From Cavities to Channels in Host:Guest ... - ACS Publications

Crystal Growth & Design. Advanced Search .... Publication Date (Web): April 1, 2016 ... Upon solvent removal, the porous crystal structure of the host...
1 downloads 0 Views 5MB Size
Subscriber access provided by UNIV OF CALIFORNIA SAN DIEGO LIBRARIES

Article

From cavities to channels in host:guest complexes of bridged trianglamine and aliphatic alcohols Agnieszka Janiak, Mateusz Bardzinski, Jacek Gawronski, and Urszula Rychlewska Cryst. Growth Des., Just Accepted Manuscript • DOI: 10.1021/acs.cgd.6b00100 • Publication Date (Web): 01 Apr 2016 Downloaded from http://pubs.acs.org on April 1, 2016

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.

Crystal Growth & Design 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 33

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

Crystal Growth & Design

From cavities to channels in host:guest complexes of bridged trianglamine and aliphatic alcohols Agnieszka Janiak,* Mateusz Bardziński, Jacek Gawroński and Urszula Rychlewska Department of Chemistry, Adam Mickiewicz University, Umultowska 89B, 61-614 Poznań, Poland.

KEYWORDS. Host-guest complexes, trianglamine, X-ray diffraction, microporosity

ABSTRACT. We demonstrate that covalently bonded chiral organic hexaamine, rigidified by methylene bridge assembles into crystalline inclusion compounds and microporous materials. The inclusion of primary alcohol molecules belonging to the homologous series from ethanol to n-octanol is intrinsic to the triangular molecular shape and columnar stacking of these triangular units. Our studies show that through the choice of differently sized guest molecules (short- or long-chain) we were able to increase the accessible solvent volume from cavities, located inside columnar stacking of triangular units, to one-dimensional undulating channels. While the shortchain molecules occupy voids, the long-chain molecules are included in the channels. By using branched chain solvent during crystallization we were able to isolate dimorphic apohost forms that give rise to two different types of new materials: porous crystals with 1D channels and crystals that possess isolated 0D pockets.

ACS Paragon Plus Environment

1

Crystal Growth & Design

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 2 of 33

INTRODUCTION Organic molecular materials with permanent porosity are still rare compared to materials with extended porous networks because molecules held together in crystals mainly by dispersion forces tend to pack in the efficient manner and leave as little empty space as possible in the solid state.[1-5] On the other hand, organic frameworks self-assembled from either oddly shaped molecules or from organic building blocks that are formed via strong and directional noncovalent interactions often contain channels or cavities that are filled by solvent molecules. Upon desolvation, they undergo a natural tendency to transformation or collapse to the non-porous solids.[6-8] This phenomenon is usually rationalized by taking into account that non-covalent interactions are typically too weak to stabilize the open molecular framework and to establish permanent porosities. However, there are also numerous groups of organic compounds known to maintain porous framework structure after guest removal: their infinite channels or isolated voids, both are enabling the removal of original and the transport of ‘new’ guest molecules through the host lattice.[2-5,9-15] Such group of compounds has been classified by Barbour as permanent or transient porous materials.[2] Among them are calix[4]aren derivatives,[16-20] cucurbit[n]urils,[21-23] some dipetides,[24] or Dianin’s compounds.[25-27] Recently, there has been a significant progress in the development of organic porous materials derived from macrocyclic Schiff bases due to their diversity with respect to size, geometry and functionality.[28-38] A general method for the synthesis of chiral, symmetric polyimine macrocyclic products of triangular,[39-49] rectangular,[50-52] rhombic[53-58] or vase-like shape [59-66] is based on cyclocondensation reaction between vicinal diamine and various unsubstituted or substituted aromatic and aliphatic dialdehydes. A [3+3] cyclocondensation reaction of (1R,2R)-diaminocyclohexane and aromatic dicarboxaldehydes, first suggested by

ACS Paragon Plus Environment

2

Page 3 of 33

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

Crystal Growth & Design

Gawroński and co-workers,[39] provided a new diverse class of imine macrocycles based on dynamic covalent bond formation[39-49,67-69] and offered a great promise in the generation of a new class of chiral inclusion compounds or cage-type materials.[70-77] To make the imine macrocycles chemically stable, they have been reduced to amines.[40] Significantly, polyamine macrocycles show receptor properties[78] and have been used as chiral derivatizing agents in NMR spectroscopy [79-81] as well as chiral ligands in asymmetrically catalyzed addition reactions.[82-86] However, the amination process resulted in an increased macrocycle flexibility [40] and in a complete loss of permanent solid-state porosity of the cage-type molecules.[87] Rigidification of such polyamines has been achieved by either full protonation or by joining together by the methylene or aminol linkers the macrocycle or cage vertices consisting of diamine fragments.[40,87] Tied in this way by the methylene bridge, the chiral trianglamine macrocycle shows an enhanced rigidity and higher symmetry compared to its parent hexaamine, and an ability to include ethanol molecule in its inner cavity.[40] This finding has prompted us to investigate the possible inclusion of aliphatic alcohol molecules belonging to a homologous series from methanol to octanol and a prospect of obtaining the apohost crystal(s). Herein we demonstrate that recrystallization of triangular-shaped chiral hexaamine macrocycle from unbranched monoalcohols provides characteristic inclusion compounds while recrystallization from branched alcohol solvents gives rise to dimorphic apohost crystals. The inclusion properties of this material are a consequence of the presence of fixed, prefabricated intramolecular voids and from a self-assembly process. We anticipate that our findings will stimulate the development of new materials showing well defined inclusion and porosity properties, and the possibility of interconversion between the two states.

ACS Paragon Plus Environment

3

Crystal Growth & Design

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 4 of 33

EXPERIMENTAL SECTION The bridged-trianglamine macrocycle (Scheme 1) was synthesized according to a known synthetic procedure[40] and was initially recrystallized from dichloromethane. Crystals suitable for single-crystal X-ray diffraction experiments were obtained by spontaneous slow evaporation from a series of homologous aliphatic alcohols under ambient conditions. Single-Crystal X-ray Diffraction. Reflection intensities for all of the crystals, except for 1n-PentOH and 1DCM, were measured on a SuperNova diffractometer equipped with Cu microfocus source (λ=1.54178 Å) and 135 mm Atlas CCD detector. The data for remaining crystals were collected with a KM4CCD kappa-geometry diffractometer equipped with a graphite monochromator and MoKα radiation (λ=0.71073 Å). Data reduction and analysis for all of the crystals were carried out with the CrysAlisPro program.[88] In all experiments the sample temperature was controlled with an Oxford Instruments Cryosystem cold nitrogen-gas blower. The structures were solved by direct methods using SHELXS-97[89] or SIR-2011,[90] and refined by the full-matrix least-squares techniques with SHELXL-2014.[91] During the refinement of structure 1n-HexOH, the ROTAX[92] procedure was used to identify the relationship between the two domains of the twinned crystal. This was expressed by the matrix (100, 0-10, 0.529 0-1), which corresponds to a rotation of 180° about the [100] direct lattice direction. Subsequent refinement indicated that the twin fraction of the second domain was 0.200(2). The program WinGX[93] was used to generate the HKLF5 file based on the twin matrix. All heavy atoms were refined anisotropically, except for one of the three positions of dichloromethane (DCM) molecules in 1DCM, octanol molecule in 1n-OctOH and fraction occupancy of water molecules in 1n-HexOH and 1n-HeptOH which were refined isotropically. The hydrogen atoms of the

ACS Paragon Plus Environment

4

Page 5 of 33

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

Crystal Growth & Design

macrocycles were placed at calculated positions and refined using the riding model, and their isotropic displacement parameters were assigned a value 20% higher than the isotropic equivalent for the atom to which they were attached. Owing to the disorder of the included solvent molecules, not all H atoms contained in these molecules have been located. The high wR2 value in the most of inclusion crystals may be related to guest disorder and comparatively poor data quality, resulting from rapid loss of solvent. This event was particularly dramatic in 1n-BuOH, crystals but recrystallization and recollection of the diffraction intensities did not result in any significant improvement the quality of the data. The occupancy of the guest molecule in the previously reported room-temperature crystal structure of 1EtOH was refined to 0.67, hence the host:guest ratio was established as 3:2.[40] Herein we report the results obtained from a new data-set collected for the freshly made 1EtOH crystals at 130K. No phase transition upon cooling was observed but the newly established host:guest ratio has been established as 1:1. In 1EtOH, the ethanol molecules display a disorder where carbon chain is split into two positions with final occupation factors of 0.65 and 0.35 while hydroxyl oxygen atom is disordered over three positions with final occupancy factors of 0.35, 0.35 and 0.3. In the crystal structure of 1n-PrOH the propanol molecule was disordered over two positions that differ in the conformation of the aliphatic chain (O-C-C-C). The final occupancy factors were 0.66 and 0.34 for trans and gauche conformations, respectively. In 1nBuOH

the hydroxyl group displays a disorder where oxygen atom is split into two positions

situated on the opposite site of the carbon chain with the final occupation factors of 0.65 and 0.35. In 1i-BuOH each of the two molecules of iso-butanol that occupy two different structural voids was found to be disordered over two positions and refined with fixed occupancies of 0.6 and 0.4 for one molecule and 0.5 for the other. In 1n-PentOH and 1n-HexOH the disordered regions

ACS Paragon Plus Environment

5

Crystal Growth & Design

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 6 of 33

include hydroxyl group that could be split into three positions with site occupation factors refined to 0.25, 0.35 and 0.4 in 1n-PentOH and 0.2, 0.4 and 0.4 in 1n-HexOH. In 1n-HeptOH the solvent molecule displays a disorder where three carbon atoms are split over two positions while the remaining four carbon atoms are common. The hydroxyl group is also disordered over two sites. The occupancies for the two positions of O and C were refined to 0.5 each. In the crystal structures of 1i-BuOH, 1n-PentOH, 1n-HexOH and 1n-HeptOH the positions of water molecules were determined by analyzing the contacts between the significant peaks of electron density and the closest hydroxy oxygen (in 1i-BuOH) or nitrogen atom of the amine group (in 1n-PentOH, 1n-HexOH and 1n-HeptOH). We determined site occupancies for water molecules as 100% in 1i-BuOH, 20% in 1n-PentOH and 1n-HexOH, and 25% in 1n-HeptOH. Disordered n-octanol molecules in 1n-OctOH were modeled under the assumption that the repeat period of these guest molecules spans three unit cells in the [001] direction. In 1DCM dichloromethane molecule displays a disorder where chlorine atoms are split into two positions while carbon atom is common. The occupancies of the two positions were refined to 0.67 and 0.16, respectively, summing up to the occupancy of 0.83. Also DCM atoms were found in the third position with the site occupancy factor of 0.17. In the crystal structure of 1ap1_EtOH the ethanol molecule was disordered over two positions of opposite orientation. The final occupancy factors were 0.55 and 0.45. In the crystal structures of 1i-PrOH and 12-BuOH solvent molecules were highly disordered and could not be properly modelled therefore their contributions were removed from the diffraction data using SQUEEZE as implemented in Platon.[94] The estimated electron count in 1i-PrOH is 123 in an accessible void volume of 747.0 Å3 and can indicate squeezing of 4 molecules of iso-propanol per unit cell (123/34~4). Taking into account the number of solvent molecules which were squeezed we determined the host to guest ratio as 2:1. The estimated electron count in 12-BuOH is 109 in an

ACS Paragon Plus Environment

6

Page 7 of 33

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

Crystal Growth & Design

accessible void volume of 422.5 Å3 and is correlated with approximately two molecules of 2butanol per unit cell (109/42~2), hence the host:guest ratio is 1:1. The site occupancy factors for the solvent molecules in all inclusion crystals were first based on the indications of their atomic displacement parameters and then were confirmed by the electron density summation in SQUEEZE. Where necessary, restraints for the 1,2- and 1,3-distances as well as the ADP restraint (SIMU) and rigid-bond restraint (DELU) were applied. In cases where the value of the Flack parameter[95] was meaningless, the absolute structure of the investigated crystals was assumed from the known absolute configuration of the (R,R)-1,2-diaminecyclohexane which was used as a starting material in the syntheses. Graphical images were produced in Xseed[96] using Pov-Ray[97] and Mercury[98] programs. The solvent accessible volumes shown as pink Connolly surfaces[99] were calculated using a probe radius of 1.5 Å. The relevant crystal data and refinement parameters are listed in Tables 1 and 2.

Powder X-ray Diffraction (PXRD). X-ray powder diffractogram for the apohost phase of 1 in the 2θ range 5−40° was collected at 295K on a PANalytical Empyrean diffractometer equipped with the PIXcel3D detector and X-ray focusing mirror with an anti-scatter device. The sample was loaded in a standard sample holder for transmission using Kapton foils. The data were collected in Bragg–Brentano fixed sample theta-theta geometry, using CuKα radiation. The apohost phase was verified by comparing the experimental PXRD pattern with those simulated on the basis of single crystal data for phases 1 and 2. Nuclear Magnetic Resonance spectroscopy (NMR). NMR spectra were measured on a Bruker Ascend 400 MHz NanoBay spectrometer for 1H in CDCl3 while TMS was used as

ACS Paragon Plus Environment

7

Crystal Growth & Design

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 8 of 33

internal standard. Each sample was dried using filter paper and dissolved in CDCl3. Relative integrations of appropriate signals correspond to the solvent content in the crystals. The small differences between NMR and other data are caused by wide-ranging solvent volatility and inability to distinguish in this method the amount of solvent on the crystal surface from the amount incorporated into the crystal (see the SI file, where all NMR spectra available have been deposited). Thermogravimetric analysis (TGA). TGA was carried out using a TA Instruments Discovery Series thermogravimetric analyzer. The sample was heated at a constant heating rate of 10 °C/min from room temperature to 450°C. The furnace was purged with N2 gas flowing at a rate of 25 cm3 min-1. Registered mass losses are adequate to determine the solvent content in the crystals (see the SI file, where all TGA curves available have been deposited). The solvent content in the samples from TGA experiments was calculated directly from the percentage of the mass loss. Based on weighing accuracy for Disovery TGA, which is 0.1% and using a type B evaluation of standard uncertainty for one measurement, we have determined the mass percent error of 0.2%.

ACS Paragon Plus Environment

8

Page 9 of 33

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

Crystal Growth & Design

Table 1. Crystal data, structure refinement parameters and H:G ratio for inclusion crystals of 1.

formula

1EtOH

1n-PrOH

1i-PrOH

1n-BuOH

12-BuOH

1i-BuOH

1n-PentOH

1n-HexOH

1n-HeptOH

1n-OctOH

1DCM

C45H60N6* C2H6O

C45H60N6* C3H8O

2(C45H60N6) *C3H8O

2(C45H60N6) *C4H10O

C45H60N6* C4H10O

C45H60N6* C4H10O *0.5(H2O)

2(C45H60N6) *C5H12O *0.4(H2O)

2(C45H60N6) *C6H14O *0.2(H2O)

2(C45H60N6) *C7H16O *0.5(H2O)

5(C45H60N6) *C8H20O

C45H60N6* CH2Cl2

2:1

2:1

2:1

2:1

2:1

5:1

H:G

1:1

1:1

1:1

1:1

1:1

a, b, c [Å]

25.5194(3) 9.2302(1) 17.9734(2)

9.1427(1) 18.2851(2) 25.7424(4)

11.1457(2) 18.0200(2) 42.0267(5)

10.9510(4) 18.1280(4) 42.293(1)

10.9374(2) 14.1707(3) 17.0023(4)

10.9317(2) 14.1485(3) 16.9895(4)

35.415(3) 11.618(1) 22.285(2)

35.0040(6) 11.7755(2) 22.6927(4)

34.928(1) 12.0282(3) 22.494(1)

17.7852(5) 41.4719(15) 5.7519(2)

9.089(2) 18.440(4) 25.594(5)

α, β, γ [°]

90 90 90

90 90 90

90 90 90

90 90 90

107.230(2) 105.589(2) 103.497(2)

107.247(2) 105.800(2) 103.413(2)

90 112.04(1) 90

90 114.094(2) 90

90 114.074(2) 90

90 90 90

90 90 90

V [Å3]

4233.62(8)

4303.49(9)

8440.9(2)

8396.0(4)

2279.73(8)

2270.65(9)

8499.3(13)

8538.8(3)

8628.2(5)

4242.5(2)

4289.6(16)

Space group

P212121

P212121

P212121

P212121

P1

P1

C2

C2

C2

P21212

P212121

T [K]

130

130

150

150

130

130

130

150

150

150

150

Dx [g/cm3]

1.147

1.150

1.078

1.142

0.998

1.123

1.140

1.148

1.148

1.113

1.192

Z, Z’

4, 1

4, 1

4, 1

4, 1

2, 2

2, 2

4, 1

4, 1

4, 1

4, 1

4, 1

F(000)

1592

1624

2976

3144

744

838

3174

3214

3250

1547

1656

R1, wR2 I>2σ(I)

0.0345 0.0926

0.0463 0.1197

0.0516 0.1478

0.0803 0.2302

0.0488 0.1424

0.0538 0.1541

0.0796 0.1774

0.0529 0.1464

0.0612 0.1739

0.0705 0.1954

0.0397 0.0821

GOF on (F2)

1.032

1.018

1.029

1.024

1.042

1.028

1.009

1.027

1.068

1.047

0.916

ρmin ρmax

-0.173 0.165

-0.380 0.307

-0.189 0.6235

-0.341 0.410

-0.395 0.990

-0.246 0.498

-0.287 0.279

-0.402 0.487

-0.309 0.283

-0.253 0.309

-0.186 0.165

ACS Paragon Plus Environment

9

Crystal Growth & Design

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 10 of 33

Table 2. Crystal data and structure refinement parameters for the two apohost phases of 1, namely 1ap_phase1 and 1ap_phase2, and 1ap1_EtOH obtained by soaking1ap_phase1 in ethanol.

formula

1ap_phase1

1ap_phase2

1ap1_EtOH

C45H60N6

C45H60N6

C45H60N6*C2H6O

H:G

-

-

1:1

a, b, c [Å]

17.8822(2) 41.5241(5) 5.6520(1)

9.5207(1) 16.5265(2) 26.1919(4)

17.9156(4) 41.5770(9) 5.6889(1)

α, β, γ [°]

90 90 90

90 90 90

90 90 90

V [Å3]

4196.85(10)

4121.13(9)

4237.53(15)

Space group

P21212

P212121

P21212

T [K]

130

150

150

Dx [g/cm3]

1.084

1.104

1.146

Z, Z’

4, 1

4, 1

4, 1

F(000)

1488

1488

1592

R1, wR2 I>2σ(I)

0.0343 0.0898

0.0338 0.0879

0.0467 0.1250

GOF on (F2)

1.041

1.042

1.030

ρmin ρmax

-0.161 0.031

-0.190 0.128

-0.161 0.203

RESULTS AND DISCUSSION Subsequent crystallization from a series of homologous mono-alcohol molecules provided ten (1EtOH to 1n-octOH) inclusion compounds with guest molecules ranging from ethanol to n-octanol and dichlorometane molecules in place of the expected methanol guests (1DCM), as well as two polymorphic forms of apohost crystals (1ap_phase1 and 1ap_phase2) and the crystals obtained by immersing crystals of 1ap_phase1 in ethanol (1ap1_EtOH). All these compounds have been subjected

ACS Paragon Plus Environment

10

Page 11 of 33

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

Crystal Growth & Design

to low temperature single crystal X-ray diffraction analysis. Selected compounds were also characterized by Thermogravimetric Analysis (TGA) and powder X-ray diffraction (PXRD). Crystal structure data are provided in Tables 1 and 2, while TGA curves are included in the ESI.

Scheme 1. The methylene-bridged trianglamine host molecule (1).

The core structure of bridged trianglamine crystals constituted of triangular-molecules with a mean diameters ranging from 15.2 to 15.4 Å, and containing inner cavities of approximate size ranging from 5.0 to 7.2 Å (see ESI, Table 1S as well as Figs. 18S and 19S for an explanation of how the molecular diameters and cavities were calculated). The top and side views of the most representative conformation of the host molecule are displayed in Fig. 1. Methylene bridging makes the molecule more symmetrical and more rigid compared to its parent unbridged trianglamine. This is demonstrated in making all torsion angles around C-N bonds close to 180° (Table 2S, ESI), while in unbridged trianglamine both both gauche and trans conformations around these bonds were observed due to the formally unrestricted rotation around any of the macrocycle single bonds, the only limitation being the requirement of a closure of the

ACS Paragon Plus Environment

11

Crystal Growth & Design

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 12 of 33

macrocyclic ring. Selected torsion angles describing the bridged trianglamine conformation, as observed in the investigated crystal structures, are accumulated in Table 2S of ESI. As can be seen from this table, for molecules having all R configuration the torsion angles around the C–N bonds are close to 180° and possess a positive sign. The minus signs appear only in crystals solvated with either 2-BuOH or i-BuOH that are, in fact, isostructural. Quite unexpectedly, inversion of the signs of the torsion angles close to 180° makes a noticeable difference in the overall shape of the macrocycle, which becomes less symmetrical than in the other bridged trianglamines. The main difference in the molecular conformation of the host molecules stems from the rotation around the C(sp2)-C(sp3) bonds. This rotation modifies the orientation of the phenyl linker with respect to the macrocycle mean plane. With all three phenyl rings similarly oriented, i.e. nearly perpendicular to the macrocycle, the entrance to the macrocycle cavity is widely open, while it is partially closed from one side of the macrocycle when one or two of the phenyl rings are inclined. This implies that the size of the entrance to the inner cavity depends from which side of the triangular frame it is measured. This is shown in Table 1S of ESI where both sets of values have been accumulated.

ACS Paragon Plus Environment

12

Page 13 of 33

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

Crystal Growth & Design

Figure 1. Top and side view of the host molecule 1.

The bridged-trianglamine host molecule possesses an ability to incorporate into its inner cavity non-branched aliphatic alcohol molecules belonging to the homologous series from ethanol to noctanol. Incorporation of methanol as well as longer than n-octanol chain alcohols has not been so far successful, the reason being the low solubility of trianglamines in these alcohols. In order to get around with this problem it was necessary to dissolve the trianglamine sample in a cosolvent. We have used dichloromethane for this purpose. However, it turned out that dichloromethane molecule competes with the alcohol molecules in the process of entering into the inner cavity of the host leading solely to the formation of 1:1 bridgedtrianglamine:dichloromethane inclusion compound.

ACS Paragon Plus Environment

13

Crystal Growth & Design

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 14 of 33

The principal supramolecular motif observed in all the inclusion compounds, irrespective of the size of the guest molecules and the crystal system or space group is the arrangement of the host molecules in parallel stacks (Fig. 2).

Figure 2. Basic supramolecular motifs consisting of host molecules arranged in parallel stacks and mutual orientation of these stacks (van der Waals representation) in the inclusion compounds with short-chain alcohols (a) and (c), and long-chain alcohols (b) and (d). The host matrices in (c) and (d) are seen along lattice parameters of, respectively, 9.0 Å and 5.6 Å (or an integral multiple of it). H atoms have been omitted for clarity.

With host molecules arranged in stacks, the intramolecular voids are arranged in columns, inside which the aliphatic chain molecules may be adopted as guests. However, depending on the

ACS Paragon Plus Environment

14

Page 15 of 33

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

Crystal Growth & Design

size of the included guest molecules, the macrocycles in a stack are oriented differently with respect to the axis of the column. We can distinguish two classes of crystals formed by primary alcohol molecules, namely those formed by short chain alcohols (and other small molecules, like DCM) and those formed by long-chain alcohol molecules. Short chain molecules pack in channels along which the unit translation is around 9 Å with solvent molecules situated in intramolecular voids. These guest molecules stimulate such arrangement of the host molecules in a stack, in which the stack axis, defined as perpendicular to the macrocycle mean plane, and the column axis do not coincide (Fig. 3). Such an arrangement of columnar stacks is observed in crystals of P212121 symmetry and host:guest ratio of 1:1.

ACS Paragon Plus Environment

15

Crystal Growth & Design

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 16 of 33

Figure 3. Relative orientation of the host and guest molecules (stick and spacefill representation, respectively) in the inclusion compounds: from left to right ethanol, n-propanol and DCM. H atoms of the host have been omitted for clarity.

Meanwhile, longer chain aliphatic alcohols better accommodate in stacks, in which the stack axis forms a small angle with the column axis, so the host molecules forming a column lie nearly perpendicular to the column axis. (Fig. 4). As can be seen from Figure 4, longer-chain alcohols pack in channels along which the repetition period for the host molecules is about 5 Å but for the gust molecules it is an integral multiple of it. In other words, the guest molecules spread over two (Fig. 4) or three (Fig. 5) consecutive host molecules forming inclusion complexes in which the H:G ratio is 2:1 or 5:1, respectively. The two neighboring host molecules are symmetry independent and possess slightly different molecular conformation as has been illustrated in Fig. 6. Inclusion crystals with H:G ratio 2:1 possess either P212121 or C2 symmetry while those with the H:G ratio 5:1 display P21212 symmetry.

Figure 4. Relative orientation of the host and guest molecules in H:G 2:1 inclusion compounds with long-chain alcohols: from left to right n-pentanol, n-hexanol and n-heptanol.

ACS Paragon Plus Environment

16

Page 17 of 33

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

Crystal Growth & Design

Figure 5. Relative orientation of the host and guest molecules in inclusion compound with noctanol, H:G ratio of 5:1.

ACS Paragon Plus Environment

17

Crystal Growth & Design

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 18 of 33

Figure 6. Two conformations of the host molecule 1, present in the crystals of n-BuOH and representative of 1n-PentOH, 1n-HexOH and 1n-HeptOH inclusion crystals. Perpendicular and inclined orientations of the phenyl rings with respect to the macrocycle plane allow opening or partial closing of the entrance to the macrocycle.

Significant difference between the two types of intra-columnar inclusion is that small guest molecules are accommodated in intramolecular voids, while longer-chain alcohol molecules are placed in channels formed inside the columnar stacks. The process of transformation from cavities to channels is illustrated in Fig. 7. It proceeds without any drastic structural changes within the major supramolecular motif but it is preceded by a steady increase of the size of the intramolecular void from 23 through 43 and 50 to 54 Å3. Lastly, worth to note is a disorder of the

ACS Paragon Plus Environment

18

Page 19 of 33

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

Crystal Growth & Design

guest molecules included within the columns. In all Figures only the major component of the disorder is illustrated.

Figure 7. Reorganization of voids ((a), (b), (c) and (d)) into 1D channels ((e), (f) and (g)) stimulated by longer-chain primary alcohol guest molecules. The solvent accessible volumes shown as pink Connolly surfaces [99] were calculated using a probe radius of 1.5 Å.

ACS Paragon Plus Environment

19

Crystal Growth & Design

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 20 of 33

Contrary to the primary alcohol molecules, the branched-chain guests experience serious difficulties in getting into the inner cavity of the host. Only iso-propanol molecules are allowed to enter the host cavity. The topologically identical molecules of iso-butanol and 2-butanol form a new type of inclusion compounds, in which the guest molecules are included in intermolecular voids with a volume of 114 Å3. This type of inclusion is accompanied by significant changes in the arrangement of the host molecules, which no longer pack in stacks but form a sort of a gridlike structure, as shown in Fig. 8. Although 2-butanol is a chiral alcohol we have not observed any separation of its enantiomers by crystallization i.e. both enantiomers are present in the crystals.

Figure 8. Intermolecular inclusion of either iso-butanol or 2-butanol molecules in a host lattice.

Crystallization from tert-butanol and tert-pentanol leads to the formation of the apohost crystals of phase 1 in which the host molecules are arranged in stacks. The same apohost structure is also obtained from iso-propanol and n-butanol by spontaneous slow evaporation of the solvent under ambient conditions for a prolonged period of time. Connected with this process is a change in the crystal symmetry from P212121 to P21212 and the concomitant decrease, by half

ACS Paragon Plus Environment

20

Page 21 of 33

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

Crystal Growth & Design

of its value, of the lattice parameter along the direction of the two-fold axis. This, in turn, is a manifestation of the changes in the relative orientation of the neighbouring stacks, which are no longer shifted with respect to one another. Despite of this reorientation of host stacks the original channel-type structure is retained after desolvation. Moreover, the apohost crystals thus obtained are stable at room temperature for a prolonged period of time. The apohost crystal structure with empty channels is presented in Fig. 9a. It contains one-dimensional undulating pore channel, coloured in pink in Fig. 9a, running through the inner cavities of the host molecules arranged in columnar stacks . The necks of these channels have an approximate diameter of 5.3 Å, whereas the widest part of the channel has a diameter of 7.6 Å. The total accessible solvent area calculated using Platon [94] amounts to 337.7 Å3 per unit cell, which means that approximately 8.0% of the crystal volume is available to guest molecules. Crystallization from other, less common solvents such as dimethylacetamide or ethyl methyl ketone leads to a new polymorphic form of the apohost crystals, which possesses a grid-like structure (Fig. 9b) featuring isolated voids. However, unlike in the apohost phase 1, in the apohost phase 2 the voids do not pass through the inner cavities of the host molecules. The total accessible solvent area amounts to 300 Å3 and corresponds to a value of 7.3% of the unit cell volume, very much the same as in the apohost phase 1. Similarity in the volume accessible for solvent molecules in the two apohost crystals might suggest that the crystals of the apohost phase 2 are also porous. However no attempt was made to provide an evidence of the crystal porosity.[2] Powder X-ray diffraction data correlate well with simulations from single-crystal X-ray diffraction data (Fig. 10) thus demonstrating that the bulk materials were structurally represented by the single crystals chosen for analysis. It should be noted that the apohost X-ray structures were obtained for samples under ambient conditions, rather than for materials that were activated under high vacuum.

ACS Paragon Plus Environment

21

Crystal Growth & Design

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 22 of 33

Figure 9. Packing diagram of the dimorphic apohost crystals with structural voids marked in pink. Apohost with channel-type structure (a) and with grid-like structure (b).

Figure 10. Comparison of experimental (red) and calculated PXRD patterns (blue and green) of apohost crystalline phases.

Furthermore, the crystals of apohost phase 1 were immersed in ethanol at room temperature for 14 days. The single crystal thus obtained has been subjected to X-ray analysis which revealed a

ACS Paragon Plus Environment

22

Page 23 of 33

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

Crystal Growth & Design

new crystalline phase of the ethanol inclusion compound 1ap1_EtOH. The new phase retained the apohost crystal structure while the ethanol guest molecules penetrated the apohost channels forming a new ethanol solvate phase with the host:guest ratio of 1:1 as is illustrated in Fig. 11.

Figure 11. Ethanol molecules migrate through the channel in the phase 1 of the apohost. The H atoms of the host have been omitted for clarity.

Conclusions We have demonstrated that it is possible to direct the type of the guest inclusion and porosity in solely molecular organic material from discrete voids to 1D undulated channels through the judicious choice of the chain alcohol solvent molecules. Furthermore, we have provided evidence of sorption after soaking the empty host with ethanol guest molecules for 14 days. The two packing modes observed in the porous apohost polymorphs (columnar stacking in 1ap_phase1 and grid-like packing in 1ap_phase2) suggest the possibility of materials where the type of porosity can be controlled by means of solvent molecules and/or desolvation process.

ASSOCIATED CONTENT Supporting Information.

ACS Paragon Plus Environment

23

Crystal Growth & Design

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 33

Electronic Supplementary Information (ESI) available: descriptions of the crystal structures, TGA curves and NMR spectra. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author *E-mail: [email protected]

Notes CCDC-1448448-1448461 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data%5Frequest/cif

ACKNOWLEDGMENT The

authors

thank

the

National

Science

Center

in

Poland,

grant

no.

UMO-

2012/06/A/ST5/00230, for financial support. The authors also would like to thank PANalytical, in particular Dr Olga Narygina, for performing PXRD experiment on apohost phase of 1 as well as Artur Strzelecki and TA Instruments for providing us the access to the TGA Discovery Series.

ACS Paragon Plus Environment

24

Page 25 of 33

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

Crystal Growth & Design

REFERENCES 1. Kitajgorodskii, A. I. Molecular Crystals and Molecules; Academic Press, New York, 1973. 2. Barbour, L. J. Chem. Commun., 2006, 1163-1168. 3. Maly, K. E. J. Mater. Chem., 2009, 19, 1781-1787. 4. Holst, J. R.; Trewin, A.; Cooper, A. I. Nature Chemistry, 2010, 2, 915-920. 5. Tian, J.; Thallapally, P. K.; McGrail, B. P. CrystEngComm, 2012, 14, 1909-1919. 6. Wuest, J. D. Chem. Commun., 2005, 5830-5837. 7. Weber, J.; Antonietti, M.; Thomas, A. Macromolecules, 2007, 40, 1299-1304. 8. Thomas, A.; Goettmann, F.; Antonietti, M. Chem. Mater., 2008, 20, 738-755. 9. Mastalerz, M.; Oppel, I. M. Angew. Chem., Int. Ed., 2012, 51, 5252-5255. 10. Luo, X.-Z.; Jia, X.-J.; Deng, J.-H.; Zhong, J.-L.; Liu, H.-J.; Wang, K.-J.; Zhong, D.-C.; J. Am. Chem. Soc., 2013, 135, 11684-11687. 11. Li, P.; He, Y.; Arman, H. D.; Krishna, R.; Wang, H.; Weng, L.; Chen, B. Chem. Commun., 2014, 50, 13081-13084. 12. Li, P.; He, Y.; Guang, J.; Weng, L.; Zhao, J. C.-G.; Xiang, S.; Chen, B. J. Am. Chem. Soc. 2014, 136, 547-549. 13. Li, P.; He, Y.; Zhao, Y.; Weng, L.; Wang, H.; Krishna, R.; Wu, H.; Zhou, W.; O'Keeffe, M.; Han, Y.; Chen, B. Angew. Chem., Int. Ed., 2015, 54, 574-577.

ACS Paragon Plus Environment

25

Crystal Growth & Design

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 26 of 33

14. Wang, H.; Li, B.; Wu, H.; Hu, T.-L.; Yao, Z.; Zhou, W.; Xiang, S.; Chen, B. J. Am. Chem. Soc., 2015, 137, 9963-9970. 15. Taylor, R. G. D.; Bezzu, C. G.; Carta, M.; Msayib, K. J.; Walker, J.; Short, R.; Kariuki, B. M.; McKeown, N. B. Chem. Eur. J. 2016, 22, 2466-2472. 16. Atwood, J. L.; Barbour, L. J.; Jerga, A.; Schottel, B. L. Science, 2002, 298, 1000-1002. 17. Thallapally, P. K.; Lloyd, G. O.; Atwood, J. L.; Barbour, L. J. Angew. Chem., Int. Ed., 2005, 44, 3848-3851. 18. Dalgarno, S. J.; Thallapally, P. K.; Barbour, L. J.; Atwood, J. L. Chem. Soc. Rev., 2007, 36, 236-245. 19. Thallapally, P. K.; McGrail, B. P.; Atwood, J. L.; Gaeta, C.; Tedesco, C.; Neri, P. Chem. Mater., 2007, 19, 3355-3357. 20. Herbert, S. A.; Janiak, A.; Thallapally, P. K.; Atwood, J. L.; Barbour, L. J. Chem. Commun., 2014, 50, 15509-15512. 21. Miyahara, Y.; Abe, K.; Inazu, T. Angew. Chem., Int. Ed. 2002, 41, 3020-3023. 22. Lim, S.; Kim, H.; Selvapalam, N.; Kim, K.-J.; Cho, S. J.; Seo G.; Kim, K. Angew. Chem., Int. Ed., 2008, 47, 3352-3355. 23. Tian, J.; Liu, J.; Liu, J.; Thallapally, P. K. CrystEngComm, 2013, 15, 1528-1531. 24. Comotti, A.; Bracco, S.; Distefano G.; Sozzani, P. Chem. Commun., 2009, 284-286. 25. Barrer, R. M.; Shanson, V. H. J. Chem. Soc., Chem. Commun., 1976, 333-334.

ACS Paragon Plus Environment

26

Page 27 of 33

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

Crystal Growth & Design

26. Imashiro, F.; Yoshimura, M.; Fujiwara, T. Acta Cryst. C, 1998, 54, 1357-1360. 27. Jacobs, T.; Smith, V. J.; Thomas, L. H.; Barbour, L. J. Chem. Commun., 2014, 50, 85-87. 28. Gallant, A. J.; Yun, M.; Sauer, M.; Yeung, C. S.; MacLachlan, M. J.Org. Lett., 2005, 7, 4827-4830. 29. Bru, M.; Alfonso, I.; Burguete, M. I.; Luis, S. V. Angew. Chem.,Int. Ed., 2006, 45, 61556159. 30. Hui, J. K. H.; MacLachlan, M. J. Chem. Commun., 2006, 2480-2482. 31. Boden, B. N.; Hui, J. K.-H.; MacLachlan, M. J. J. Org. Chem. 2008, 73, 8069–8072. 32. Jiang, J.; MacLachlan, M. J. J. Org. Lett. 2010, 12, 1020-1023. 33. Guieu, S.; Crane, A. K.; MacLachlan, M. J. Chem. Commun., 2011, 47, 1169-1171. 34. Guieu, S.; Crane, A. K.; MacLachlan, M. J. Chem. Commun., 2011, 4, 1169–1171. 35. Bru, M.; Alfonso, I.; Bolte, M.; Burguete, M. I.; Luis, S. V., Chem. Commun., 2011, 47, 283-220. 36. Leeland, J. W.; White, F. J.; Love, J. B. J. Am. Chem. Soc. 2011, 133, 7320-7323. 37. Okochi, K. D.; Han, G. S.; Aldridge, I. M.; Liu, Y.; Zhang, W. Org. Lett. 2013, 15, 42964299. 38. Stadler, A.-M.; Jiang, J.-J.; Wang, H.-P.; Bailly, C. Chem. Commun. 2013, 49, 3784-3786. 39. Gawronski, J.; Kołbon, H.; Kwit, M.; Katrusiak, A. J. Org. Chem. 2000, 65, 5768-5773.

ACS Paragon Plus Environment

27

Crystal Growth & Design

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 28 of 33

40. Gawronski, J.; Gawronska, K.; Grajewski, J.; Kwit, M.; Plutecka, A.; Rychlewska, U. Chem. Eur. J, 2006, 12, 1807-1817. 41. Kwit, M.; Plutecka, A.; Rychlewska, U.; Gawronski, J.; Khlebnikov, A. F.; Kozhushkov, S. I.; Rauch, K.; Meijere, A. Chem. Eur. J. 2007, 13, 8688- 8695. 42. Chadim, M.; Budesınsky, M.; Hodacova, J.; Zavada, J.; Junk, P. C. Tetrahedron: Asymmetry, 2001, 12, 127-133. 43. Akine, S.; Taniguchi, T.; Nabeshima, T. Tetrahedron Letters, 2001, 42, 8861–8864. 44. Kuhnert, N.; Lopez-Periago, A. M. Tetrahedron Letters, 2002, 43, 3329-3332. 45. Akine, S.; Hashimoto, D.; Saiki, T.; Nabeshima, T. Tetrahedron Letters, 2004, 45, 42254227. 46. Kuhnert, N.; Patel, Ch.; Jami, F. Tetrahedron Letters, 2005, 46, 7575-7579. 47. Gallant, A. J.; Hui, J. K.-H.; Zahariev, F. E.; Wang, Y. A.; MacLachlan, M. J. J. Org. Chem., 2005, 70, 7936-7946. 48. Shopsowitz, K. E.; Edwards, D.; Gallant, A. J.; MacLachlan, M. J. Tetrahedron, 2009, 65, 8113–8119. 49. Lopez-Periago, A. M.; Garcia-Gonzalez, C. A.; Domingo, C. Chem. Commun. 2010, 46, 4315-4317. 50. Gregolinski, J.; Lisowski, J.; Lis, T. Org. Biomol. Chem., 2005, 3, 3161-3166. 51. Cameron, S. A.; Brooker, S. Inorg. Chem., 2011, 50, 3697-3706.

ACS Paragon Plus Environment

28

Page 29 of 33

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

Crystal Growth & Design

52. Gregolinski, J.; Slepokura, K.; Packowski, T.; Lisowski, J. Org. Lett., 2014, 16, 43724375. 53. Jazwinski, J.; Lehn, J.-M.; Meric, R.; Vigneron, J.-P.; Cesario, M.; Guilhem, J.; Pascard, C. Tetrahedron Lett., 1987, 28, 3489-3492. 54. Li, Z.; Jablonski, C. Chem. Commun., 1999, 1531-1532. 55. Gawronski, J.; Brzostowska, M.; Kwit, M.; Plutecka, A.; Rychlewska, U. J. Org. Chem., 2005, 70, 10147-10150. 56. Srimurugan, B. Viswanathan, T. K. Varadarajana, B. Vargheseb, Org. Biomol. Chem., 2006,4, 3044-3047. 57. Gualandi, A.; Cerisoli, L.; Stoeckli-Evans, H.; Savoia, D. J. Org. Chem., 2011, 76, 33993408. 58. Ide, T.; Takeuchi, D.; Osakada, K. Chem Commun., 2012, 48, 278-280. 59. Korupoju, S. R.; Zacharias, P. S. Chem. Commun., 1998, 12, 1267- 1268. 60. Gao, J.; Reibenspies, J. H.; Zingaro, R. A.; Woolley, F. R.; Martell, A. E.; Clearfield, A. Inorg. Chem., 2005, 44, 232–241. 61. Paluch, M.; Lisowski, J.; Lis, T. Dalton Trans., 2006, 381-388. 62. Chu, Z.; Huang, W.; Wang, L.; Gou, S. Polyhedron, 2008, 27, 1079-1092. 63. Tanaka, K.; Shimoura, R.; Caira, M. R. Tetrahedron Lett., 2010, 51, 449-452. 64. Sarnicka, A.; Starynowicz, P.; Lisowski, J. Chem. Commun.; 2012, 48, 2237-2239.

ACS Paragon Plus Environment

29

Crystal Growth & Design

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 33

65. Janiak, A.; Petryk, M.; Barbour, L. J.; Kwit, M. Org. Biomol. Chem., 2016, 14, 669-673. 66. Petryk,

M.;

Biniek,

K.;

Janiak,

A.;

Kwit,

M.

CrystEgComm,

2016,

doi:

10.1039/c6ce00256k. 67. Lee, S. J.; Lin, W. In Encyclopedia of Nanoscience and Nanotechnology; Nalwa, H. S.,Ed.; American Scientific Publishers, 2001, Vol. 1, pp. 863-875. 68. Borisova, N. E.; Reshetova, M. D.; Ustynyuk, Y. A. . Chem. Rev. 2007, 107, 46-79. 69. Srimurugan, S.; Suresh, P.; Babu, B.; Pati, H. N. Mini-Reviews in Organic Chemistry, 2008, 5, 228-242. 70. Tozawa, T.; Jones, J. T.A.; Swamy, S. I.; Jiang, S.; Adams, D. J.; Shakespeare, S.; Clowes, R.; Bradshaw, D.; Hasell, T.; Chong, S. Y.; Tang, Ch.; Thompson, S.; Parker, J.; Trewin, A.; Bacsa, J.; Slawin, A. M. Z.; Steiner, A.; Cooper, A. I. Nature Materials, 2009, 8, 973978. 71. Mastalerz, M.; Schneider, M. W.; Oppel, I. M.; Presly, O. Angew. Chem., Int. Ed. 2010, 50, 1046-1051. 72. Mastalerz, M.; Schneider, M. W.; Oppel, I. M.; Presly, O. Angew. Chem., Int. Ed., 2011, 50, 1046-1051. 73. Jones, J. T. A.; Hasell, T.; Wu, X.; Bacsa, J.; Jelfs, K. E.; Schmidtmann, M.; Chong, S. Y.; Adams, D. J.; Trewin, A.; Schiffman, F.; Cora, F.; Slater, B.; Steiner, A.; Day, G. M.; Cooper, A. I. Nature, 2011, 474, 367-371. 74. Mastalerz, M. Chem. Eur. J. 2012, 18, 10082-10091.

ACS Paragon Plus Environment

30

Page 31 of 33

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

Crystal Growth & Design

75. Skowronek, P.; Warżajtis, B.; Rychlewska, U.; Gawroński, J. Chem Commun., 2013, 49, 2524-2526. 76. Zhang, G.; Presly, O.; White, F.; Oppel, I. M.; Mastalerz, M. Angew. Chem., Int. Ed. 2014, 53, 1516-1520. 77. Ding, H.; Yang, Y.; Li, B.; Pan, F.; Zhu, G.; Zeller, M.; Yuan, D.; Wang, C. Chem. Commun., 2015, 51, 1976-1979. 78. Alfonso, I. Curr. Org. Synth, 2010, 7, 1-23. 79. Tanaka, K.; Nakai, Y.; Takahashi, H. Tetrahedron: Asymmetry, 2011, 22, 178-184. 80. Periasamy, M.; Dalai, M.; Padmaja, M. J. Chem. Sci. 2010, 122, 561-569. 81. Tanaka, K.; Fukuda, N.; Fujiwara, T. Tetrahedron: Asymmetry, 2007, 18, 2657-2661. 82. Tanaka, K.; Hachiken, S. Tetrahedron Lett. 2008, 49, 2533-2536. 83. Song, W.; Dong, L; Zhou, Y.;Fu, Y.; Xu, W. Synth, Commun. 2015, 45, 70-77. 84. Gajewy, J.; Kwit, M.; Gawronski, J. Adv. Synth. Catal. 2009, 351, 1055-1063. 85. Gajewy, J.; Gawronski, J.; Kwit, M. Org. Biomol. Chem. 2011, 9, 3863-3870. 86. Tanaka, K.; Asakura, A.; Muraoka, T.; Kalicki, P.; Urbanczyk-Lipkowska, Z. New J. Chem. 2013, 37, 2851-2855. 87. Liu, M.; Little, M. A.; Jelfs, K. E.; Jones, J. T. A.; Schmidtmann, M.; Chong, S. Y.; Hasell, Y.; Cooper, A. I. J. Am. Chem. Soc., 2014, 136, 7583-7586. 88. CrysAlisPro, version. 1.171.35.4, Agilent Technologies, Ltd, Yarnton, England, 2010.

ACS Paragon Plus Environment

31

Crystal Growth & Design

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 32 of 33

89. Sheldrick, G. M. Acta Cryst. A, 2008, 64, 112-120. 90. Burla, M. C.; Caliandro, R.; Camalli, M.; Carrozzini, B.; Cascarano, G.L.; Giacovazzo, C.; Mallamo, M.; Mazzone, A.; Polidori, G.; Spagna, R. J. Appl. Cryst. 2012, 45, 351-356. 91. Sheldrick, G.M. Acta Cryst. C, 2015, 71, 3-8. 92. Cooper, R. I., Gould, R. O., Parsons, S., Watkin, D. J. J.Appl. Cryst. 2002, 35, 168–174. 93. Farrugia, L. J. J. Appl. Cryst, 1997,30, 565. 94. Spek, T. L. Acta Cryst. A, 1990, 46, 194. 95. Flack, H. D.; Bernardinelli, G. J. Appl. Cryst. 2000, 33, 1143–1148. 96. Barbour, L. J. J. Supramol. Chem. 2001, 1, 189-191. 97. POV-RayTM for Windows, version 3.6, Persistence of Vision Raytracer Pty Ltd, Williamstown, Australia, 2004, http://www.povray.org 98. Bruno, I. J; Cole, J. C.; Edgington, P. R.; Kessler, M.; Macrae, C. F.; McCabe, P.; Pearson, J.; Taylor, R. Acta Cryst. B, 2002, 58, 389–397. 99. Connolly, M. L. Science, 1983, 221, 709-713.

ACS Paragon Plus Environment

32

Page 33 of 33

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

Crystal Growth & Design

Table of Contents Use Only From cavities to channels in host:guest complexes of bridged trianglamine and aliphatic alcohols Agnieszka Janiak, Mateusz Bardziński, Jacek Gawroński, and Urszula Rychlewska

Chiral, rigid hexaamine exhibits well defined inclusion and porosity properties. Crystallization from unbranched monoalcohols leads to solvated crystals while crystallization form branched alcohols or other non-alcoholic solvents provides dimorphic apohost crystals. Upon solvent removal, the porous crystal structure of the host is retained.

ACS Paragon Plus Environment

33