupis, E.; Nair, S.; Sholl, D. S. Finding MOFs for Highly Selective CO2/N2 Adsorption Using Materials Screening Based on Efficient Assignment of Atomic Point Charges. J. Am. Chem. Soc. 2012, 134, 4313-4323. (27) Watanabe, T.; Sholl, D. S. Accelerating Applications of Metal-Organic Frameworks for Gas Adsorption and Separation by Computational Screening of Materials. Langmuir 2012, 28, 14114-14128. (28) Li, S.; Chung, Y. G.; Snurr, R. Q. High-Throughput Screening of Metal-Organic Frameworks for CO2 Capture in the Presence of Water. Langmuir 2016, 32, 10368-10376. (29) Qiao, Z.; Peng, C.; Zhou, J.; Jiang, J. High-Throughput Computational Screening of 137953 Metal-Organic Frameworks for Membrane Separation of a CO2/N2/CH4 Mixture. J. Mater. Chem. A 2016, 4, 15904-15912. (30) Grajciar, L.; Wiersum, A. D.; Llewellyn, P. L.; Chang, J.-S.; Nachtigall, P.

22 ACS Paragon Plus Environment

Page 22 of 38

Page 23 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Understanding CO2 Adsorption in CuBTC MOF: Comparing Combined DFT-Ab Initio Calculations with Microcalorimetry Experiments. J. Phys. Chem. C 2011, 115, 1792517933. (31) Poloni, R.; Smit, B.; Neaton, J. B. Ligand-Assisted Enhancement of CO2 Capture in Metal-Organic Frameworks. J. Am. Chem. Soc. 2012, 134, 6714-6719. (32) Yu, D.; Yazaydin, A. O.; Lane, J. R.; Dietzel, P. D.; Snurr, R. Q. A Combined Experimental and Quantum Chemical Study of CO2 Adsorption in the Metal-Organic Framework CPO-27 with Different Metals. Chem. Sci. 2013, 4, 3544-3556. (33) Poloni, R.; Lee, K.; Berger, R. F.; Smit, B.; Neaton, J. B. Understanding Trends in CO2 Adsorption in Metal-Organic Frameworks with Open-Metal Sites. J. Phys. Chem. Lett. 2014, 5, 861-865. (34) Horike, S.; Shimomura, S.; Kitagawa, S. Soft Porous Crystals. Nat. Chem. 2009, 1, 695-704. (35) Schneemann, A.; Bon, V.; Schwedler, I.; Senkovska, I.; Kaskel, S.; Fischer, R. A. Flexible Metal-Organic Frameworks. Chem. Soc. Rev. 2014, 43, 6062-6096. (36) Sato, H.; Kosaka, W.; Matsuda, R.; Hori, A.; Hijikata, Y.; Belosludov, R. V.; Sakaki, S.; Takata, M.; Kitagawa, S. Self-Accelerating CO Sorption in a Soft Nanoporous Crystal. Science 2014, 343, 167-170. (37) Getman, R. B.; Bae, Y.-S.; Wilmer, C. E.; Snurr, R. Q. Review and Analysis of Molecular Simulations of Methane, Hydrogen, and Acetylene Storage in Metal-Organic Frameworks. Chem. Rev. 2012, 112, 703-723. (38) Yang, Q.; Liu, D.; Zhong, C.; Li, J. R. Development of Computational Methodologies for Metal-Organic Frameworks and Their Application in Gas Separations. Chem. Rev. 2013, 113, 8261-8323.

23 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 38

(39) Odoh, S. O.; Cramer, C. J.; Truhlar, D. G.; Gagliardi, L. Quantum-Chemical Characterization of the Properties and Reactivities of Metal-Organic Frameworks. Chem. Rev. 2015, 115, 6051-6111. (40) Sagara, T.; Klassen, J.; Ganz, E. Computational Study of Hydrogen Binding by MetalOrganic Framework-5. J. Chem. Phys. 2004, 121, 12543-12547. (41) Sagara, T.; Klassen, J.; Ortony, J.; Ganz, E. Binding Energies of Hydrogen Molecules to Isoreticular Metal-Organic Framework Materials. J. Chem. Phys. 2005, 123, 01471. (42) Lee, T. B.; Kim, D.; Jung, D. H.; Choi, S. B.; Yoon, J. H.; Kim, J.; Choi, K.; Choi, S.H. Understanding the Mechanism of Hydrogen Adsorption into Metal-Organic Frameworks. Catal. Today 2007, 120, 330-335. (43) Kuc, A.; Heine, T.; Seifert, G.; Duarte, H. A. On the Nature of the Interaction between H2 and Metal-Organic Frameworks. Theor. Chem. Acc. 2008, 120, 543-550; H2 adsorption in Metal-Organic Frameworks: Dispersion or Electrostatic Interactions? Chem. Eur. J. 2008, 14, 6597-6600. (44)

Pianwanit,

A.;

Kritayakornupong,

C.;

Vongachariya,

A.;

Selphusit,

N.;

Ploymeerusmee, T.; Remsungnen, T.; Nuntasri, D.; Fritzsche, S.; Hannongbua, S. The Optimal Binding Sites of CH4 and CO2 Molecules on the Metal-Organic Framework MOF5: ONIOM Calculations. Chem. Phys. 2008, 349, 77-82. (45) Deshmukh, M. M.; Sakaki, S. Binding Energy of Gas Molecule with Two Pyrazine Molecules as Organic Linker in Metal-Organic Framework: Its Theoretical Evaluation and Understanding of Determining Factors. Theor. Chem. Acc. 2011, 130, 475-482. (46) Rubes, M.; Grajciar, L.; Bludsky, O.; Wiersum, A. D.; Llewellyn, P. L.; Nachtigall, P. Combined Theoretical and Experimental Investigation of CO Adsorption on Coordinatively Unsaturated Sites in CuBTC MOF. ChemPhysChem, 2012, 13, 488-495.

24 ACS Paragon Plus Environment

Page 25 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(47) Deshmukh, M. M.; Ohba, M.; Kitagawa, S.; Sakaki, S. Absorption of CO2 and CS2 into the Hofmann-Type Porous Coordination Polymer: Electrostatic versus Dispersion Interactions. J. Am. Chem. Soc. 2013, 135, 4840-4849. (48) Supronowicz, B.; Mavrandonakis, A.; Heine, T. Interaction of Small Gases with the Unsaturated Metal Centers of the HKUST-1 Metal Organic Framework. J. Phys. Chem. C, 2013, 117, 14570-14578. (49) Hijikata, Y.; Sakaki, S. Interaction of Various Gas Molecules with Paddle-Wheel-Type Open Metal Sites of Porous Coordination Polymers: Theoretical Investigation. Inorg. Chem. 2014, 53, 2417-2426. (50) Lee, K.; Howe, J. D.; Lin, L.-C.; Smit, B.; Neaton, J. B. Small-Molecule Adsorption in Open-Site Metal-Organic Frameworks: A Systematic Density Functional Theory Study for Rational Design. Chem. Mater. 2015, 27, 668-678. (51) Mavrandonakis, A.; Vogiatzis, K. D.; Boese, A. D.; Fink, K.; Heine, T.; Klopper, W. Ab Initio Study of the Adsorption of Small Molecules on Metal-Organic Frameworks with Oxo-centered Trimetallic Building Units: The Role of the Undercoordinated Metal Ion. Inorg. Chem. 2015, 54, 8251-8263. (52) Grimme, S. Accurate Description of van der Waals Complexes by Density Functional Theory including Empirical Corrections. J. Comput. Chem. 2004, 25, 1463-1473; Semiempirical GGA-type Density Functional Constructed with a Long-range Dispersion Correction. J. Comput. Chem. 2006, 27, 1787-1799. (53) Dapprich, S.; Komáromi, I.; Byun, K. S.; Morokuma, K.; Frisch, M. J. A New ONIOM Implementation in Gaussian98. Part I. The Calculation of Energies, Gradients, Vibrational Frequencies and Electric Field Derivatives. J. Mol. Struct.(THEOCHEM) 1999, 461, 1-21.

25 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(54) Chung, L. W.; Sameera, W. M. C.; Ramozzi, R.; Page, A. J.; Hatanaka, M.; Petrova, G. P.; Harris, T. V.; Li, X.; Ke, Z.; Liu, F., et al. The ONIOM Method and Its Applications. Chem. Rev. 2015, 115, 5678-5796. (55) Chai, J.-D.; Head-Gordon, M. Long-Range Corrected Hybrid Density Functionals with Damped Atom-Atom Dispersion Corrections. Phys. Chem. Chem. Phys. 2008, 10, 66156620. (56) Hay, P. J.; Wadt, W. R. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for the Transition Metal Atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270-283. (57) In Dm-R-CO, Dm-Me-N2, Dm-Me-NO, we found several imaginary frequencies around 10 cm-1, arising from vibration of the framework, because the Cuα positions were fixed to be the same as the experimental geometry. Since these imaginary frequencies are quiet small and do not influence the position of gas molecule, we neglected them in the present work. (58) Deshmukh, M. M.; Sakaki, S. Two-Step Evaluation of Binding Energy and Potential Energy Surface of Van der Waals Complexes. J. Comput. Chem. 2012, 33, 617-628. (59) Dolg, M.; Wedig, U.; Stoll, H.; Preuss, H. Energy-Adjusted Ab Initio Pseudopotentials for the First Row Transition Elements. J. Chem. Phys. 1987, 86, 866-872. (60) Figgen, D.; Rauhut, G.; Dolg, M.; Stoll, H. Energy-Consistent Pseudopotentials for Group 11 and 12 Atoms: Adjustment to Multi-Configuration Dirac-Hartree-Fock Data. Chem. Phys. 2005, 311, 227-244. (61) Boys, S. F.; Bernardi, F. The Calculation of Small Molecular Interactions by the Differences of Separate Total Energies. Some Procedures with Reduced Errors. Mol. Phys. 1970, 19, 553-566. (62) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.;

26 ACS Paragon Plus Environment

Page 26 of 38

Page 27 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Cheeseman, J. R.; Scalmani, G.; Barone, V.; B., M.; Petersson, G. A., et al. Gaussian 09, revision D.01; Gaussian, Inc: Wallingford, CT, 2013. (63) Su, P.; Li, H. Energy Decomposition Analysis of Covalent Bonds and Intermolecular Interactions. J. Chem. Phys. 2009, 131, 014102. (64) Schmidt, M. W.; Baldridge, K. K.; Boatz, J. A.; Elbert, S. T.; Gordon, M. S.; Jensen, J. H.; Koseki, S.; Matsunaga, N.; Nguyen, K. A.; Su, S., et al. General Atomic and Molecular Electronic Structure System. J. Comput. Chem. 1993, 14, 1347-1363. (65) Tachikawa, H.; Iyama, T.; Hamabayashi, T. Metal-Ligand Interactions of the Cu-NO Complex at the Ground and Low-Lying Excited States: An Ab Initio MO Study. Electron. J. Theor. Chem. 1997, 2, 263-267. (66) McCleverty, J. A. Chemistry of Nitric Oxide Relevant to Biology. Chem. Rev. 2004, 104, 403-418. (67) Wang, Q. M.; Shen, D.; Bulow, M.; Lau, M. L.; Deng, S.; Fitch, F. R.; Lemcoff, N. O.; Semanscin, J. Metallo-Organic Molecular Sieve for Gas Separation and Purification. Micropor. Mesopor. Mater. 2002, 55, 217-230. (68) Jee, B.; Petkov, P. S.; Vayssilov, G. N.; Heine, T.; Hartmann, M.; Poppl, A. A Combined Pulsed Electron Paramagnetic Resonance Spectroscopic and DFT Analysis of the

13

CO2 and

13

CO Adsorption on the Metal-Organic Framework Cu2.97Zn0.03(Btc)2. J.

Phys. Chem. C 2013, 117, 8231-8240. (69) The Cβ-CPh distance is 2.764 Å at the dihedral angle of 3° and 3.029 Å at the dihedral angle of 13°, where the Cβ is the carbon atom of CO2 and the CPh is the closest carbon atom of the phenyl moiety of the neighbour [Cu2(O2CC6H4-R)4] unit. (70) Because the COβ position is distant from the Cuα-Cuβ axis, the COβ is not favourable for the CT from the CO lone pair to the Cuβ 4pz orbital. However, the CT+Pol+Mix

27 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

becomes moderately more attractive at the deviating COβ position because the Cuβ-COβ distance is shorter than Cuα-COα. The reason why the Cuβ-COβ distance is shorter than Cuα-COα is not clear at this moment. (71) The reason for the Cu-N elongation at the deviating position is not clear at this moment. The change in the Cuα-Cuβ-Nβ angle also contributes to the change in EXR, as follows: The smaller Cuα-Cuβ-Nβ angle than the Cuβ-Cuα-Nα (Figure 2c and 2d) increases the overlap between the lone pair with dπ and dπ* MOs of Cu(aip)-PCP at the deviating position, which leads to the increase in EXR at the deviating position like in COβ. But, the Cu-N bond elongation simultaneously occurs to influence more the EXR because the σ−overlap is larger than others. Thus, the EXR decreases at the deviating position. (72) The Cuβ-O(CO2β) distance becomes shorter than the Cuα-O(CO2α), which is responsible for the larger CT+Pol+Mix. However, it is not clear why the EXR is small at the short Cu-O distance and why the Cuβ-O(CO2β) distance is shorter than the CuαO(CO2α). At this moment, the explanation about these issues is difficult. (73) Sillar, K.; Hofmann, A.; Sauer, J. Ab Initio Study of Hydrogen Adsorption in MOF-5. J. Am. Chem. Soc. 2009, 131, 4143-4150. (74) Amirjalayer, S.; Schmid, R. Adsorption of Hydrocarbons in Metal-Organic Frameworks: A Force Field Benchmark on the Example of Benzene in Metal-Organic Framework 5. J. Phys. Chem. C 2012, 116, 15369-15377.

28 ACS Paragon Plus Environment

Page 28 of 38

Page 29 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Table 1. Binding energy (BE, kcal mol-1) of gas molecule L (L = CO2, CO, N2, and NO) with Dm-R and Mm-R (R = H, Me) calculated by the ONIOM[MP4(SDQ):ωB97XD] method. a L CO N2 NO CO2 a

Mm-H BEMm –5.86 –3.13 –3.63 –4.99

Dm-H BEα BEβ –5.58 –6.11 –2.98 –3.31 –3.37 –4.55 –5.15 –6.16

Mm-Me BEMm –5.80 –3.11 –3.38 –4.82

Dm-Me BEα BEβ –5.38 –7.28 –2.84 –4.09 –3.10 –4.81 –4.38 –6.49

BEα and BEβ denote binding energies for Lα and Lβ with Dm-R. BEMm is BE of L with

Mm-R. Counterpoise correction was made.

Table 2. Various interaction terms (kcal mol-1) of gas molecule with the SM1.a L CON COβ N2N N2β NON NOβ CO2N CO2β a

HF BE SM1

–3.47 –2.91 –1.14 –0.75 0.03 –0.09 –3.94 –3.92

ES –14.14 –15.04 –6.23 –5.85 –5.50 –4.87 –10.96 –10.88

MP4(SDQ) EXR CT+Pol+Mix DIS BE SM1 16.97 –6.30 –3.42 –6.89 18.86 –6.74 –3.63 –6.54 8.21 –3.12 –2.57 –3.71 8.01 –2.91 –2.59 –3.34 9.30 –3.78 –3.86 –3.83 8.42 –3.64 –3.58 –3.67 11.19 –3.78 –1.42 –5.36 10.84 –3.87 –1.44 –5.36

The geometry was taken to be the same as that of R = H; see Table S3 in SI for R = Me.

The binding energy was analyzed at both the normal position (LN) and deviating one (Lβ).

29 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 38

Table 3. Various interaction terms (kcal mol-1) of gas molecule with the SM2.a HF BE SM2 L CON 11.32

COβ

3.94

N2N

13.29

N2β

3.41

NON

9.05

NOβ

2.83

CO2N 9.55 CO2β a

2.46

ES –9.46 –2.72 (–71%)b –7.20 –1.94 (–73%) –5.22 –2.69 (–48%) –8.65 –3.42 (–60%)

EXR 23.02 7.23 (–69%) 22.26 5.72 (–74%) 17.03 6.51 (–62%) 20.64 6.78 (–67%)

CT+Pol+Mix –2.25 –0.74 (–67%) –1.78 –0.36 (–80%) –3.03 –0.99 (–67%) –2.44 –0.86 (–65%)

DIS –5.61 –3.99 (–29%) –7.09 –3.57 (–50%) –5.39 –3.61 (–33%) –5.15 –3.66 (–29%)

MP4(SDQ) BE SM2

5.71 –0.05 6.20 –0.16 3.66 –0.78 4.40 –1.20

The geometry was taken to be the same as that of R = H; see Table S3 in SI for R = Me.

The binding energy was analyzed at both the normal position (LN) and deviating one (Lβ). b

The extent of decrease in relevant interaction term from the normal position (LN) to

deviating one (Lβ) is presented in parentheses.

30 ACS Paragon Plus Environment

Page 31 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Scheme 1. (a) Experimental structurea of activated Cu(aip)-PCP, (b) that of CO-adsorbed Cu(aip)-PCP, and (c) computational dimer model (Dm-R) of Cu(aip)-PCP with adsorbed gas molecule L, and (d) that of monomer model Mm-R.

a

These experimental structures

were re-drown from CIF files of experimental report (ref. 36).

31 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Scheme 2. Small models (SM1 and SM2) employed for high-level calculations in the ONIOM method.

32 ACS Paragon Plus Environment

Page 32 of 38

Page 33 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Scheme 3. Schematic representation of interaction between CO2 and Cu open-metal site.

33 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1. Optimized structures of Cu paddle-wheel unit with (a) CO, (b) N2, (c) NO, and (d) CO2 molecules. Distances are in angstroms. a R = H; see Table S1 for details and those for R = Me.

34 ACS Paragon Plus Environment

Page 34 of 38

Page 35 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 2. Optimized structures of two Cu paddle-wheel units, Dm-H, with (a) CO, (b) N2, (c) NO, and (d) CO2 molecules. Distances are in angstroms. a R = H; see Table S3 for details and those for R = Me.

35 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3. MP2-calculated potential energy surfaces for CO interaction with SM1 and SM2.

36 ACS Paragon Plus Environment

Page 36 of 38

Page 37 of 38

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 4. Overlap between the lone-pair orbital of CO and (a) dπ and (b) dπ* orbitals of Cu(aip)-PCP. Hydrogen atoms are omitted for clarity.

37 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TOC Graphic

38 ACS Paragon Plus Environment

Page 38 of 38