research communications\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Crystal structures of N-[4-(tri­fluoro­meth­yl)phen­yl]benzamide and N-(4-meth­­oxy­phen­yl)benz­amide at 173 K: a study of the energetics of conformational changes due to crystal packing

crossmark logo

aChemistry Department, United States Naval Academy, 572 Holloway Rd, Annapolis, MD 21402, USA
*Correspondence e-mail: waynehp21662@gmail.com

Edited by M. Weil, Vienna University of Technology, Austria (Received 5 October 2021; accepted 27 January 2022; online 8 February 2022)

As a part of our study of the syntheses of aryl amides, the crystal structures of two benzamides were determined from single-crystal X-ray data at 173 K. Both crystal structures contain mol­ecular units as asymmetric units with no solvent in the unit cells. Crystal structure I, TFMP, is the result of the crystallization of N-[4-(tri­fluoro­meth­yl)phen­yl]benzamide, C14H10F3NO. Crystal structure II, MOP, is composed of N-(4-meth­oxy­phen­yl)benzamide, C14H13NO2, units. TFMP is triclinic, space group P[\overline{1}], consisting of two mol­ecules in the unit cell related by the center of symmetry. MOP is monoclinic, space group P21/c, consisting of four mol­ecules in the unit cell. Both types of mol­ecules contain three planar regions; a phenyl ring, an amide planar region, and a para-substituted phenyl ring. The orientations of these planar regions within the asymmetric units are compared to their predicted orientations, in isolation, from DFT calculations. The aryl rings are tilted approximately 60° with respect to each other in both experimentally determined structures, as compared to 30° in the DFT results. These conformational changes result in more favorable environments for N—H⋯O hydrogen bonding and aryl ring π-stacking in the crystal structures. Inter­molecular inter­actions were examined by Hirshfeld surface analysis and qu­anti­fied by calculating mol­ecular inter­action energies. The results of this study demonstrate that both hydrogen bonding and dispersion are essential to the side-by-side stacking of mol­ecular units in these crystal structures. Weaker dispersion inter­actions along the axial directions of the mol­ecules reveal insight into the melting mechanisms of these crystals.

1. Chemical context

Numerous methodologies have been developed to form amide C—N bonds due to their prevalence in biomolecules, such as peptides and proteins, and in synthetic targets (Seward & Jakubke, 2002[Seward, N. & Jakubke, H. D. (2002). Peptides: Chemistry and Biology. Weinheim: Wiley-VCH.]; Greenberg et al., 2000[Greenberg, A., Brenemanm, C. M. & Liebman, J. F. (2000). Amide Linkage: Selected Structural Aspects in Chemistry, Biochemistry, and Materials Science. New York: Wiley-Interscience.]). In particular, aryl amides can be found in a variety of pharmaceutical drugs and in polymers such as KevlarTM (Masse et al., 1998[Masse, C. E., Yang, M., Solomon, J. & Panek, J. S. (1998). J. Am. Chem. Soc. 120, 4123-4134.]; Evano et al., 2004[Evano, G., Schaus, J. V. & Panek, J. S. (2004). Org. Lett. 6, 525-528.], 2008[Evano, G., Blanchard, N. & Toumi, M. (2008). Chem. Rev. 108, 3054-3131.]; Satyanarayana et al., 2007[Satyanarayana, K., Srinivas, K., Himabindu, V. & Reddy, G. M. (2007). Org. Process Res. Dev. 11, 842-845.]; Tanner et al., 1989[Tanner, D., Fitzgerald, J. A. & Phillips, B. R. (1989). Angew. Chem. Int. Ed. Engl. 28, 649-654.]). A series of aryl amides were synthesized and isolated during the development of a copper-mediated concurrent tandem catalytic methodology for the amidation of aryl chlorides (Chang et al., 2019[Chang, R. K., Clairmont, B. P., Lin, S. & MacArthur, A. H. R. (2019). Organometallics, 38, 4448-4454.]). The crystal structures of two of these aryl amides, derived from the cross-coupling of either 4-chloro­benzotrifluoride or 4-chloro­anisole with benzamide, are reported here.

[Scheme 1]

2. Structural commentary

The reported compounds are substituted benzamides containing a para-substituted phenyl ring in place of one of the hydrogen atoms of the amide nitro­gen. In both crystal structures, the asymmetric unit is a single mol­ecule of the compound. Crystal structure I, TFMP, contains an asymmetric unit with a tri­fluoro­methyl­phenyl ring. Crystal structure II, MOP, has an asymmetric unit with a meth­oxy­phenyl ring. The mol­ecular structures in the form of ellipsoid plots are shown in Fig. 1[link]. There is nothing remarkable about the individual bond lengths, bond angles, or planarity of the aryl rings in these mol­ecules.

[Figure 1]
Figure 1
The mol­ecules present in the asymmetric units in (a) TFMP and (b) MOP. Displacement ellipsoids are drawn at the 50% probability level. Hydrogen atoms are represented by spheres of 0.20 Å radius.

Fig. 2[link] contains the unit cells for both crystal structures. Both mol­ecules assume chiral configurations. Because the space groups are centrosymmetric, the unit-cell contents are racemic mixtures containing the enanti­omers of the mol­ecules in symmetry-related positions. In both crystal structures, the mol­ecules align along the mol­ecular axes. This alignment results in the long axes in both unit cells, c = 14.415 (3) Å in TFMP and a = 26.783 (2) Å in MOP.

[Figure 2]
Figure 2
Unit-cell packing of (a) TFMP and (b) MOP.

Both mol­ecules contain three planar regions; a phenyl ring, an amide linkage, and the para-substituted phenyl ring. Rotation of the rings relative to each other can lead to conformations that exist in the crystal structures that differ from the native mol­ecular conformations. The relationship between the conformations of organic mol­ecules and crystal structures has been reported extensively and summarized in the review article by Cruz-Cabeza & Bernstein (2014[Cruz-Cabeza, J. & Bernstein, J. (2014). Chem. Rev. 114, 2170-2191.]). Tilt angles were determined by comparing the angles between normals to least-squares planes as defined by the non-hydrogen atoms in a planar region. Significant tilt angles exist between the planar regions in both mol­ecules in the experimentally determined structures as shown in Fig. 3[link].

[Figure 3]
Figure 3
Views of orientations of planar regions and their dihedral angles (in °) from experimental results (top) and DFT calculations (bottom) for (a) TFMP and (b) MOP. Blue = phenyl ring; green = amide plane; mauve = para-substituted phenyl ring.

3. DFT calculations and results for isolated mol­ecules

Quantum-chemical density functional theory (DFT) calculations were performed to find the conformations of global minimum energy for the two mol­ecules in isolation. Calculations were performed with the GAUSSIAN09 (Frisch et al., 2016[Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Mennucci, B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratchian, H. P., Izmaylov, A. F., Bloino, J., Zheng, G., Sonnenberg, J. L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery, J. A., Jr., Peralta, J. E., Ogliaro, F., Bearpark, M., Heyd, J. J., Brothers, E., Kudin, K. N., Staroverov, V. N., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Rega, N., Millam, J. M., Klene, M., Knox, J. E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E., Yazyev, O., Austin, A. J., Cammi, R., Pomelli, C., Ochterski, J. W., Martin, R. L., Morokuma, K., Zakrzewski, V. G., Voth, G. A., Salvador, P., Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, Ö., Foresman, J. B., Ortiz, J. V., Cioslowski, J. & Fox, D. J. (2016). GAUSSIAN 09, Revision A. 02. Gaussian Inc., Wallingford CT, USA.]) program suite on DoD High Performance Modernization resources. Initial conformer searching was performed at the mol­ecular mechanics level with the MMFF force field as implemented in SPARTAN mol­ecular modeling software (Wavefunction, 2014[Wavefunction (2014). SPARTAN. Wavefunction Inc., Irvine, CA, USA.]). Viable structures were then subjected to complete geometry optimizations in GAUSSIAN09 at the M06-2X/6-31+G(d) level (Zhao & Truhlar, 2008[Zhao, Y. & Truhlar, D. J. (2008). Theor. Chem. Acc. 120, 215-241.]). Frequency calculations were performed at M06-2X/6-31+G(d) to confirm that all stational points were minima. Comparisons of bond lengths and angles between the experimentally determined structures and the DFT calculations can be found in the supporting information.

Tilts of the planar regions from the DFT calculations are also shown in Fig. 3[link]. The amide plane/phenyl ring angles are approximately 29° in the experimental results and 27° in the calculated mol­ecules. In the experimentally determined structures, the angles between para-substituted phenyl rings and the amide planes are 31.4 (2)° in TFMP and 38.4 (4)° in MOP. The DFT calculations yield much smaller angles of 8.5 and 7.9°, respectively. These results indicate that the conformational change due to crystal packing in both mol­ecules is primarily due to ring tilts around the N1—C5 bonds while the rings joined by C8—C9 bonds are essentially oriented the same as in the isolated mol­ecules. A search of benzamides in the Cambridge Structural Database (version 2020.3; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) revealed a number of compounds with similar phenyl ring/amide plane tilts. For example, N-phenyl­benzamide (Wang et al., 2014[Wang, J.-L., Xu, J.-S., Wang, D.-Y., Wang, H., Li, Z.-T. & Zhang, D.-W. (2014). CrystEngComm, 16, 2078-2084.]), N-(4-hy­droxy­phen­yl)benzamide (Tothadi & Desiraju, 2012[Tothadi, S. & Desiraju, G. R. (2012). Cryst. Growth Des. 12, 6188-6198.]), benzamide (Blake & Small, 1972[Blake, C. C. F. & Small, R. W. H. (1972). Acta Cryst. B28, 2201-2206.]) and N-(4-nitro­phen­yl)benzamide (du Plessis et al., 1983[Plessis, M. P. du, Modro, T. A. & Nassimbeni, L. R. (1983). J. Crystallogr. Spectrosc. Res. 13, 179-189.]) all possess amide plane/phenyl ring angles between 28 and 31°. A likely explanation for the consistent amide plane/phenyl ring tilt would be the balance of the attractive O1⋯H14 inter­actions and the repulsive inter­actions of H1⋯H10.

Additional DFT calculations were performed to determine approximate energy differences between the mol­ecules in isolation and conformations found in the crystal structures. To best approximate the conformations in the experimentally determined structures, dihedral angles around the amide linkage were constrained to crystallographic values while all other geometrical parameters were allowed to vary. Tilt angles between phenyl and para-substituted phenyl rings are in good agreement between the X-ray models and DFT calculations. For TFMP, the angles are 59.7 (1)° in the crystal structure and 59.6° in the DFT calculation. For MOP, the angles are 67.4 (1)° in the crystal structure and 66.8° in the DFT calculation. The results of the DFT calculations show that the energies of the conformations in the experimentally determined structures are slightly above those in the isolated mol­ecules, viz. 3.2 kJ mol−1 higher for TFMP and 2.5 kJ higher for MOP.

4. Supra­molecular features

Close packing in both crystal structures is the result of hydrogen bonding, dipole inter­actions and dispersion. Hydrogen bonds were revealed by using the HTAB command in SHELXL (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]) and verified using PLATON (Spek, 2020[Spek, A. L. (2020). Acta Cryst. E76, 1-11.]). The H⋯O contacts are listed in Tables 1[link] and 2[link] and shown in Fig. 4[link]. There is only one type of N—H⋯O inter­action in both crystal structures, in the direction parallel to the a axis for TFMP and the b axis in MOP. There are non-classical carbon-based hydrogen bonds that exist as intra­molecular inter­actions (C6—H6⋯O1) in both crystal structures and inter­molecular contacts (C4—H4⋯O1) in TFMP only. The longer H4⋯O1 contact in MOP (2.95 Å) is a result of the larger ring twist angle between the para-substituted phenyl ring and the amide linkage in MOP, 38.4° versus 31.4° in TFMP.

Table 1
Hydrogen-bond geometry (Å, °) for TFMP[link]

D—H⋯A D—H H⋯A DA D—H⋯A
C6—H6⋯O1 0.95 2.44 2.938 (4) 112
C4—H4⋯O1i 0.95 2.57 3.240 (4) 128
N1—H1⋯O1i 0.99 (1) 2.23 (2) 3.138 (3) 151 (3)
Symmetry code: (i) [x-1, y, z].

Table 2
Hydrogen-bond geometry (Å, °) for MOP[link]

D—H⋯A D—H H⋯A DA D—H⋯A
C6—H6⋯O1 0.95 2.49 2.912 (2) 107
N1—H1⋯O1i 0.96 (1) 2.16 (1) 3.108 (2) 166 (2)
Symmetry code: (i) [x, y-1, z].
[Figure 4]
Figure 4
Hydrogen bonding contacts (in Å) in (a) TFMP and (b) MOP.

Comparisons of hydrogen-bonding regions from the experimentally determined structure and DFT results are shown in Fig. 5[link] for TFMP and MOP. In both cases, the molecules in the crystal structures have a more open environment with larger angles around the donor and acceptor sites and larger donor and acceptor cavities. The increased planar tilt between para-substituted phenyl rings and amide planes is a contributor to the more open hydrogen-bonding environments in the experimentally determined structures.

[Figure 5]
Figure 5
Comparison of hydrogen-bonding environments (in Å) from X-ray results and DFT calculations for (a) TFMP and (b) MOP.

The increased tilt angles between the amide and para-substituted phenyl planes also facilitate the π-stacking in both crystal structures (Table 3[link]). Neighboring environments around aryl rings are shown in Fig. 6[link]. Each aryl ring has close contacts with six other aryl rings. In TFMP, there are contacts between tri­fluoro­methyl­phenyl rings and phenyl rings. In MOP, phenyl rings have close contacts with phenyl rings while meth­oxy­phenyl rings have contacts with other meth­oxy­phenyl rings on neighboring mol­ecules. There are a total of six inter­actions surrounding each aryl ring, with four T-shaped inter­actions and two being a parallel displacement of rings. Neighboring mol­ecules that have parallel displaced rings are involved in the N—H⋯O hydrogen bonding. A qu­anti­tative discussion of the π stacking geometries based upon the approach of Banerjee et al. (2019[Banerjee, A., Saha, A. & Saha, B. K. (2019). Cryst. Growth Des. 19, 2245-2252.]) can be found in the supporting information.

Table 3
π-stacking parameters for TFMP and MOP

All distances are in Å with estimated uncertainties of 0.004. Angles are in ° with estimated uncertainties of 0.2.

Centroid Normal Offset Twist angle
TFMP – surrounding both rings      
4.774 4.672 0.982 59.7
4.718 4.649 0.804 59.7
4.711 4.646 0.780 59.7
4.698 4.611 0.900 59.7
5.361 2.666 4.651 0.0
       
MOP – surrounding phenyl rings      
4.781 4.757 0.478 64.6
4.901 4.875 0.504 64.6
5.248 2.802 4.437 0.0
       
MOP – surrounding meth­oxy­phenyl rings      
4.849 4.658 1.348 68.1
4.831 4.64 1.345 68.1
5.248 2.938 4.349 0.0
[Figure 6]
Figure 6
Hydrogen bonding and π-stacking (in Å) in (a) TFMP and (b) MOP (s.u.'s for centroid distances are approximately 0.005 Å). Riding H atoms are omitted for clarity.

Inter­molecular inter­actions in the remaining axial direction, c in TFMP and a in MOP, are shown in Fig. 7[link]. In TFMP, the axial inter­actions are between a tri­fluoro­methyl group on one mol­ecule and a phenyl ring on its neighbor. In MOP, the closest inter­actions are of two types, meth­oxy­phen­yl–meth­oxy­phenyl inter­actions and phen­yl–phenyl inter­actions. The neighboring phenyl rings have a centroid distance of 6.4 Å and do not overlap.

[Figure 7]
Figure 7
View of contacts (in Å) along the mol­ecular axes in (a) TFMP and (b) MOP.

5. Hirshfeld surfaces and mol­ecular pair inter­action energies

To further examine the supra­molecular environments, Hirshfeld surfaces and mol­ecular pair inter­action energies were calculated for both crystal structures. All of these calculations were performed using the CE-B3LYP method via the TONTO program (Jayatilaka & Grimwood, 2003[Jayatilaka, D. & Grimwood, D. J. (2003). Computational Science|ICCS 2003, edited by P. M. A. Sloot, D. Abramson, A. V. Bogdanov, Y. E. Gorbachev, J. J. Dongarra & A. Y. Zomaya, pp. 142-151. Berlin, Heidelberg: Springer.]) as implemented in CrystalExplorer17 (Spackman et al., 2021[Spackman, P. R., Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Jayatilaka, D. & Spackman, M. A. (2021). J. Appl. Cryst. 54, 1006-1011.]). Inter­action energies use benchmarked models based upon B3LYP/6-31G(d,p) functionals, coupled with appropriate scale factors for electrostatic, polarization, dispersion and repulsion energies. The CE-B3LYP model is benchmarked against B3LYP-D2/6-31G(d,p) counterpoise-corrected energies and has been found to give very good agreement with CCSD(T)/CBS (Turner et al., 2014[Turner, M. J., Grabowsky, S., Jayatilaka, D. & Spackman, M. A. (2014). J. Phys. Chem. Lett. 5, 4229-4255.]).

Hirshfeld surfaces and mol­ecular inter­action energies are shown in Fig. 8[link]. The neighboring mol­ecules fall within 3.8 Å from the mol­ecule inside the Hirshfeld surface. The color coding keys and scaled energies are found in Fig. 9[link]. Although the energy values are reported to 0.1 of a kJ mol−1, the authors of CrystalExplorer17 recommend that the reliability is on the order of 1 kJ mol−1. As a result, the total inter­action energies (Etot) are rounded to a kJ mol−1. As expected, the major Etot energies occur for the side-by-side inter­actions for TFMP (# 1–5) and MOP (# 1–3). The percent contributions to the Eattract from the electrostatic, polarization and dispersion components are reported. Dispersion is the major attractive inter­action in both crystal structures. For mol­ecules with hydrogen-bonded close contacts and for some inter­actions along the mol­ecular axes directions, the electrostatic energies are roughly equal to the dispersion energies. Videos showing 360° rotations of the static views in Fig. 8[link] can be found in the supporting information.

[Figure 8]
Figure 8
Inter­molecular inter­action energies and Hirshfeld surfaces with electrostatic potential (top) and dnorm (bottom) plots are shown for (a) TFMP and (b) MOP. Scales for electrostatic potential are red (−0.0788) to blue (0.1227) au for TFMP and red (−0.0875) to blue (0.1219) au for MOP. Scales for dnorm are red (−0.2905) to blue (0.9711) for TFMP and red (−0.3719) to blue (1.1524) for MOP. The color code for mol­ecular inter­actions is shown in Fig. 9[link].
[Figure 9]
Figure 9
Key for the inter­molecular inter­action energies for TFMP and MOP. Energy units are kJ mol−1.

In Fig. 8[link], the electrostatic potentials, plotted on the Hirshfeld surfaces, show regions of negative charge (red) and positive charge (blue) for both compounds. For TFMP, in Fig. 8[link]a, the electrostatic inter­action of the hydrogen-bonding region is evident but so is the head-to-tail stacking of neighboring mol­ecules due to the attraction of negative tri­fluoro­methyl groups with neighboring positive phenyl hydrogens. For MOP, in Fig. 8[link]b, the electrostatic inter­action of the hydrogen bonding is apparent but the polar nature in the remaining segments of the mol­ecule is localized in the meth­oxy group, contributing to the preference for association of meth­oxy­phenyl rings in the crystal structure.

Fig. 8[link] also includes Hirshfeld surfaces with dnorm surface plots. Inter­molecular contacts less than a van der Waals contact are colored red, roughly equal contacts are white, and contacts longer than a van der Waals contact are blue. White or red contacts should indicate some degree of inter­molecular inter­action of inner and outer atoms at those positions on the Hirshfeld surfaces. Specific close contacts are shown in the supporting information.

Insight into the melting process for these crystals can be obtained from the energy analysis. Melting of these crystals would require overcoming the weak inter­molecular inter­actions along the direction of the mol­ecular axes. In the case of TFMP, the energy required would be on the order of 8–9 kJ mol−1 (inter­actions #6 and #7 in Fig. 9[link]). In MOP, the energy required would only be around 7 kJ mol−1 (inter­actions #5 and #6). Although these energy values are inter­nal energies and not enthalpies, they are reasonable values for heats of fusion and correlate with the melting points of the two crystal structures, 478 K for TFMP and 425 K for MOP (Chang et al., 2019[Chang, R. K., Clairmont, B. P., Lin, S. & MacArthur, A. H. R. (2019). Organometallics, 38, 4448-4454.]). However, for TFMP, mol­ecules should separate equally at the melting point on either side of a mol­ecule. In MOP, mol­ecules will separate first at the phenyl ends of the mol­ecules while the meth­oxy­phenyl ends would be predicted to persist into the liquid phase until enough energy was applied to overcome the 11 kJ mol−1 inter­action energy (inter­action #4).

6. Database survey

The Cambridge Structural Database was searched for possible crystal structures of these compounds. No entries were found for a crystal structure of N-[4-(tri­fluoro­meth­yl)phen­yl]benzamide. A room-temperature crystal structure was found for the N-(4-meth­oxy­phen­yl)benzamide compound (du Plessis et al., 1983[Plessis, M. P. du, Modro, T. A. & Nassimbeni, L. R. (1983). J. Crystallogr. Spectrosc. Res. 13, 179-189.]). The CIF file associated with this study, BUTDOJ, included only atom positions with no atomic displacement parameters. The R factor was listed as 0.106. In the published article, the authors noted that the overlapping reflections made it difficult to make an accurate background correction. This resulted, in the authors' words, `in a somewhat poor set of intensity data for this compound'. For these reasons, we opted to use our redetermination of the crystal structure for the purpose of this publication.

7. Synthesis and crystallization

Details of the synthesis of the title compounds can be found in Chang et al. (2019[Chang, R. K., Clairmont, B. P., Lin, S. & MacArthur, A. H. R. (2019). Organometallics, 38, 4448-4454.]). Product crystals for both compounds were grown by slow diffusion of hexa­nes into a concentrated solution of the amide in ethyl acetate.

8. Refinement

Crystal data, data collection and refinement details are summarized in Table 4[link]. All hydrogen atoms were located in difference-Fourier maps. Final positions for most of the hydrogen atoms were calculated and included in a riding model relative to the bonded, non-hydrogen atoms by use of AFIX commands. Methyl hydrogen atoms were fixed at 0.98 Å from bonded carbon atoms, and phenyl hydrogen atoms were located 0.95 Å from bonded carbon atoms. Hydrogen displacement parameters were isotropic and set at 1.20 times the bonded phenyl carbons and 1.50 times the bonded methyl carbon in MOP. The amide hydrogens were treated differently because of their participation in the hydrogen bonding in these crystal structures. DFIX commands were set at 1.00 Å for these hydrogen atoms to allow for better comparison with the DFT-calculated N—H bond lengths. These hydrogen positions were then refined with independent isotropic displacement parameters. Isotropic extinction was refined in MOP. Although the `standard' independent atom model was used for our analysis, alternative models were considered. Refinements with librationally corrected bond lengths and high-angle refinements were performed. These refinements had no significant effects on the structural results or the energy calculations.

Table 4
Experimental details

  TFMP MOP
Crystal data
Chemical formula C14H10F3NO C14H13NO2
Mr 265.23 227.25
Crystal system, space group Triclinic, P[\overline{1}] Monoclinic, P21/c
Temperature (K) 173 173
a, b, c (Å) 5.3606 (11), 7.7831 (16), 14.415 (3) 26.7830 (15), 5.2477 (3), 8.1343 (5)
α, β, γ (°) 77.170 (7), 79.421 (7), 89.719 (7) 90, 97.594 (2), 90
V3) 576.0 (2) 1133.24 (11)
Z 2 4
Radiation type Mo Kα Mo Kα
μ (mm−1) 0.13 0.09
Crystal size (mm) 0.24 × 0.08 × 0.06 0.70 × 0.26 × 0.08
 
Data collection
Diffractometer Bruker APEXII CCD Bruker APEXII CCD
Absorption correction Multi-scan (SADABS; Bruker, 2018[Bruker (2018). APEX3, SAINT and SADABS. Bruker AXS Inc, Madison, Wisconsin, USA.]) Multi-scan (SADABS; Bruker, 2018[Bruker (2018). APEX3, SAINT and SADABS. Bruker AXS Inc, Madison, Wisconsin, USA.])
Tmin, Tmax 0.700, 1.000 0.518, 1
No. of measured, independent and observed [I > 2σ(I)] reflections 11667, 2201, 1415 36318, 2333, 1678
Rint 0.086 0.106
(sin θ/λ)max−1) 0.612 0.626
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.068, 0.189, 1.11 0.049, 0.126, 1.08
No. of reflections 2201 2333
No. of parameters 176 160
No. of restraints 1 1
H-atom treatment H atoms treated by a mixture of independent and constrained refinement H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.26, −0.32 0.20, −0.19
Computer programs: APEX3 and SAINT (Bruker, 2018[Bruker (2018). APEX3, SAINT and SADABS. Bruker AXS Inc, Madison, Wisconsin, USA.]), SHELXT (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), WinGX (Farrugia, 2012[Farrugia, L. J. (2012). J. Appl. Cryst. 45, 849-854.]), Mercury (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]), CrystalExplorer17 (Spackman et al., 2021[Spackman, P. R., Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Jayatilaka, D. & Spackman, M. A. (2021). J. Appl. Cryst. 54, 1006-1011.]), and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]).

Supporting information


Computing details top

For both structures, data collection: APEX3 (Bruker, 2018); cell refinement: SAINT (Bruker, 2018); data reduction: SAINT (Bruker, 2018); program(s) used to solve structure: SHELXT (Sheldrick, 2015a); program(s) used to refine structure: WinGX (Farrugia, 2012), SHELXL (Sheldrick, 2015b); molecular graphics: Mercury (Macrae et al., 2020) and CrystalExplorer17 Spackman et al., 2021); software used to prepare material for publication: publCIF (Westrip, 2010).

N-[4-(Trifluoromethyl)phenyl]benzamide (1) top
Crystal data top
C14H10F3NOF(000) = 272
Mr = 265.23Dx = 1.529 Mg m3
Dm = 1.46 (2) Mg m3
Dm measured by flotation in K2CO3(aq) solution
Triclinic, P1Melting point = 477–478 K
a = 5.3606 (11) ÅMo Kα radiation, λ = 0.71073 Å
b = 7.7831 (16) ÅCell parameters from 2253 reflections
c = 14.415 (3) Åθ = 2.7–23.2°
α = 77.170 (7)°µ = 0.13 mm1
β = 79.421 (7)°T = 173 K
γ = 89.719 (7)°Regular parallelpiped, colourless
V = 576.0 (2) Å30.24 × 0.08 × 0.06 mm
Z = 2
Data collection top
Bruker APEXII CCD
diffractometer
2201 independent reflections
Radiation source: sealed X-ray tube1415 reflections with I > 2σ(I)
Detector resolution: 8.53 pixels mm-1Rint = 0.086
rotating crystal scansθmax = 25.8°, θmin = 1.5°
Absorption correction: multi-scan
(SADABS; Bruker, 2018)
h = 66
Tmin = 0.700, Tmax = 1.000k = 99
11667 measured reflectionsl = 1717
Refinement top
Refinement on F2Primary atom site location: structure-invariant direct methods
Least-squares matrix: fullHydrogen site location: structure-invariant direct methods
R[F2 > 2σ(F2)] = 0.068H atoms treated by a mixture of independent and constrained refinement
wR(F2) = 0.189 w = 1/[σ2(Fo2) + (0.0726P)2 + 0.6085P]
where P = (Fo2 + 2Fc2)/3
S = 1.11(Δ/σ)max < 0.001
2201 reflectionsΔρmax = 0.26 e Å3
176 parametersΔρmin = 0.32 e Å3
1 restraint
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
C10.4290 (8)0.2664 (5)0.9270 (3)0.0361 (10)
C20.5101 (6)0.2716 (5)0.8223 (3)0.0259 (8)
C30.3572 (6)0.3479 (5)0.7572 (3)0.0274 (9)
H30.2072490.4037880.7794240.033*
C40.4208 (6)0.3432 (5)0.6611 (3)0.0269 (8)
H40.3144670.3956720.6174580.032*
C50.6407 (6)0.2620 (4)0.6270 (2)0.0228 (8)
C60.7958 (6)0.1861 (5)0.6926 (3)0.0260 (8)
H60.9472770.1317440.6702700.031*
C70.7301 (6)0.1899 (5)0.7885 (2)0.0253 (8)
H70.8349780.1363840.8324380.030*
C80.9209 (6)0.2491 (5)0.4709 (3)0.0257 (8)
C90.9138 (6)0.2390 (5)0.3692 (3)0.0238 (8)
C100.7131 (6)0.1552 (5)0.3439 (3)0.0265 (8)
H100.5720740.1034210.3920700.032*
C110.7225 (6)0.1487 (5)0.2481 (3)0.0302 (9)
H110.5871550.0916570.2309320.036*
C120.9254 (7)0.2236 (5)0.1773 (3)0.0304 (9)
H120.9289320.2189030.1117500.036*
C131.1249 (7)0.3062 (5)0.2021 (3)0.0303 (9)
H131.2654180.3579140.1535740.036*
C141.1180 (6)0.3127 (5)0.2979 (3)0.0280 (9)
H141.2550990.3684670.3147110.034*
F10.3071 (5)0.4105 (3)0.94336 (17)0.0520 (7)
F20.6234 (5)0.2548 (4)0.97444 (18)0.0599 (8)
F30.2713 (5)0.1294 (4)0.97311 (18)0.0606 (8)
N10.6886 (5)0.2533 (4)0.5288 (2)0.0253 (7)
O11.1224 (4)0.2563 (4)0.49980 (18)0.0359 (7)
H10.541 (5)0.261 (6)0.495 (3)0.052 (13)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
C10.039 (2)0.033 (2)0.037 (2)0.0094 (19)0.0088 (18)0.0111 (18)
C20.0212 (17)0.027 (2)0.030 (2)0.0040 (15)0.0059 (15)0.0067 (16)
C30.0187 (17)0.031 (2)0.033 (2)0.0076 (15)0.0031 (15)0.0104 (17)
C40.0178 (17)0.030 (2)0.036 (2)0.0046 (15)0.0092 (15)0.0095 (16)
C50.0179 (16)0.022 (2)0.029 (2)0.0015 (14)0.0057 (14)0.0065 (15)
C60.0159 (17)0.032 (2)0.031 (2)0.0082 (15)0.0052 (14)0.0077 (16)
C70.0204 (17)0.027 (2)0.029 (2)0.0061 (15)0.0095 (15)0.0031 (15)
C80.0196 (18)0.026 (2)0.032 (2)0.0036 (14)0.0044 (15)0.0080 (16)
C90.0191 (17)0.024 (2)0.0293 (19)0.0114 (14)0.0054 (14)0.0070 (15)
C100.0158 (16)0.029 (2)0.035 (2)0.0062 (14)0.0043 (14)0.0084 (16)
C110.0206 (18)0.034 (2)0.040 (2)0.0090 (16)0.0123 (16)0.0127 (17)
C120.029 (2)0.037 (2)0.030 (2)0.0127 (17)0.0098 (16)0.0119 (17)
C130.0233 (19)0.033 (2)0.032 (2)0.0101 (16)0.0007 (15)0.0054 (17)
C140.0180 (17)0.032 (2)0.036 (2)0.0071 (15)0.0073 (15)0.0097 (17)
F10.0738 (18)0.0462 (16)0.0401 (14)0.0300 (13)0.0083 (12)0.0207 (12)
F20.0573 (16)0.093 (2)0.0404 (15)0.0265 (15)0.0240 (13)0.0268 (14)
F30.0735 (19)0.0528 (17)0.0442 (16)0.0077 (14)0.0159 (13)0.0089 (12)
N10.0147 (14)0.0351 (19)0.0279 (17)0.0076 (13)0.0064 (12)0.0093 (13)
O10.0162 (12)0.062 (2)0.0331 (15)0.0062 (12)0.0074 (11)0.0161 (13)
Geometric parameters (Å, º) top
C1—F31.335 (5)C8—O11.233 (4)
C1—F21.340 (4)C8—N11.370 (4)
C1—F11.340 (4)C8—C91.493 (5)
C1—C21.484 (5)C9—C141.384 (5)
C2—C31.391 (5)C9—C101.405 (5)
C2—C71.397 (5)C10—C111.385 (5)
C3—C41.373 (5)C10—H100.9500
C3—H30.9500C11—C121.378 (5)
C4—C51.396 (5)C11—H110.9500
C4—H40.9500C12—C131.391 (5)
C5—C61.403 (5)C12—H120.9500
C5—N11.409 (4)C13—C141.387 (5)
C6—C71.370 (5)C13—H130.9500
C6—H60.9500C14—H140.9500
C7—H70.9500N1—H10.993 (10)
F3—C1—F2105.9 (3)O1—C8—N1122.7 (3)
F3—C1—F1106.1 (3)O1—C8—C9122.0 (3)
F2—C1—F1105.7 (3)N1—C8—C9115.3 (3)
F3—C1—C2112.7 (3)C14—C9—C10119.3 (3)
F2—C1—C2113.1 (3)C14—C9—C8117.8 (3)
F1—C1—C2112.7 (3)C10—C9—C8122.9 (3)
C3—C2—C7119.0 (3)C11—C10—C9119.4 (3)
C3—C2—C1120.0 (3)C11—C10—H10120.3
C7—C2—C1120.9 (3)C9—C10—H10120.3
C4—C3—C2120.7 (3)C12—C11—C10121.0 (3)
C4—C3—H3119.7C12—C11—H11119.5
C2—C3—H3119.7C10—C11—H11119.5
C3—C4—C5120.5 (3)C11—C12—C13119.8 (3)
C3—C4—H4119.8C11—C12—H12120.1
C5—C4—H4119.8C13—C12—H12120.1
C4—C5—C6118.9 (3)C14—C13—C12119.7 (3)
C4—C5—N1117.8 (3)C14—C13—H13120.1
C6—C5—N1123.3 (3)C12—C13—H13120.1
C7—C6—C5120.4 (3)C9—C14—C13120.8 (3)
C7—C6—H6119.8C9—C14—H14119.6
C5—C6—H6119.8C13—C14—H14119.6
C6—C7—C2120.6 (3)C8—N1—C5127.0 (3)
C6—C7—H7119.7C8—N1—H1115 (3)
C2—C7—H7119.7C5—N1—H1117 (3)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
C6—H6···O10.952.442.938 (4)112
C4—H4···O1i0.952.573.240 (4)128
N1—H1···O1i0.99 (1)2.23 (2)3.138 (3)151 (3)
Symmetry code: (i) x1, y, z.
N-(4-Methoxyphenyl)benzamide (2) top
Crystal data top
C14H13NO2Dx = 1.332 Mg m3
Dm = 1.29 (2) Mg m3
Dm measured by flotation in aqueous KI
Mr = 227.25Melting point = 424–425 K
Monoclinic, P21/cMo Kα radiation, λ = 0.71073 Å
a = 26.7830 (15) ÅCell parameters from 5514 reflections
b = 5.2477 (3) Åθ = 3.1–26.0°
c = 8.1343 (5) ŵ = 0.09 mm1
β = 97.594 (2)°T = 173 K
V = 1133.24 (11) Å3Regular parallelpiped, colourless
Z = 40.70 × 0.26 × 0.08 mm
F(000) = 480
Data collection top
Bruker APEXII CCD
diffractometer
2333 independent reflections
Radiation source: sealed X-ray tube1678 reflections with I > 2σ(I)
Detector resolution: 8.53 pixels mm-1Rint = 0.106
rotating crystal scansθmax = 26.4°, θmin = 1.5°
Absorption correction: multi-scan
(SADABS; Bruker, 2018)
h = 3333
Tmin = 0.518, Tmax = 1k = 66
36318 measured reflectionsl = 1010
Refinement top
Refinement on F2Hydrogen site location: structure-invariant direct methods
Least-squares matrix: fullH atoms treated by a mixture of independent and constrained refinement
R[F2 > 2σ(F2)] = 0.049 w = 1/[σ2(Fo2) + (0.0247P)2 + 0.8083P]
where P = (Fo2 + 2Fc2)/3
wR(F2) = 0.126(Δ/σ)max = 0.001
S = 1.08Δρmax = 0.20 e Å3
2333 reflectionsΔρmin = 0.19 e Å3
160 parametersExtinction correction: SHELXL-2018/3 (Sheldrick, 2015b), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4
1 restraintExtinction coefficient: 0.0099 (18)
Primary atom site location: structure-invariant direct methods
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Refinement. isotropic extinction correction applied

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
C10.53199 (8)0.7401 (5)0.1647 (3)0.0544 (7)
H1A0.5292490.7452500.2835570.082*
H1B0.4982950.7247880.1017860.082*
H1C0.5479520.8971140.1323790.082*
C20.61069 (7)0.5196 (4)0.2084 (2)0.0350 (5)
C30.63993 (8)0.3201 (4)0.1635 (3)0.0388 (5)
H30.6257950.1976600.0847490.047*
C40.68972 (8)0.2992 (4)0.2334 (3)0.0369 (5)
H40.7096050.1611540.2034340.044*
C50.71075 (7)0.4793 (4)0.3471 (2)0.0319 (4)
C60.68127 (8)0.6755 (4)0.3937 (3)0.0362 (5)
H60.6952840.7963020.4737470.043*
C70.63122 (7)0.6971 (4)0.3240 (3)0.0361 (5)
H70.6111370.8332350.3556790.043*
C80.79514 (7)0.6427 (4)0.4517 (3)0.0367 (5)
C90.84768 (7)0.5666 (4)0.5223 (3)0.0343 (5)
C100.85869 (8)0.3512 (4)0.6189 (3)0.0412 (5)
H100.8323430.2402780.6408840.049*
C110.90810 (9)0.2974 (5)0.6833 (3)0.0493 (6)
H110.9154650.1514210.7513180.059*
C120.94644 (9)0.4548 (5)0.6491 (3)0.0502 (6)
H120.9802940.4161640.6922920.060*
C130.93588 (8)0.6682 (5)0.5525 (3)0.0495 (6)
H130.9624510.7764710.5287360.059*
C140.88664 (8)0.7252 (4)0.4899 (3)0.0417 (5)
H140.8794400.8736780.4242170.050*
N10.76211 (6)0.4491 (3)0.4173 (2)0.0358 (4)
O20.56174 (5)0.5266 (3)0.13011 (19)0.0477 (4)
O10.78412 (6)0.8674 (3)0.4262 (2)0.0572 (5)
H10.7748 (8)0.277 (2)0.424 (3)0.041 (6)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
C10.0349 (12)0.0616 (16)0.0642 (17)0.0071 (11)0.0026 (11)0.0003 (13)
C20.0296 (10)0.0395 (11)0.0348 (11)0.0032 (9)0.0001 (8)0.0042 (9)
C30.0404 (12)0.0366 (12)0.0377 (12)0.0039 (9)0.0014 (9)0.0053 (9)
C40.0365 (11)0.0321 (11)0.0420 (12)0.0014 (9)0.0053 (9)0.0020 (9)
C50.0310 (10)0.0309 (10)0.0334 (11)0.0017 (8)0.0024 (8)0.0033 (8)
C60.0347 (11)0.0341 (11)0.0382 (12)0.0002 (9)0.0010 (9)0.0050 (9)
C70.0323 (11)0.0356 (11)0.0402 (12)0.0021 (9)0.0044 (9)0.0026 (9)
C80.0331 (11)0.0313 (11)0.0447 (13)0.0016 (9)0.0020 (9)0.0001 (9)
C90.0321 (11)0.0316 (10)0.0380 (11)0.0013 (8)0.0007 (9)0.0042 (9)
C100.0410 (12)0.0359 (12)0.0449 (13)0.0010 (9)0.0010 (10)0.0002 (10)
C110.0505 (14)0.0432 (13)0.0504 (14)0.0103 (11)0.0079 (11)0.0015 (11)
C120.0349 (12)0.0584 (15)0.0539 (15)0.0113 (11)0.0066 (10)0.0111 (12)
C130.0336 (12)0.0557 (15)0.0576 (15)0.0045 (11)0.0009 (10)0.0062 (12)
C140.0365 (11)0.0394 (12)0.0479 (13)0.0023 (9)0.0011 (10)0.0016 (10)
N10.0297 (9)0.0281 (9)0.0479 (11)0.0017 (7)0.0014 (7)0.0007 (8)
O20.0301 (8)0.0552 (10)0.0548 (10)0.0004 (7)0.0060 (7)0.0063 (8)
O10.0376 (9)0.0302 (8)0.0996 (14)0.0016 (7)0.0063 (9)0.0055 (9)
Geometric parameters (Å, º) top
C1—O21.424 (3)C8—O11.227 (2)
C1—H1A0.9800C8—N11.352 (3)
C1—H1B0.9800C8—C91.502 (3)
C1—H1C0.9800C9—C101.386 (3)
C2—O21.380 (2)C9—C141.387 (3)
C2—C71.385 (3)C10—C111.387 (3)
C2—C31.385 (3)C10—H100.9500
C3—C41.383 (3)C11—C121.375 (3)
C3—H30.9500C11—H110.9500
C4—C51.389 (3)C12—C131.376 (3)
C4—H40.9500C12—H120.9500
C5—C61.381 (3)C13—C141.382 (3)
C5—N11.426 (2)C13—H130.9500
C6—C71.389 (3)C14—H140.9500
C6—H60.9500N1—H10.965 (10)
C7—H70.9500
O2—C1—H1A109.5O1—C8—C9120.83 (18)
O2—C1—H1B109.5N1—C8—C9115.72 (17)
H1A—C1—H1B109.5C10—C9—C14119.21 (19)
O2—C1—H1C109.5C10—C9—C8123.50 (19)
H1A—C1—H1C109.5C14—C9—C8117.28 (19)
H1B—C1—H1C109.5C9—C10—C11120.1 (2)
O2—C2—C7124.29 (19)C9—C10—H10119.9
O2—C2—C3115.73 (18)C11—C10—H10119.9
C7—C2—C3119.98 (18)C12—C11—C10120.1 (2)
C4—C3—C2119.97 (19)C12—C11—H11119.9
C4—C3—H3120.0C10—C11—H11119.9
C2—C3—H3120.0C11—C12—C13120.1 (2)
C3—C4—C5120.28 (19)C11—C12—H12119.9
C3—C4—H4119.9C13—C12—H12119.9
C5—C4—H4119.9C12—C13—C14120.0 (2)
C6—C5—C4119.58 (18)C12—C13—H13120.0
C6—C5—N1121.95 (18)C14—C13—H13120.0
C4—C5—N1118.42 (18)C13—C14—C9120.4 (2)
C5—C6—C7120.34 (19)C13—C14—H14119.8
C5—C6—H6119.8C9—C14—H14119.8
C7—C6—H6119.8C8—N1—C5124.71 (17)
C2—C7—C6119.81 (19)C8—N1—H1118.4 (13)
C2—C7—H7120.1C5—N1—H1116.2 (13)
C6—C7—H7120.1C2—O2—C1116.87 (17)
O1—C8—N1123.44 (19)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
C6—H6···O10.952.492.912 (2)107
N1—H1···O1i0.96 (1)2.16 (1)3.108 (2)166 (2)
Symmetry code: (i) x, y1, z.
π-stacking parameters for TFMP and MOP top
All distances are in Å with estimated uncertainties of 0.004. Angles are in ° with estimated uncertainties of 0.2.
CentroidNormalOffsetTwist angle
TFMP – surrounding both rings
4.7744.6720.98259.7
4.7184.6490.80459.7
4.7114.6460.78059.7
4.6984.6110.90059.7
5.3612.6664.6510.0
MOP – surrounding phenyl rings
4.7814.7570.47864.6
4.9014.8750.50464.6
5.2482.8024.4370.0
MOP – surrounding methoxyphenyl rings
4.8494.6581.34868.1
4.8314.641.34568.1
5.2482.9384.3490.0
 

Footnotes

Retired.

Acknowledgements

The authors thank Megan Mohadjer Beromi of the Chemistry Department at USNA for reviewing the manuscript and offering helpful insights. We also thank Chip Nataro of Lafayette College for performing the CSD search.

Funding information

Funding for this research was provided by: Office of Naval Research (award No. N0001421WX01619 to DWLC); James Kinnear (USNA Class of 1950) (gift to SL, AHRM); Naval Academy Research Council (grant to SL, AHRM); DoD High Performance Modernization Program (award No. USNAM07923005 to JJU); USNA Faculty Enhancement Center (award to JJU, AHRM, and SL); USNA Faculty Development Fund (JJU, SL, and AHRM).

References

First citationBanerjee, A., Saha, A. & Saha, B. K. (2019). Cryst. Growth Des. 19, 2245–2252.  Web of Science CrossRef CAS Google Scholar
First citationBlake, C. C. F. & Small, R. W. H. (1972). Acta Cryst. B28, 2201–2206.  CSD CrossRef IUCr Journals Web of Science Google Scholar
First citationBruker (2018). APEX3, SAINT and SADABS. Bruker AXS Inc, Madison, Wisconsin, USA.  Google Scholar
First citationChang, R. K., Clairmont, B. P., Lin, S. & MacArthur, A. H. R. (2019). Organometallics, 38, 4448–4454.  Web of Science CrossRef CAS Google Scholar
First citationCruz-Cabeza, J. & Bernstein, J. (2014). Chem. Rev. 114, 2170–2191.  Web of Science CAS PubMed Google Scholar
First citationEvano, G., Blanchard, N. & Toumi, M. (2008). Chem. Rev. 108, 3054–3131.  Web of Science CrossRef PubMed CAS Google Scholar
First citationEvano, G., Schaus, J. V. & Panek, J. S. (2004). Org. Lett. 6, 525–528.  Web of Science CrossRef PubMed CAS Google Scholar
First citationFarrugia, L. J. (2012). J. Appl. Cryst. 45, 849–854.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationFrisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Mennucci, B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratchian, H. P., Izmaylov, A. F., Bloino, J., Zheng, G., Sonnenberg, J. L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery, J. A., Jr., Peralta, J. E., Ogliaro, F., Bearpark, M., Heyd, J. J., Brothers, E., Kudin, K. N., Staroverov, V. N., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Rega, N., Millam, J. M., Klene, M., Knox, J. E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E., Yazyev, O., Austin, A. J., Cammi, R., Pomelli, C., Ochterski, J. W., Martin, R. L., Morokuma, K., Zakrzewski, V. G., Voth, G. A., Salvador, P., Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, Ö., Foresman, J. B., Ortiz, J. V., Cioslowski, J. & Fox, D. J. (2016). GAUSSIAN 09, Revision A. 02. Gaussian Inc., Wallingford CT, USA.  Google Scholar
First citationGreenberg, A., Brenemanm, C. M. & Liebman, J. F. (2000). Amide Linkage: Selected Structural Aspects in Chemistry, Biochemistry, and Materials Science. New York: Wiley-Interscience.  Google Scholar
First citationGroom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171–179.  Web of Science CrossRef IUCr Journals Google Scholar
First citationJayatilaka, D. & Grimwood, D. J. (2003). Computational Science|ICCS 2003, edited by P. M. A. Sloot, D. Abramson, A. V. Bogdanov, Y. E. Gorbachev, J. J. Dongarra & A. Y. Zomaya, pp. 142–151. Berlin, Heidelberg: Springer.  Google Scholar
First citationMacrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226–235.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMasse, C. E., Yang, M., Solomon, J. & Panek, J. S. (1998). J. Am. Chem. Soc. 120, 4123–4134.  Web of Science CrossRef CAS Google Scholar
First citationPlessis, M. P. du, Modro, T. A. & Nassimbeni, L. R. (1983). J. Crystallogr. Spectrosc. Res. 13, 179–189.  CSD CrossRef Google Scholar
First citationSatyanarayana, K., Srinivas, K., Himabindu, V. & Reddy, G. M. (2007). Org. Process Res. Dev. 11, 842–845.  Web of Science CrossRef CAS Google Scholar
First citationSeward, N. & Jakubke, H. D. (2002). Peptides: Chemistry and Biology. Weinheim: Wiley-VCH.  Google Scholar
First citationSheldrick, G. M. (2015a). Acta Cryst. A71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015b). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSpackman, P. R., Turner, M. J., McKinnon, J. J., Wolff, S. K., Grimwood, D. J., Jayatilaka, D. & Spackman, M. A. (2021). J. Appl. Cryst. 54, 1006–1011.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSpek, A. L. (2020). Acta Cryst. E76, 1–11.  Web of Science CrossRef IUCr Journals Google Scholar
First citationTanner, D., Fitzgerald, J. A. & Phillips, B. R. (1989). Angew. Chem. Int. Ed. Engl. 28, 649–654.  CrossRef Web of Science Google Scholar
First citationTothadi, S. & Desiraju, G. R. (2012). Cryst. Growth Des. 12, 6188–6198.  Web of Science CSD CrossRef CAS Google Scholar
First citationTurner, M. J., Grabowsky, S., Jayatilaka, D. & Spackman, M. A. (2014). J. Phys. Chem. Lett. 5, 4229–4255.  Web of Science CrossRef Google Scholar
First citationWang, J.-L., Xu, J.-S., Wang, D.-Y., Wang, H., Li, Z.-T. & Zhang, D.-W. (2014). CrystEngComm, 16, 2078–2084.  Web of Science CSD CrossRef CAS Google Scholar
First citationWavefunction (2014). SPARTAN. Wavefunction Inc., Irvine, CA, USA.  Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationZhao, Y. & Truhlar, D. J. (2008). Theor. Chem. Acc. 120, 215–241.  Web of Science CrossRef CAS Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890
Follow Acta Cryst. E
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds