SlideShare a Scribd company logo
1 of 28
Download to read offline
This journal © The Royal of Dalton r
Dalton
Transactions
REPORT
Received 5th June 2015
www.rsc.org/dalton
Structural Studies on Pyrene-Based Blue-Light Emitting
Molecules and Ion-Sensing Calixarenes
Thomas G. Warwick* and Dr. Mark R. J. Elsegood
The awarding of the physics Nobel Prize to Shuji Nakamura and colleagues for their role in the
development of the blue LED has drawn attention to the field of organic LED, or OLED design, where
there is a keenness to develop blue-light emitting organic substrates with desirable emissive properties.
Calixarenes meanwhile have been an important family of hosts in host-guest chemistry in recent
decades, with well-established synthetic procedures and nomenclature. This wealth of literature makes
calixarenes appealing compounds when considering solutions to new ion-sensing/molecular
recognition-type problems. Single crystal X-ray diffraction experiments can provide absolute structure
determinations of crystalline solids, and is therefore an invaluable technique when the properties of a
material are highly dependent upon their structure, as with the above. Single crystal X-ray diffraction
experiments were performed on six calixarenes and two pyrene-based blue-light emitting materials,
submitted by a collaborator in Japan, and ten additional structures. The results include both structures
which appear to be highly, and poorly optimized for their intended purpose, the reasons for which are
discussed here. Most saliently, irrefutable proof of 2:1 host-to-guest Cl
-
binding behavior for a
thiacalix[4]arene has been discovered, and these results have been included in a 2015 research paper.
Introduction
Calixarenes
The chemistry of calix[n]arenes has been thoroughly explored subsequent to the work of Gutsche and co-workers,
who were the first to tabulate assorted general synthetic routes, their rudimentary chemical properties, and
nomenclature in the late 70s.1
Success in ―mapping‖ out calixarene chemistry met with an abundance of prospective
applications for these new materials and spawned a research area in which Gutsche remains a seminal figure, with
collaborative papers published over a period spanning four decades,2a~b
and more recently, three books collating the
key concepts resulting from this work.3a~c
In the broadest terms then, calixarenes are cyclic oligomers formed by
*Student Number B126328, Department of Chemistry, Loughborough, Leicester, LE11 3UN. Email: t.warwick-11@student.lboro.ac.uk.
Supervisor Email: m.r.j.elsegood@lboro.ac.uk
†Department of Applied Chemistry, Faculty of Science and Engineering, Saga University, Honjo-machi 1, Saga 840-8502, Japan. E-mail:
yamatot@cc.saga-u.ac.jp
†Herein we use R? to denote compounds external to our experiments that are available in the literature. CCDC deposition numbers etc. are
available in appendix ii
†All figures generated by the author in XP interactive molecular graphics, available as part of the SHELXTL suite released under license by
BRUKER
21
†Non-standard structure codes, TY?, are used at the request of the supervisor for the sake of cross-referencing and internal consistency
This journal © The Royal of Dalton r
acid- or base-catalysed condensation of a finite number of simple aromatic alcohols, such as phenol or resorcinol,
with an aldehyde, wherein water is eliminated. [n] is a variable denoting the number of aromatic alcohol monomers
incorporated into the oligomer. The aldehyde, which effectively forms a bridging unit between aromatic alcohols,
may be switched to (say) elemental sulfur, to afford sulphur-bridged thiacalix[n]arenes,4
or to a suitable
dielectrophile to afford oxacalix[n]arenes.5
Additionally, more complex derivatives of the starting aromatic alcohols
may be chosen to yield functionalized calixarene products, with the conformation, as well as the connectivity of these
starting materials affecting the ultimate properties of the products. Calixarenes therefore are highly tuneable
molecules, with the interplay of these controllable parameters offering a means to synthesize calixarenes which are
highly optimized for their purpose. These purposes include various types of catalysis, such as α-olefin
polymerization and lactone ring opening,6
in which calixarenes with novel bridging units, e.g. dimethyleneoxa
(-H2COCH2-), coordinated to vanadium metal centers are employed. Calixarenes bearing Lewis base moieties can
access the d orbitals of transition metal ions/centers,7
whilst the size of the central cavity of the calixarene may be
fine-tuned to optimize specificity towards species of a particular size, and so calixarenes are often employed as ion-
sensors/purification agents.8
These molecular recognition-type properties of calixarenes of course have biomedical
applications too, and recently, phosphonate-derivatized thiacalix[4]arenes were demonstrated to successfully inhibit
Protein Tyrosine Phosphatase, an enzyme which catalyzes the dephosphorylation of phosphotyrosine protein
residues, in bubonic plague-causing Yersinia bacteria, in which these Protein Tyrosine Phosphatases are the primary
cause of the plague symptoms.9
Pyrene-Based Blue Light Emitters
The first evidence of the photovoltaic, or PV, effect occurring in an organic material was published in 1906 by
Pochettino who, inspired by Edmond Becquerel, was performing similar experiments on anthracene crystals.10
The
conversion efficiency was however poor and so the field of photovoltaics remained focused on exploiting the PV
properties of inorganic materials, culminating in the first solar cell in 1954.11
In the mid-80s, Tang reported an
organic PV cell with a relatively high (ca. 1%) conversion efficiency, which was a consequence of the cell‘s unique,
two-layer design.12
Here, two organic layers were placed in-between two electrodes of different work functions, and
excitons, produced by the absorption of photons by one of the layers, dissociated towards the low-work function
cathode, generating a potential difference.13
Of course, Organic Light Emitting Diodes, or OLEDs, operate via the
exact reverse principle; electrons and electron holes are introduced to the organic layers via application of an electric
current, and Coulombic forces bring these together. Once both electrons and electron holes exist within a single layer
(the emissive layer), recombination occurs to produce light with a wavelength inversely proportional to the energy
separation of that material‘s Frontier Molecular Orbitals (its band gap).
With the 2014 Nobel Prize for physics awarded to Shuji Nakamura for his role in the development of the
blue LED,14
research into the development of organic molecules which emit blue light upon application of an electric
This journal © The Royal of Dalton r
field has been brought sharply into focus. Necessarily, blue emission bands are hard to access due to; the large band
gap required to produce high energy light; low-energy excimer emission occurring when π-clouds on adjacent
molecules approach each other at a distance less than ca. 4.5 Å resulting from π···π stacking;15
and homolysis of
excited species whose recombination pathways include non-radiative processes. The latter issue presents the further
condition of having to avoid, where possible, inclusion of weaker carbon-heteroatom bonds in the material‘s
structure. Pyrene is a highly stable Poly Aromatic Hydrocarbon, PAH, with a long fluorescence lifetime of >100ns16
and high quantum yield, Φf, of 0.32 in cyclohexane17
whose emission from the S1 state is red, or bathochromically,
shifted into the blue/visible region of the electromagnetic spectrum by rational substitution. Choice of substituent,
and its location on the pyrene ring can effectively suppress π···π stacking/excimer emission in the solid state, and so
pyrene derivatives are a promising class of compounds for use as emissive materials in blue OLEDs. These OLEDs,
once fully optimized and in sufficient demand, may replace current Liquid Crystal Display technology, reducing the
energy consumption of televisions, monitors and such like by circumventing the need for backlighting. Furthermore,
displays may be printable18
and flexible,19
complimenting research into dye-sensitized solar cells.
Aims
Clearly, success or failure of both classes of materials is intimately connected to their solid state structure, which in
turn affects their solid state properties. Determination of these structures can, in the case of the blue-light emitters,
provide arguments as to why a prospective material emits (say) green or yellow light, by quantifying the distance
between excimer-forming π···π stacked units. Similarly, unanticipated charge-transfer stabilization between proximal
non-hydrocarbon substituents may be simultaneously proven and quantified. For calixarenes, solid state structure
determination is a companion to spectrophotometric and NMR studies in assessing whether a host complex has, or
has not captured a guest species. Moreover, a fully solved structure can yield precisely where host-guest interactions
occur in multi-functional host complexes such as prospective catalysts. Precise data of this kind is essential for
maintaining oversight on the frequent serendipitous results in these fields of research, and the ability to tabulate
unexpected/chance results is arguably the single most beneficial contribution provided by a solid state structure
determination.
As both classes of materials readily form crystalline solids, single crystal X-ray diffraction experiments were
to be performed on a variety of samples representing both classes of materials, provided by Takehiko Yamato† and
co-workers. Each sample was to be subjected to data collection, correction, structure solution, modeling, refinement
and validation in accordance with established crystallographic procedures. Parameters and variables governing these
procedures were to be tabulated and presented as part of a results ―package‖. The crystal structures, and key
structural features, were to be presented diagrammatically, textually and quantitatively, as part of this package.
Adherence to established procedures and in-house styles were to be rigorously upheld to a) provide high-quality
This journal © The Royal of Dalton r
results, and b) provide a benchmark against which improvements in crystallographic knowledge and skill of the
author could be compared.
Experimental
A general experimental procedure representing the most common choice of experimental conditions is given.
Deviations from this procedure for each sample will then be presented sample-wise. The reader is directed to tables
fully detailing experimental conditions for each sample provided in appendix i.
General X-ray Crystallography Procedure
Crystal data were collected at the ALS, station 11.3.1 on a Bruker APEX 2 and CCD diffractometer using silicon 111
monochromated X-radiation with ω-scans.20
Data were corrected for Lorentz and polarization effects and absorption,
based on repeated symmetry-equivalent reflections, and structures solved by dual-space iterative methods.21,22
Structures were refined by full matrix least squares on F2
.21,22
H atoms were geometrically constrained (unless stated
otherwise below) using a riding model, in which all hydrogen atom Uiso values set at 120% of that of the carrier
atom, except for methyl and hydroxyl H atoms (set as 150%). Observed disorder was modeled as multiple sets of
point atoms distributed over discreet locations with an occupancy sum of 1. Restraints were placed on geometries
and Uaniso values of atoms in disordered moieties as appropriate.
Sample-Specific Procedures
In TY44 the chloroform solvent carbon and its adjoining H atom were modeled as fully disordered over two sets of
positions without employing any restraints. In TY49 badly disordered solvent molecules were modeled as diffuse
clouds of electron density using the Platon ―Squeeze‖ procedure.23
This gave a solution of three methanol molecules
and one water molecule. The o-pyridyl group C(53) > N(6) and the t
butyl group C(7) > C(10) were modelled as fully
disordered over two sets of positions for the Me groups. Geometric and Uaniso restraints were used. H(1) and H(3)
were freely refined, whilst H(5) was modelled as two partial point atoms, with each one belonging to one of the o-
pyridyl disorder components. For TY51 the t
butyl group containing C(17) on one pyrene ring was modelled as fully
disordered over two sets of positions for the Me groups using geometric and Uaniso restraints. The badly disordered
methanol molecule was modelled as a diffuse cloud of electron density using the Platon ―Squeeze‖ procedure.23
In
TY53 diffraction data were collected on a Bruker APEX 2 CCD diffractometer using graphite monochromated Mo-
Kα X-radiation with narrow slice, 0.3° ω-scans. The structure was solved by direct methods.21,22
The structure was
heavily disordered overall. Two of the ethanol solvent molecules were modelled as point atoms, both of which are
disordered over two positions. The third ethanol molecule was modelled as a diffuse cloud of electron density using
the Platon ―Squeeze‖ procedure.23
The water molecule was modelled partially as a single, dehydrated point O partial
This journal © The Royal of Dalton r
point atom, and partially as a diffuse cloud of electron density using the squeeze procedure. The point atom
constitutes the major disorder component, with occupancy of ca. 63%. The terminal ethyl ester carbon atom C(49)
also exhibits two-part disorder. Geometric and Uaniso restraints were heavily applied. In TY54 diffraction data were
collected on a Bruker D8 diffractometer, equipped with a PHOTON 100 CMOS detector, with shutterless ω-scans.
Full matrix least squares refinement on F2
was performed as a two-component non-merohedral twin, on ‗twin5‘
data,21,22
with the major component at 53.5(2)%. Rotation is about axis [0 1 -1]. For TY55 diffraction data were
collected on a Bruker D8 diffractometer, equipped with a PHOTON 100 CMOS detector, with shutterless ω-scans.
Again, the structure was badly disordered. The benzyl group containing C(58) > C(64) was modeled as fully
disordered over two positions. The 1-methylimidazole group attached to C(53), and C(53) itself, were also modeled
as fully disordered over two positions. The ethanol molecule containing O(5) was modeled as fully disordered over
three positions, with an occupancy sum restraint employed. Geometric and Uaniso restraints were heavily applied. For
TY57 diffraction data were collected on a Bruker APEX 2 CCD diffractometer using graphite monochromated Mo-
Kα X-radiation with narrow slice, 0.3° ω-scans. The structure was solved by direct methods.21,22
NH atoms were
freely refined using geometric restraints. In TY59 diffraction data were collected on a Bruker D8 diffractometer,
equipped with a PHOTON 100 CMOS detector, with shutterless ω-scans. Two t
butyl groups on calixarenes were
modelled as disordered over two sets of positions for the Me groups. Two n
butyl chains in the cations exhibit some
signs of disorder, but this was not modelled. The chloroform solvent molecule was modelled as fully disordered over
two sets of positions. NH atoms were freely refined using geometric restraints.
For GW114 diffraction data were collected on a Bruker APEX 2 CCD diffractometer using graphite
monochromated Mo-Kα X-radiation with narrow slice, 0.3° ω-scans. The carbazole ring N(3) > C(29) was modeled
as disordered over two sets of positions. Both substituted pyridine molecules in the asymmetric were also modeled as
disordered over two sets of positions. Geometric and Uaniso restraints were used. In GW123 diffraction data were
collected on a Bruker APEX 2 CCD diffractometer using graphite monochromated Mo-Kα X-radiation with 0.5° ω-
scans. A single t
butyl group C(29) > C(30) was modelled as fully disordered over two sets of positions for the Me
groups. For GW115 diffraction data were collected on a Bruker APEX 2 CCD diffractometer using graphite
monochromated Mo-Kα X-radiation with narrow slice, 0.3° ω-scans. The structure was solved by direct methods.21,22
The chloroform hydrogen H(35) and NH hydrogen atoms H(6) and H(8), were freely refined using geometric
constraints. In GWALS11 diffraction data were collected on a Bruker D8 diffractometer, equipped with a PHOTON
100 CMOS detector, with shutterless ω-scans. For GWALS12 diffraction data were collected on a Bruker D8
diffractometer, equipped with a PHOTON 100 CMOS detector, with shutterless ω-scans. No solvent of
crystallization was present in the asymmetric unit, which exhibited full molecule, two-fold disorder. This was
modelled using geometric restraints, and the major disorder component refined to have an occupation of ca. 66%.
For GWALS13, GWALS14, GWALS15 and GWALS16, diffraction data were collected on a Bruker D8
diffractometer, equipped with a PHOTON 100 CMOS detector, with shutterless ω-scans. No disorder or solvent of
This journal © The Royal of Dalton r
crystallization was present in the asymmetric unit and NH atoms were freely refined where present. For RSH29
diffraction data were collected on a Bruker APEX 2 CCD diffractometer using graphite monochromated Mo-Kα X-
radiation with narrow slice, 0.3° ω-scans. Hydroxyl and water H atoms were located by inspection of the difference
maps and electron density peaks. H atoms H(2), H(4), H(10), H(18), H(20), H(25), H(25A), H(26), H(26A), H(27)
and H(27A) were freely refined with geometric restraints.
Programs used were Bruker APEX 2 (data collection), Bruker SAINT (integration and cell refinement),
SHELXTL suite (solution, refinement and graphics), SHELXLe (refinement), and local programs. These data, along
with abundant supplementary figures and descriptions of each sample generated by the author are available at
ws8.lboro.ac.ukcm-xray-datawarwick or by request. Crystal data will become available free of charge at
www.ccdc.cam.ac.uk/data_request/cif when and if structures are published. All of these data have been separately
archived within the university and on the accompanying DVD ROM.
Results and Discussion
TY57 & TY59
Crystal structures of calixarene samples were determined in order to asses their ability to bind guests through
hydrogen bonding. For TY59 the asymmetric unit (Fig. 1) contains four calixarenes, two chloride anions, two tetra-
n-butylammonium cations, two acetonitrile and one chloroform, molecules of crystallisation. Within each of the four
calixarenes are pairs of N—H···O H-bonds between urea moieties to a single carbonyl O atom. Looking down on the
square-shaped S4 planes of the four unique calixarenes, three are approximately geometrically aligned, whist one,
containing S(1A), is slightly twisted. Most importantly, pairs of calixarene molecules are linked via four H-bonds
between both urea NH moieties on each calixarene and a chloride ion. The structure provides unequivocal proof of
2:1 host-to-guest binding properties of a 1,3-alternate thiacalix[4]arene bearing two phenyl-ureido moieties bound,
through ether O atoms, to the p-position of two opposing aryl rings which are part of the central calixarene cavity
(herein referred to as oligomeric rings). A full host-guest unit is shown in Fig. 2a. These results corroborate UV-vis
analyses performed by Kumar and co-workers, who demonstrated 2:1 host-to-guest binding behaviour in an
analogous calix[4]arene.24a
Structural differences between this analogous calix[4]arene, R1†, and TY59 include CH2
instead of S bridges; two n
propoxy instead of benzyloxy substituents bound to oligomeric rings; and most saliently,
p-nitrophenyl-ureido, instead of phenyl-ureido, moieties bound through ether O atoms.
Structurally, TY57 (Fig. 2b) differs from TY59 only through bearing p-tolyl-ureido, not phenyl-ureido,
moieties. Its crystal structure is however absent of any chloride anions and solvent of crystallisation. Such was the
significance of our X-ray diffraction studies on these samples, they have been included in a paper due for publication
in the New Journal of Chemistry later in 2015.25
It builds the work of Kumar24a~c
and a study by Lhoták and co-
workers which demonstrates the necessity of both aryl-ureido moieties to be bound to the p-, and not m-, or both
This journal © The Royal of Dalton r
p- and m- positions of the oligomeric rings.26
Our study assesses the effect of electron donating or withdrawing
groups bound to the p- position of the aryl-ureido moieties, on the anion binding ability of a series of 1,3-alternate
thiacalix[4]arene-based receptors, of which TY57 and TY59 and R3 are a part. Association constant data (Table 1)
for these receptors clearly demonstrate the enhancing effect of electron withdrawing substituents attached to the
p- position of the R-p-aryl-ureido moieties, on the acidity of the urea hydrogen atoms. R3, which was not subject to
X-ray analysis, shares the same ‗scaffold‘ as TY57 and TY59, and so these data show an increasing binding ability
of 1,3-alternate thiacalix[4]arene receptors towards Cl-
with increasing electronegativity of the R substituent on the
R-p-aryl-ureido chains. This occurs as more electronegative substituents increase the acidity of the urea hydrogen
Fig. 1 The asymmetric unit of sample TY59 highlighting its anion-binding properties. Non-NH hydrogen atoms are omitted for clarity. H-
bonds are shown as dashed lines.
Host TY57 TY59 R3 R1a
R2b
Guest Cl-
Cl-
Cl-
Cl-
Cl-
R-p-aryl-ureido; R= tolyl none trifluoromethyl nitro nitro
Association constant 1,500±110c
2,940±210c
6,590±460c
3,470,000±90,000d
270±70e
Table 1 Selected binding constant data from24a,25,26
highlighting the electronic effects of p- substituents on binding ability. a
Receptor 1 from24a
as
described above. b
Receptor 9b from26
with p-nitrophenyl-ureido moieties substituted at the m- position (i.e. both leaning off to one side) of
oligomeric rings. c
Calculated from Δδ of NH(x) protons by titration of host at 4.0 x 10-3
M with the guest, in CDCl3–CD3CN (10:1, v/v) at 298K,
300MHz.25 d
Calculated from Δabsmax, at 300 & 342nm, of host at 5 x 10-5
M upon incremental addition of 0–100 mol. equivalents of guest at
298K, using non-linear regression analysis program SPECFIT.24a,27 e
Calculated from Δδ of NH(x) protons by titration of host at 0.5–2.0mM with
guest to afford host:guest ratios 1- 20:1, in DMSO-d6 at 298K, 300MHz.26
Guests added as tetra-n-butylammonium (TBA) salts.24a,25,26
This journal © The Royal of Dalton r
Fig. 2a (left) The full bifurcated H-bonding environment of Cl(1) is shown for TY59 with the tetra-n-butylammonium cation observed
approaching from behind. 2b (right) The asymmetric unit of TY57 is shown from the other side, highlighting the continuous parallel
arrangement of urea moieties which runs continuously along the crystallographic b axis to give an extended, single helical H-bond array. Non-
H-bonding H atoms are omitted for clarity.
atoms, strengthening the H-bonds between the host and the guest through stabilization of the conjugate base. This is
verified by H-bond length data obtained via our diffraction experiments. For the TY59 fragment shown, the chloride
anion is held in place by H-bonds: N(1)—H(1)···Cl(1), N(2)—H(2)···Cl(1), N(3A)—H(3A)···Cl(1) and N(4A)—
H(4A)···Cl(1) of A···H length 2.26(2), 2.81(3), 2.87(3) and 2.30(2) Å respectively. TY57 meanwhile has
intermolecular N(3)—H(3)···O(5A) and N(4)—H(4)···O(5A) H-bonds of A···H length 2.13(3) and 2.35(7) Å
respectively. These slightly shorter distances are likely a consequence of the diminished donor capability of urea N
atoms due to the presence of the immediately adjacent, electron donating tolyl, instead of phenyl substituents. For
completion, the low association constant for receptor R2 is a consequence of the nitrophenyl-ureido moieties being
attached to the m- positions of oligomeric rings, forbidding any cooperative H-bonding within each calixarene by
lengthening the A—D distances. This in turn allows only facile 1:1 host-to-guest binding behavior for this receptor.
The 1000-fold increase in selectivity towards Cl-
of R1 compared to TY57 and TY59 meanwhile, is most likely due
to the presence of n
propoxy groups. These would have less severe steric impositions than the benzyloxy groups of
both TY samples, perhaps therefore allowing a more favourable enclave for the captured Cl-
anion. Without
diffraction data for both R1 and R2 however, these notions remain debatable.
This journal © The Royal of Dalton r
TY53 & TY55
Both samples are 1,3-alternate thiacalix[4] arenes bearing two 2-(1-methyl-1H-imidazolyl)methoxy groups bound to
the p- positions of oligomeric rings, differing by the remaining oligomeric rings being bound to benzyloxy groups for
TY55 (Fig. 3a) and 2-(ethyl ethanoyl)oxy groups for TY53 (Fig. 3b). Both samples crystallize in space group P1 ,
and exhibit remarkably similar disorder. Firstly, TY55 exhibits two-part disorder about a single C(58) > C(64)
benzyl substituent (major occupancy 50%), whilst TY53 exhibits two-part disorder about a single C(49) atom in an
ethanoyl moiety (major occupancy 56%). A single C(53) > C(57) 1-methyl-1H-imidazole ring in TY55 also exhibits
two-part disorder. As this 1-methyl-1H-imidazole ring moves between the two positions, cavities are formed in the
positions temporarily left unoccupied by the 1-methyl-1H-imidazole ring. It is in these two cavities that the disorder
components of one of the ethanol molecules, which is three-fold disordered, are found. The two ethanol disorder
components containing O(5X) and O(5Y) are permitted when the minor component C(54X) > N(4X) 1-methyl-1H-
imidazole ring is present. The ethanol disorder component containing O(5) is permitted when the major component
C(54) > N(4) 1-methyl-1H-imidazole ring is present. This arrangement of disorder components in turn permits a
single O(5)—H(5)···N(4) bond of length 2.75(2) Å. The connection between 1-methyl-1H-imidazole disorder and
ethanol disorder was however not modelled. Instead this ethanol molecule was modelled as fully disordered over all
three positions irrespective of 1-methyl-1H-imidazole ring site occupation using an occupancy sum restraint. The
major component occupancy for the 1-methyl-1H-imidazole ring is 56% whilst the restrained major component
occupancy of the ethanol molecule is 42%. There is a remaining non-disordered ethanol molecule within the
asymmetric unit, and all disorder components are shown (Fig. 3c).
Two of the three ethanol molecules in sample TY53 were modelled as two-fold disordered with both major
and minor disorder components of both ethanol molecules H-bonding, albeit through different H atoms, to N atoms
in the immediately adjacent 1-methyl-1H-imidazole rings C(42) > N(2) and C(51) > N(4). A single disordered water
molecule is also present in the asymmetric unit through which ethanol disorder components also H-bond to form
centrosymmetric dimers, in which calixarenes approach one another head-to-head (Fig. 3d). The remaining ethanol
molecule was modelled as a diffuse cloud of electron density using the Platon ―Squeeze‖ procedure.23
This disorder is surprising, given that an analogous 1,3-alternate thiacalix[4]arene, R4, bound at the p-
position of all four oligomeric rings by 2-(1-methyl-1H-imidazolyl)methoxy groups, refined to completion with an
R-factor of 4.01% without the use of restrains, and showed no signs of disorder.28
By comparison, R-factors of TY53
and TY55 are less good, at 6.84 and 9.18% respectively. In R4 all four 1-methyl-1H-imidazole rings lie away from
the central calixarene cavity, whereas in both TY samples, the 1-methyl-1H-imidazole ring C(42) > N(2) nestles
within the central calixarene cavity, parallel to the two flanking oligomeric rings. Remarkably, this has little effect on
the dimensions of the S4 units which remain square-shaped, and have areas 30.714 Å2
for R428
and 31.147 A2
for
TY53. Inclusion of 2-(ethyl ethanoyl)oxy and benzyloxy groups does seem to affect the shape of the overall
This journal © The Royal of Dalton r
calixarene structure however with both TY53 and TY55 having a more conical shape than R4 , whose cuboid
shape and tetrahomosubstitution with lone pair-donating moieties is essential in allowing 1:2 host-to-guest binding of
transition metals.28
TY49
Oxacalix[3]arene derivatives whose electron-accepting/donating substituents all reside on one side of the O3 cavity
can selectively bind anionic/cationic guests with C3 symmetry,29
and TY49 (Fig. 4a) is an example of such a
calixarene. The three oligomeric rings are functionalized through the p- position with [N-(2-Pyridinyl)acetamyl]oxy
Fig. 3a (top left) The asymmetric unit of TY55 with only major disorder components shown. 3b (top right) The asymmetric unit of TY53
with only major disorder components shown. 3c (bottom left) Sample TY55 with all disorder shown. 3d (bottom right) The H-bonding array
observed in sample TY53. Non-H-bonding H atoms are omitted for clarity.
This journal © The Royal of Dalton r
groups, two of which reside within the central calixarene cavity, whilst the cavity itself adopts a collapsed, or
squashed conformation to facilitate a single intramolecular N(3)—H(3)···O(8) H-bond, of length 2.29(2) Å. So
distorted is the cavity that a π···π interaction is observed, between oligomeric rings C(13) > C(18) and C(25) >
C(30), with a distance of 3.87 Å. The o-pyridyl groups C(46) > N(4) and C(53) > N(6) meanwhile also align face-to-
face, but at a distance of 4.80 Å which is too far for π···π stacking. This conformation appears to give rise to a very
favourable self-assembly of molecules in the solid state, with a further intermolecular C—H···O H-bond, bond of
length 2.28 Å, occurring between the m-pyridyl H(56) atom and a carbonyl O(7) atom (Fig. 4b & c). The benzyloxy
analogue, R5,30
of TY49 adopts much more open conformation by comparison. For example, the narrowest angle
between the three cavity-located O atoms is ∠O(1)-O(2)-O(3) = 47.71° for TY49 whilst the analogous angle for R5
is 52.81°.30
This wider angle necessitates greater π···π distances between oligomeric rings, with the closest distance
between any two being 4.51 Å. Moreover, the angle between the two planes of the π···π-stacked C(13) > C(18) and
C(25) > C(30) rings in TY49 is 23.97°, whilst the two closest oligomeric rings in R5 are far less parallel in their
alignment, with the angle between their respective planes being 41.35°.30
A tentative conclusion therefore is that the
cation binding ability of TY49 may be diminished by competitive intramolecular H-bonding between potential
binding sites, and poorly optimized cavity geometry.
Fig. 4a (left) Receptor TY49 showing the labelling scheme and the distorted central cavity. π···π stacking occurs vertically in this view. 4b
(middle) The intra- and intermolecular H-bonding of receptor molecules, highlighting the ‗dog-legged‘ [N-(2-Pyridinyl)acetamyl]oxy
substituent. 4c (right) A packing plot for receptor TY49. Non-H-bonding H atoms are omitted for clarity.
This journal © The Royal of Dalton r
TY44
This sample is a 1,3-alternate thiacalix[4]arene substituted with two pyridyl-4-methoxy groups, and two benzyloxy
groups (Fig. 5a). This receptor is one of a family or receptors, which were studied elsewhere, in which the pyridyl-4-
methoxy groups are replaced by pyridyl-3- or pyridyl-2-methoxy groups.31
i.e. the position of the pyridyl N atom
changes with each member of this family, resulting in some measurable change in binding properties of the members.
The TY44 thiacalixarenes align in columns running parallel to the crystallographic c axis, ca. 45° from the b axis,
with chloroform solvent molecules of crystallization residing in voids between any four thiacalixarenes, generated by
the observed vertex-to-vertex packing of the square-shaped S4 units, which all lie in a common plane. This packing
affords S···S interactions of length 3.32–3.63 Å between sulfur atoms on adjacent thiacalixarenes (Fig. 5b). In
contrast, the pyridyl-2-methoxy-substiuted analogue, R6, packs with offset S4 planes, occluding any S···S
interactions. The S···S distances in R6 are all 4.25 Å.31
Both TY44 and R6 however pack such that adjacent
thiacalixarene molecules have their pyridyl substituents aligned in opposite directions along the normal to the S4
plane, probably to minimise lone pair interactions between N atoms on adjacent molecules. This, in the case of
TY44, allows molecules to pack more closely, permitting the S···S interactions.
Some C—H···π interactions are observed between chloroform molecules and the oligomeric rings of TY44.
In particular, the non-disordered chloroform molecule exhibits a C(67)—H(67)···π{C(21) > C(26)} interaction of
distance 3.10 Å. The disordered chloroform meanwhile makes one of two mutually exclusive
Fig. 5a (left) A thermal ellipsoid plot of receptor TY44, highlighting the substituent conformation about the calixarene cavity. H atoms and
solvent of crystallization are omitted for clarity. 5b (right) A packing plot of TY44 with S···S contacts drawn. Non-chloroform H atoms are
omitted for clarity. Thermal ellipsoids are drawn at the 50% probability level.
This journal © The Royal of Dalton r
C—H···π interactions, with the major component giving C(68)—H(68)···π{C(11A) > C(16A)} and the minor
component giving C(68X)—(H68X)···π{C(31) > C(36)} (Fig. 5b). These have distances of 3.02 and 3.09 Å
respectively. The distance between N atoms in TY44 is long, at 8.29 Å, due to the pyridine rings lying outside of the
central thiacalixarene cavity. This distance is comparable to the analogous distance in R6, which is 9.08 Å.31
The
binding capability of TY44 is likely therefore to be poor, given that the upward-pointing N atoms would only be able
to interact with host species through weak, singular N···Guest contacts, and not strong N,N‘···Guest···S,S‘ contacts
demonstrated for R6,31
bearing in mind that, for pyridyl-2-methoxy substituents, the N occupies the o- position.
Rotation of the pyridine ring about the methoxy bridge, which is possible in solution, would create therefore an
extremely electron-rich, S-, O- and N-containing enclave, suitable for small cationic species with accessible orbitals
for electron donation.
TY51
Moving on to the blue-light emitting samples, TY51 is formally 1,2-di-(7-t-butylpyrene)-3,4,5,6-tetraphenylbenzene
(Fig. 6a). The two 7-t-butylpyrenyl and four phenyl substituents align in a propeller-like fashion around the central
arene ring C(41) > C(46), with the two pyrene rings aligning without overlap to avoid unfavourable steric interactions
between their respective t
butyl groups. The torsion angles of the arene substituents relative to the central ring are
summarized in Table 2.
Ring Torsion angle, in°
C(1) > C(16) 80.11(8)
C(21) > C(36) 86.97(8)
C(47) > C(52) 75.96(12)
C(53) > C(58) 65.19(13)
C(59) > C(64) 67.61(15)
C(65) > C(70) 70.34(16)
Table 2 Torsion angles of arene substituents relative to the central arene ring C(41) > C(46) of TY51.
The central arene rings all align with their C6 axes parallel to one another, with the C(47) > C(52) ring nestling
between the two pyrene rings on adjacent molecules (Fig. 6b). This permits multiple C–H···π interactions which lie in
the range of 2.56–3.39 Å. No π···π stacking is present in the crystal structure, however, due to the steric impositions
of the propeller-like topology of the individual molecules. Only the central C(47) > C(52) rings are aligned in such a
way as to allow π···π stacking between adjacent molecules, however the centroid···centroid separation of these rings
is too large, at 7.54 Å, to allow π···π stacking to occur.15
It is well known32a~c
that π···π stacking promotes excimer
formation between proximal molecules, and that these excimers, by definition, emit lower energy photons than the
original monomeric species. As such, excimer formation between prospective blue-light emitting molecules
This journal © The Royal of Dalton r
undermines any optimization techniques hitherto employed, as emissive wavelengths are bathochromically shifted
towards greener wavelengths, making the solid-state structure of these materials somewhat critical to their success.
Thin films of the prospective blue-light emitter 1,3,5,9-tetra(p-methoxyphenyl)-7-t-butylpyrene, R7, with the
comparable pyrene centroid···centroid separation to that of TY51, 8.57 Å, was shown to emit blue light of
wavelength 471 nm upon application of an electric current.33
Moreover, these emissions occurred with a high Φf
of 0.48 in DCM at 298K, and a small stokes shift of 2 nm, indicating that π···π stacking was neither happening in the
solid or solution state.33
It may be concluded therefore that, assuming substituent electronic effects are minimal, TY51
is a promising candidate material for blue OLEDs.
TY54
In contrast to TY51, TY54 does show significant π···π stacking between adjacent molecules. TY54 itself is a pyrene-
derivatized molecule, bearing two phenyl substituents at the 1, and 3 positions, a t
butyl group at the 7 position, and
naphthalene rings bridged by imine groups to the 4, & 5 positions, and the 9, & 10 positions (Fig. 7a). The large
torsion angles between the phenyl rings and the central pyrene, which are of the order of those observed in TY51, and
the t
butyl group, are not able to supress close approach of the flat naphthyl ‗spines‘. Hence adjacent molecules pack
with their pyrene cores co-aligned in parallel planes (Fig. 7b). The pyrene core of each molecule is approached from
above by the ring C(1) > C(6) on one immediately adjacent molecule, and from below by the ring C(11) > (16) on a
second, immediately adjacent, molecule. The centroid···centroid distances of these two π···π interactions are 3.82 and
3.75 Å respectively. Furthermore, this π···π bonding array is observed at each molecule, running continuously
Fig. 6a (left) Emitter TY51 with the full labelling scheme shown. 6b (right) A packing plot of TY51 highlighting the steric impositions of the
propeller-like topology.
This journal © The Royal of Dalton r
Fig. 7a (top) emitter TY54 with a full labelling scheme shown. 7b (bottom left) A packing plot highlighting the π···π stacking in TY54.
7c (bottom right) a side-on view of TY54 emphasizing the twisted nature of the imine-bridged naphthalene rings.
through each plane. Molecules in adjacent planes meanwhile pack with their t
butyl groups approaching one another,
and their two phenyl substituents approaching one another. Where molecules in two adjacent planes‘ phenyl groups
approach one another, a single, very weak π···π interaction of distance 4.52 Å is observed between the two C(41) >
C(46) rings on the neighbouring molecules.
Interestingly the two imine-bridged naphthyl substituents twist away from one another, about the axis which
is the perpendicular bisector of the C(2)—C(3) and C(12)—C(13) bonds. The angle between the arene rings C(1) >
C(6) and C(11) > C(16) about this axis is 3.9(2)°. This behavior is radically different from that observed in an
analogous compound, R8, which bears two p-formylphenyl substituents instead of phenyl substituents, and two
imine-bridged benzene substituents instead of naphthalene substituents. In R8, the pyrene core is bent in on itself
about the same axis, giving the molecule a characteristic banana shape. For comparison, using the labeling scheme of
This journal © The Royal of Dalton r
TY54, the acute angle between the planes of the C(22) > C(27) and C(29) > C(34) rings in R8 is 13.6°.34
This
difference in conformation is not reflected however in the π···π distances, which remain short for R8, at 3.58 Å.34
This is shorter than the π···π distances observed in TY54, and so the p-formylphenyl substituents are more co-planar
with the pyrene core, with torsion angles of 51 and 58°,34
whilst the torsion angles of the phenyl substituents of TY54
are 75.77(11) and 80.92(13)°.
Conclusion
Single crystal X-ray diffraction techniques have been employed to determine the solid state structure of eight
samples, representing two different areas of chemistry, in collaboration with Takehiko Yamato. The collaborative
nature of this work necessitated completion to a publishable standard, and indeed experimental results for two of
these eight determinations have been included in an advanced draft of a research paper. The most significant finding
is irrefutable evidence of 2:1 host-to-guest binding behavior of a 1,3-alternate thiacalix[4]arene bearing phenyl-
ureido and benzyloxy substituents, corroborating the results of a separate study which did not contain X-ray data.24a
The remainder of our structural findings add to the current structure-property investigation literature available both
for classes of compounds studied here. The most salient experimental results for each structure were abstracted from
the abundance of quantitative data collected, and presented as a package tailored to the research areas of interest to
our collaborator. Upon future publication of our X-ray work, Crystallographic Information Files, or CIFs, for all of
the structures studied will be deposited with the Cambridge Structural Database and so become freely accessible
through moiety and key-word etc. database interrogations.
Whilst the different classes and sub-classes of compounds studied here undermine the validity of direct
comparisons between them, comparisons to reference compounds highlight their currency. Researchers involved in
calixarene host-guest chemistry, and OLED engineering are encouraged to engage with crystallographic
collaborators as fully as possible, to reduce the amount of structure-property research published without X-ray
structures, which seems oxymoronic. In particular, if more dedication as given to growing crystals of a host material
with the guest entrapped, variable-temperature X-ray diffraction experiments could elucidate how a guest is released
from the host. From this, something might be learnt of what subtle structural changes need to take place for a guest to
successfully bind to a host. This may have particular significance for less rigid calix[3]arenes, and involve such
things as reconfiguration or deformation of the calixarene ‗scaffold‘, repositioning of functional groups, and
antagonistic steric effects imposed by a crystalline structures. Should such experiments afford any useful data,
materials may be optimized during synthesis for better function in either the solution state, where (say) hosts are
added to complex with chromate, or other guests, which are harmful to aquatic life. Solid state optimization
meanwhile could allow an improvement in the selectivity of ion-selective electrodes or synthetic membranes.
This journal © The Royal of Dalton r
Additional Work
Structures
Upon completion of all relevant samples submitted by the Takehiko Yamato group, refinement, modelling, validation
and description were performed on/for ten additional samples. GW114, GW115, GW123, GWALS11, GWALS12,
GWALS13 GWALS14, GWALS15, GWALS16 and RSH29 were studied as the proficiency of the author in these
steps allowed increasingly rapid turnaround of X-ray results. These samples possess radically different chemistry to
the ‗TY’ samples, and that suggested by the title of this report however. They are therefore excluded from the main
discussion. A considerable investment of time (which is not represented by this disproportionately small discussion)
and development of the crystallographic skill of the author resulted from these additional studies. The diverse range of
structures presented more challenging, or even entirely new ‗classes‘ of crystallographic disorder, a full discussion of
which cannot be presented here for reasons of brevity.
For example, GW114 contained two novel disorder scenarios which required modelling. Firstly, full-
molecule disorder about a carbazole substituent mandated the splitting of the entirety of this substituent into two
restrained disorder components occupying separate regions in space. A before-and-after for the modelling of this
disorder is shown in Figs. 8a & b.
Fig 8a (top left) a thermal ellipsoid plot of GW114 showing the
undesirable prolate displacement ellipsoids of the carbazole carbons.
These are indicative of two-fold disorder. Fig 8b (top right) the
carbazole substituent has now been modelled split over two sets of
positions for the constituent atoms, which possess much more
satisfactory displacement ellipsoids. Fig 8c (bottom left) The novel
disorder corresponding to overlapping pyridyl N and C atoms is
emphasized. Thermal ellipsoids are drawn at 50% probability and H
atoms are omitted for clarity.
This journal © The Royal of Dalton r
The cause of this disorder was related to the second class of disorder present, in which the entire molecule could adopt
one of two orientations during the crystal growth phase. For this particular case, the result was an electron density
map in which the atomic position of N(1) is partially overlapped by C(4) and vice versa. Looking at Fig 8c it can be
seen that these positions correspond to an imaginary C2 axis along the C(2)—F(2) & C(5)—F(5) bonds, in one
molecule, and the C(2A)—F(2A) & C(5A)—F(5A) bonds in the second molecule (with both of the two molecules in
the asymmetric unit experiencing this disorder, where ―?A‖ is used to distinguish between molecules). 2-fold rotation
about this axis affords a major disorder component, where, focusing on only one of the molecules, the atomic position
of N(1) is partially occupied by C(4X), and a minor disorder component, where the atomic position of C(4) is partially
occupied by N(1X). Necessarily, the F(4) atom bound to C(4) is also disordered about this axis. It was found that the
disorder observed in the carbazole substituent was due to two favorable, mutually exclusive C—H···F interactions
between the disordered carbazole substituent, and either F(4A) in one adjacent asymmetric unit, or F(2) in another.
Obviously, F(4A) is only present at the site which permits this C—H···F interaction when the entire F4A-containing
molecule adopts the appropriate orientation, hence the connection. GW115 is another example of a complex
modelling procedure, due to its extensive H-bonding array. Here, each macromolecule (Fig 9a) was H-bonded to three
other macromolecules via three unique sets of H-bonds. These are as follows: The first and second macromolecules
are H-bonded to one another by two bifurcated H-bonds between hydrogen atoms H(25B) and H(27B) on one
macromolecule, to F(4) on the other macromolecule, and vice versa. Between the first and third is a single N—H···N
H-bond between H(6) on the first macromolecule and N(2) on the third macromolecule.
Fig 9a (left) GW115 is shown. Each macromolecule is H-bonded to a single chloroform atom. Fig 9b (right) the complex H-bonding
environment of a single macromolecule. Here H atoms are omitted for clarity if not involved in H-bonding.
This journal © The Royal of Dalton r
Lastly, between the first and the fourth is a pair of parallel H-bonds. These are between O(2) on the first
macromolecule to H(14) on the fourth macromolecule, and H(8) on the first macromolecule to N(3) on the fourth
macromolecule. This H-bonding (Fig 9b) array is observed at each macromolecule, generating a nexus of
intermolecular H-bonds. Work on these samples was of course performed to the same standard as the TY samples,
and results were delivered their respective submitters in exactly the same package. These data have too been archived,
and crystal data is included in appendix i.
Guides
Three documents were written by the author, for the intended use of future X-ray project students. The first, is entitled
―Basic SHELXL Commands for Small Molecule Refinement‖ is a tabulation of SHELX (structure modelling)
commands learned throughout the entirety of the project. Commands are presented with their appropriate syntax and
function explained in short, digestible paragraphs, occasionally taking the form of short tutorials for the commonest
tasks. Examples of simplified instruction files with the necessary entries required for a command to work are also
provided where appropriate. All of the commands are presented with useful tips for efficient, tidy structure modelling
as learned by the author, and some salient in-house protocols are presented for the sake of internal consistency. An
introduction is also provided explaining how to download and set up SHELXL and its GUI SHELXLe, include these
executables in a computer‘s environment variables for command prompt loading, and some basic commands for
navigating the command prompt. It is hoped that such a section, and the document as a whole, will equip the reader
with essential computational skills and program-specific knowledge to set up, manage, and operate this software with
confidence. This guide is ca. 6,000 words in length.
The second guide, entitled ―Commands for Probing Materials in XP Interactive Molecular Graphics‖ has the
same basic layout as the above, but this time with XP (structure probing and graphics) commands described. Again,
an intro detailing program setup is provided, as well as some essential information on file extensions. Due to the
diverse combination of XP commands, tutorial-type descriptions are kept to a minimum. The most salient diagram
generation commands are presented as short tutorial. This guide is ca. 9,000 words in length.
The final guide is entitled ―Tutorial For Mapping the K Drive to an Off-Campus Machine‖. The K: drive is
the university‘s X-ray network drive, which is usually accessed via a clumsy VPN client which limits users to
uploading/downloading files one-at-a-time amongst other productivity-limiting frustrations. This guide instructs the
reader on how to map the K: drive to their network-connected PC or laptop, from where they may access their work as
if it were stored locally on their machine, and therefore ‗work from‘ the K: drive. Additionally, instruction is given on
mapping the students U: drive, provided to all Loughborough University students.
This journal © The Royal of Dalton r
Dalton Transactions Template
The template for this report was generated by retroconversion of a generic Dalton Transactions manuscript into .docx
format, and has been made available for those who wish to use it. All additional work is available by request from the
author or Dr. Mark R. J. Elsegood, and is available at ws8.lboro.ac.ukcm-xray-datawarwick where they are
provided with their raw text versions for future additions, corrections and improvements. The attached DVD ROM
also contains all work, supplementary and otherwise, completed during this project.
Acknowledgement
The Advanced Light Source is supported by the Director, Office of Science, Office of Basic Energy Sciences, of the
U.S. Department of Energy under Contract No. DE-AC02-05CH11231. Dr. Mark Elsegood is thanked for his
guidance on matters both practical and literary.
This journal © The Royal of Dalton r
Appendix i
Experimental and Crystal Data
Compound TY44 TY49 TY51
Formula C66H70N2O4S4·CHCl3 C57H66N6O9·3(COH4)·H2O C70H54·CH4O
Mr [gmol-1
] 1202.84 1093.30 927.17
Crystal system Triclinic Monoclinic Orthorhombic
Space group 1P C2/c Pna21
Unit cell dimensions
a [Å] 16.3706 (10) 26.1641 (11) 11.6362 (4)
b [Å] 19.2771 (12) 15.4995 (6) 26.6850 (8)
c [Å] 20.9368 (13) 28.5153 (11) 17.6959 (6)
α [°] 83.800 (4) 90 90
β [°] 82.809 (4) 94.063 (3) 90
γ [°] 71.799 (4) 90 90
V [Å3
] 6210.6 (7) 11534.8 (8) 5494.8 (3)
Z 4 8 4
T [K] 100 100 100
Dx(calc) [g cm-3
] 1.286 1.259 1.121
Radiation type Synchrotron Synchrotron Synchrotron
Radiation λ [Å] 0.7293 0.7085 0.7749
Crystal size [mm3
] 0.10 × 0.08 × 0.04 0.30 × 0.17 × 0.03 0.11 × 0.06 × 0.04
Extinction Coeff. N/A 0.0008 (2) 0.0066 (10)
2θmax [°] 75.0 66.2 55.2
Reflections measured 236971 113193 43351
Unique reflections 59903 22109 9808
reflections with
I > 2σ(I)
45699 16974 8404
Transmission factors 0.966, 0.986 0.974, 0.997 0.992, 0.997
Rint 0.062 0.074 0.072
Number of parameters 1475 747 669
Number of restraints 0 407 51
R[F2
> 2σ(F2
)] 0.068 0.067 0.058
wR(F2
) 0.201 0.205 0.156
This journal © The Royal of Dalton r
Compound TY53 TY54 TY55
Formula C58H72N4O8S4·3(C2H6O)·O C52H34N4 C68.30H84.45Cl0N4O6.03S4
Mr [gmol-1
] 1253.63 714.83 1186.24
Crystal system Triclinic Triclinic Triclinic
Space group 1P 1P 1P
Unit cell dimensions
a [Å] 13.9134 (18) 10.2435 (4) 14.6911 (6)
b [Å] 15.4414 (19) 13.5620 (6) 14.8673 (6)
c [Å] 17.086 (2) 13.7266 (6) 15.0303 (7)
α [°] 103.616 (2)° 106.560 (3) 85.407 (3)
β [°] 104.584 92.709 (3) 86.152 (3)
γ [°] 96.412 (2)° 102.490 (3) 79.641 (3)
V [Å3
] 3395.2 (7) 1772.45 (13) 3214.3 (2)
Z 2 2 2
T [K] 150 100 100
Dx(calc) [g cm-3
] 1.209 1.339 1.226
Radiation type Mo K radiation Synchrotron Synchrotron
Radiation λ [Å] 0.71073 0.7749 0.7749
Crystal size [mm3
] 0.88 × 0.42 × 0.20 0.30 × 0.17 × 0.05 0.11 × 0.09 × 0.01
Extinction Coeff. 0.0043(6) 0.017 (4) N/A
2θmax [°] 48.0 48.8 29.2
Reflections measured 37604 10339 22331
Unique reflections 12465 3836 8081
reflections with
I > 2σ(I)
8007 2960 5492
Transmission factors 0.844, 0.961 0.973, 0.995 0.973, 0998
Rint 0.062 0.041 N/A
Number of parameters 818 510 947
Number of restraints 103 0 784
R[F2
> 2σ(F2
)] 0.068 0.056 0.092
wR(F2
) 0.190 0.172 0.277
This journal © The Royal of Dalton r
Compound TY57 TY59
Formula C74H84N4O6S4·CHCl3·0.5(C2H3N) C144H160N8O12S8·C16H36N+
·Cl·0.5(CHCL3)·C2H3N
Mr [gmol-1
] 1393.58 2829.91
Crystal system Monoclinic Triclinic
Space group P21/c 1P
Unit cell dimensions
a [Å] 20.1570 (13) 15.1315 (5)
b [Å] 26.8064 (17) 28.8618 (11)
c [Å] 27.0262 (17) 35.9491 (13)
α [°] 90 104.778 (2)
β [°] 93.8414 (12) 94.065
γ [°] 90 92.353 (2)
V [Å3
] 14570.4 (16) 15113.8 (9)
Z 8 4
T [K] 150 100
Dx(calc) [g cm-3
] 1.271 1.244
Radiation type Mo K radiation Synchrotron
Radiation λ [Å] 0.71073 0.7749
Crystal size [mm3
] 0.26 × 0.24 × 0.08 0.16 × 0.13 × 0.04
Extinction Coeff. N/A N/A
2θmax [°] 46.2 50.6
Reflections measured 127772 116001
Unique reflections 28608 42128
reflections with
I > 2σ(I)
17146 27707
Transmission factors 0.927, 0.977 0.956, 0.989
Rint 0.093 0.062
Number of parameters 1737 3705
Number of restraints 109 256
R[F2
> 2σ(F2
)] 0.092 0.068
wR(F2
) 0.301 0.185
This journal © The Royal of Dalton r
Compound GW114 GW115 GWALS11
Formula C29.02H16.02F3N3 C34H30F6N8O2·CHCl3 C41H24F2N4
Mr [gmol-1
] 463.77 816.03 610.64
Crystal system Orthorhombic Monoclinic Orthorhombic
Space group P212121 P21/n P212121
Unit cell dimensions
a [Å] 8.1301 (5) 13.8932 (16) 7.8011 (4)
b [Å] 18.1470 (12) 14.1669 (16) 13.4885 (6)
c [Å] 29.9929 (19) 18.231 (2) 27.7849 (13)
α [°] 90 90 90
β [°] 90 99.7139 (16) 90
γ [°] 90 90 90
V [Å3
] 4425.1 (5) 3536.8 (7) 2923.7 (2)
Z 8 4 4
T [K] 150 150 100
Dx(calc) [g cm-3
] 1.392 1.532 1.387
Radiation type Mo K radiation Mo K radiation Synchrotron
Radiation λ [Å] 0.71073 0.71073 0.7749
Crystal size [mm3
] 0.40 × 0.21 × 0.05 0.54 × 0.54 × 0.36 0.30 × 0.02 × 0.02
Extinction Coeff. N/A N/A 0.0077(9)
2θmax [°] 44.0 56.8 51.6
Reflections measured 30333 35615 20833
Unique reflections 9168 8798 4347
reflections with
I > 2σ(I)
6291 5995 3904
Transmission factors 0.961, 0.995 0.839, 0.888 0.968, 0.998
Rint 0.057 0.051 0.053,
Number of parameters 760 499 425
Number of restraints 546 2 0
R[F2
> 2σ(F2
)] 0.045 0.064 0.035
wR(F2
) 0.089 0.209 0.083
This journal © The Royal of Dalton r
Compound GWALS13 GWALS14 GWALS15
Formula C34H42S4 C34H24F8N6O4 C60H42F30N6O12
Mr [gmol-1
] 578.91 732.59 1608.99
Crystal system Triclinic Monoclinic Monoclinic
Space group 1P P21/n P21/c
Unit cell dimensions
a [Å] 4.8575 (2) 12.7854 (5) 17.4161 (6)
b [Å] 9.7907 (4) 8.9381 (4) 14.3221 (5)
c [Å] 16.7849 (6) 14.1393 (5) 25.7502 (9)
α [°] 101.411 90 90
β [°] 94.356 96.872 (2) 101.207
γ [°] 96.746 (2) 90 90
V [Å3
] 773.02 (5) 1604.19 (11) 6300.5 (4)
Z 1 2 4
T [K] 150 150 150
Dx(calc) [g cm-3
] 1.244 1.517 1.696
Radiation type Synchrotron Synchrotron Synchrotron
Radiation λ [Å] 0.7749 0.7749 0.7749
Crystal size [mm3
] 0.25 × 0.07 × 0.03 0.35 × 0.18 × 0.02 0.05 × 0.03 × 0.01
Extinction Coeff. N/A N/A 0.00050(8)
2θmax [°] 80.6 67.2 51.2
Reflections measured 14583 20949 44141
Unique reflections 7360 4910 9083
reflections with
I > 2σ(I)
6299 3877 6863
Transmission factors 0.904, 0.988 0.959, 0.998 0.992, 0.998
Rint 0.024 0.030 0.056
Number of parameters 173 239 997
Number of restraints 0 0 0
R[F2
> 2σ(F2
)] 0.036 0.043 0.055
wR(F2
) 0.104 0.121 0
This journal © The Royal of Dalton r
Compound GWALS16 RSH29 GWALS12
Formula C38H50S4 C21H13HoN3O12·1.5(H2O) C12H6BrF4NO
Mr [gmol-1
] 635.02 691.30 336.09
Crystal system Triclinic Triclinic Monoclinic
Space group 1P 1P P21/n
Unit cell dimensions
a [Å] 4.8430 (2) 11.6545 (3) 7.0989 (4)
b [Å] 9.7919 (5) 13.5819 (3) 13.8009 (7)
c [Å] 18.6253 (8) 15.1150 (4) 11.7747 (7)
α [°] 81.716 (2) 90.78 90
β [°] 82.641 (2) 95.76 91.175 (4)
γ [°] 82.641 (2) 103.15 90
V [Å3
] 862.11 (7) 2316.43 (10) 1153.34 (11)
Z 1 4 4
T [K] 150 150 150
Dx(calc) [g cm-3
] 1.223 1.982 1.936
Radiation type Synchrotron Mo K radiation Synchrotron
Radiation λ [Å] 0.7749 0.71073 0.6199
Crystal size [mm3
] 0.30 × 0.12 × 0.01 0.20 × 0.18 × 0.06 0.30 × 0.10 × 0.06
Extinction Coeff. N/A N/A 0.0020(6)
2θmax [°] 80.6 61.2 50.4
Reflections measured 16243 37412 15540
Unique reflections 8185 14098 3109
reflections with
I > 2σ(I)
6239 12120 2128
Transmission factors 0.895, 0.996 0.541, 0.818 0.518, 0.863
Rint 0.029 0.024 0.044
Number of parameters 191 779 344
Number of restraints 0 14 174
R[F2
> 2σ(F2
)] 0.053 0.024 0.031
wR(F2
) 0.0147 0.057 0.079
This journal © The Royal of Dalton r
Compound GW123
Formula C32H26F8O2
Mr [gmol-1
] 594.53
Crystal system Triclinic
Space group 1P
Unit cell dimensions
a [Å] 11.674 (3)
b [Å] 21.478 (5)
c [Å] 22.335 (5)
α [°] 88.008 (3)
β [°] 85.529 (3)°
γ [°] 86.292 (3)°
V [Å3
] 5569 (2)
Z 8
T [K] 150
Dx(calc) [g cm-3
] 1.418
Radiation type Mo K radiation
Radiation λ [Å] 0.71073
Crystal size [mm3
] 0.91 × 0.15 × 0.05
Extinction Coeff. 0.0023(2)
2θmax [°] 45.0
Reflections measured 462780
Unique reflections 14560
reflections with
I > 2σ(I)
9130
Transmission factors 0.896, 0.994
Rint 0.078
Number of parameters 1565
Number of restraints 50
R[F2
> 2σ(F2
)] 0.050
wR(F2
) 00.128
Appendix ii
CIF Locations for Reference Structures Discussed
Compound CIF location [Hyperlink; click to follow]
R1 Not Available
R2 Not Available
R3 Not Available
R4 https://summary.ccdc.cam.ac.uk/structure-summary-form 997001
R5 https://summary.ccdc.cam.ac.uk/structure-summary-form 841827
R6 https://summary.ccdc.cam.ac.uk/structure-summary-form 245644
R7 http://pubs.acs.org DOI: 10.1021/ol4002653
R8 http://pubs.acs.org DOI: 10.1021/ol401438a
This journal © The Royal of Dalton r
References
1. R, Muthukrishnan, J. Org. Chem., 1978, 43, 4905.
2. (a) S. Fleming, J. M. Harrowfield, M. I Ogden, B. W. Skelton, D. F. Stewart, and A. F. White, Dalton Trans., 2003, 17, 3319; (b)
https://sites.google.com/site/cdgutsche/home/bibliography (accessed April 2015).
3. (a) C. D. Gutsche, Monographs in Supramolecular Chemistry: Calixarenes, Royal Society of Chemistry, London, 1989; (b) C. D. Gutsche,
Monographs in Supramolecular Chemistry: Calixarenes Revisited, Royal Society of Chemistry, London, 1998; (c) C. D. Gutsche, Monographs
in Supramolecular Chemistry: Calixarenes, an Introduction, Royal Society of Chemistry, London, 2008.
4. H. Kumagai, M. Hasegawa, S. Miyanari, Y. Sugawa, Y. Sato, T. Hori, S. Ueda, H. Kamiyama, and S. Miyano, Tet. Lett.,1997, 38, 3971.
5. C. Zhang, Z. Wang, S. Song, X. Meng, Y.-S. Zheng, X.-L. Yang, and H.-Bi. Xu, J. Org. Chem., 2014, 79, 2729.
6. C. Redshaw, M. A. Rowan, L. Warford, D. M. Homden, A. Arbaoui, M. R. J. Elsegood, S. H. Dale, T. Yamato, C. Pérez Casas, S. Matsui,
and S. Matsuura, Chem. Eur. J., 2007, 13, 1090.
7. J.-L. Zhao, H. Tomiyasu, X.-L. Ni, X. Zeng, M. R. J. Elsegood, C. Redshaw, S. Rahman, P. E. Georghiou, and Takehiko Yamato, New J.
Chem., 2014, 38, 6041.
8. A. Yilmaz, S. Memon, and M. Yilmaz, Tetrahedron, 2002, 58, 7735.
9. A. I. Vovk, L. A. Kononets, V. Y. Tanchuk, S. O. Cherenok, A. B. Drapailo, V. I. Kalchenko, and V. P. Kukhar, Bioorg. Med. Chem. Lett.,
2010, 20, 483.
10. A. Pochettino, Atti. Accad. Rend., 1906, 15, 355.
11. D. M. Chapin, C. S. Fuller, G. L. Pearson, J. Appl. Phys., 1954, 25, 676.
12. C. W. Tang, Appl. Phys. Lett., 1986, 48, 183.
13. H. Spanggaard, F. C. Krebs, Sol. Energy Mat. & Sol. Cells, 2004, 83, 125.
14. http://www.nobelprize.org/nobel_prizes/physics/laureates/2014/press.html (accessed April 2015).
15. G. E. Bacon. Acta Crystallogr., 1951, 4, 558.
16. F. M. Winnik, Chem. Rev., 1993, 93, 587.
17. I. B. Berlman, Handbook of fluorescence spectra of aromatic molecules, Academic Press, New York 2nd
ed., 1971.
18. A. Dino, G. E. Jabbour, and N. Peyghambarian, Advanced Mat., 2000, 12, 1249.
19. B. Park, C. H. Park, Y. Yim, and J. Park. J. Appl. Phys., 2010, 108, 84508.
20. SAINT and APEX 2 Software for CCD Diffractometers, Bruker AXS: Madison, WI, USA, 2008.
21. G. M. Sheldrick, SHELXTL User Manual, version 6.10, Bruker AXS: Madison, WI, USA, 2000.
22. G. M. Sheldrick, Acta Crystallogr., 2008, A64, 112.
23. A. L. Spek, Acta Crystallogr., 1990, Sect. A, 46, C34.
24. (a) J. N. Babu, V. Bhalla, M. Kumar, R. K. Mahajan, and R. K. Puri, Tet. lett., 2008, 49, 2772; (b) M. Kumar, R. K. Mahajan, K. V.
Bhalla, N. Singh, H. N. Sharma, and I. Kaur, Tet. Lett., 2001, 42, 5315; (c) V. Bhalla, R. Kumar, M. Kumar, and A. Dhir, Tetrahedron, 2007,
63, 11153.
25. S. Rahman, H. Tomiyasu, H. Kawozie, J.-l. Zhao, H. Cong, X.-L. Ni, X. Zeng, M. R. J. Elsegood, T. G. Warwick, S. J. Teat, C. Redshaw,
P. E. Georghiou, and T. Yamato. N. J. Chem., 2015.
26. O. Kundrat, V. Eigner, P. C. Rínova, J. Kroupa, and Pavel Lhotak, Tetrahedron., 2011, 67, 8367.
27. H. Gampp, M. Maeder, C. J. Meyer, and A. D. Zuberbulher, Talanta, 1985, 32, 95.
28. J.-L. Zhao, H. Tomiyasu, X.-L. Ni, X. Zeng, M. R. J. Elsegood, C. Redshaw, S. Rahman, P. E. Georghiou, and T. Yamato, New J. Chem.,
2014, 38, 6041.
29. A. Casnati, P. Minari, A. Pochini and R. Ungaro, J. Chem. Soc. Chem. Commun., 1991, 19, 1413.
30. X.-L. Ni, S. Rahman, S. Wang, C.-C. Jin, X. Zeng, D. L. Hughes, C. Redshaw, and T. Yamato, Org. Biomol. Chem., 2012, 10, 4618.
31. J.-L. Zhao, H. Tomiyasu, X.-L. Ni, X. Zeng, M. R. J. Elsegood, C. Redshaw, S. Rahman, P. E. Georghiou, S. J. Teat, and T. Yamato, Org.
Biomol. Chem., 2015, 13, 3476.
32. (a) F. Camerel, S. Diring, B. Donnio, T. Dintzer, S. Toffanin, R. Capelli, M. Muccini, and R. Ziessel, J. Am. Chem. Soc., 2009, 131, 18177;
(b) Z. J.-H. Zhao, X. Chen, X. Wang, P. Lu, and Y. Yang, J. Org. Chem., 2009, 74, 383; (c) K. R. J. Thomas, M. Velusamy, J. T. Lin, C. H.
Chuen, and Y.-T. Tao, J. Mater. Chem., 2005, 15, 4453.
33. X. Feng, J.-Y. Hu, F. Iwanaga, N. Seto, C. Redshaw, M. R. J. Elsegood, and T. Yamato, Org. Lett., 2013, 15(6), 1318.
34. X. Feng, F. Iwanaga, J.-Y. Hu, H. Tomiyasu, M. Nakano, C. Redshaw, M. R. J. Elsegood, and Takehiko Yamato Org. Lett., 2013, 15(14),
3594.

More Related Content

What's hot

Новая химия под давлением. Александр Гончаров
Новая химия под давлением. Александр ГончаровНовая химия под давлением. Александр Гончаров
Новая химия под давлением. Александр Гончаров
Alexander Dubynin
 
Coupling_Carrots
Coupling_CarrotsCoupling_Carrots
Coupling_Carrots
Scott Shaw
 
Correlation between corrosion inhibitive effect and quantum molecular structu...
Correlation between corrosion inhibitive effect and quantum molecular structu...Correlation between corrosion inhibitive effect and quantum molecular structu...
Correlation between corrosion inhibitive effect and quantum molecular structu...
Al Baha University
 
Supramolecular chemistry
Supramolecular chemistrySupramolecular chemistry
Supramolecular chemistry
DaffodilsInJune
 

What's hot (19)

A04 06 0108
A04 06 0108A04 06 0108
A04 06 0108
 
Новая химия под давлением. Александр Гончаров
Новая химия под давлением. Александр ГончаровНовая химия под давлением. Александр Гончаров
Новая химия под давлением. Александр Гончаров
 
Chapter 05 stereochemistry at tetrahedral centers
Chapter 05 stereochemistry at tetrahedral centersChapter 05 stereochemistry at tetrahedral centers
Chapter 05 stereochemistry at tetrahedral centers
 
Fragmentation rules mass spectroscopy
Fragmentation rules mass spectroscopyFragmentation rules mass spectroscopy
Fragmentation rules mass spectroscopy
 
A Nano Capacitor Including Graphene Layers Composed with Doped Boron and Nitr...
A Nano Capacitor Including Graphene Layers Composed with Doped Boron and Nitr...A Nano Capacitor Including Graphene Layers Composed with Doped Boron and Nitr...
A Nano Capacitor Including Graphene Layers Composed with Doped Boron and Nitr...
 
Landcastle_UIUC_REUpaper
Landcastle_UIUC_REUpaperLandcastle_UIUC_REUpaper
Landcastle_UIUC_REUpaper
 
Graphene Supported Metal Oxide for Non-Enzymatic H2O2 Sensing - Crimson Publi...
Graphene Supported Metal Oxide for Non-Enzymatic H2O2 Sensing - Crimson Publi...Graphene Supported Metal Oxide for Non-Enzymatic H2O2 Sensing - Crimson Publi...
Graphene Supported Metal Oxide for Non-Enzymatic H2O2 Sensing - Crimson Publi...
 
Structural and magnetic properties of LiMn1.5Fe0.5O4 spinel oxide
Structural and magnetic properties of LiMn1.5Fe0.5O4 spinel oxideStructural and magnetic properties of LiMn1.5Fe0.5O4 spinel oxide
Structural and magnetic properties of LiMn1.5Fe0.5O4 spinel oxide
 
Detection and identification of Extractables and leachables from Pharmaceutic...
Detection and identification of Extractables and leachables from Pharmaceutic...Detection and identification of Extractables and leachables from Pharmaceutic...
Detection and identification of Extractables and leachables from Pharmaceutic...
 
Fragmentation rule MASS
Fragmentation rule MASSFragmentation rule MASS
Fragmentation rule MASS
 
Fragmentation Pattern of Mass Spectrometry
Fragmentation Pattern of Mass SpectrometryFragmentation Pattern of Mass Spectrometry
Fragmentation Pattern of Mass Spectrometry
 
Coupling_Carrots
Coupling_CarrotsCoupling_Carrots
Coupling_Carrots
 
Structure of polymer chains
Structure of polymer chainsStructure of polymer chains
Structure of polymer chains
 
Asymmetric synthesis ii
Asymmetric synthesis iiAsymmetric synthesis ii
Asymmetric synthesis ii
 
Vol. 1 (4), 2014, 112‒122
Vol. 1 (4), 2014, 112‒122Vol. 1 (4), 2014, 112‒122
Vol. 1 (4), 2014, 112‒122
 
Correlation between corrosion inhibitive effect and quantum molecular structu...
Correlation between corrosion inhibitive effect and quantum molecular structu...Correlation between corrosion inhibitive effect and quantum molecular structu...
Correlation between corrosion inhibitive effect and quantum molecular structu...
 
T15 Thermochemistry IB
T15  Thermochemistry IB T15  Thermochemistry IB
T15 Thermochemistry IB
 
Iron, cobalt and Nickel -ligand bonding in metallocene: Differentiation betwe...
Iron, cobalt and Nickel -ligand bonding in metallocene: Differentiation betwe...Iron, cobalt and Nickel -ligand bonding in metallocene: Differentiation betwe...
Iron, cobalt and Nickel -ligand bonding in metallocene: Differentiation betwe...
 
Supramolecular chemistry
Supramolecular chemistrySupramolecular chemistry
Supramolecular chemistry
 

Similar to Structural Studies on Pyrene-Based Blue-Light Emitting Molecules and Ion-Sensing Calixarenes - Project Report

Brent Godau Chem 499 Thesis
Brent Godau Chem 499 ThesisBrent Godau Chem 499 Thesis
Brent Godau Chem 499 Thesis
Brent Godau
 
A simple photochemical_method_to_synthes
A simple photochemical_method_to_synthesA simple photochemical_method_to_synthes
A simple photochemical_method_to_synthes
franciscocabrera94
 
Organic solar cells the exciting interplay of excitons and nano-morphology
Organic solar cells the exciting interplay of excitons and nano-morphologyOrganic solar cells the exciting interplay of excitons and nano-morphology
Organic solar cells the exciting interplay of excitons and nano-morphology
vvgk-thalluri
 
2010_NanoToday_Controlling nanoscale morphology in polymer photovoltaic devices
2010_NanoToday_Controlling nanoscale morphology in polymer photovoltaic devices2010_NanoToday_Controlling nanoscale morphology in polymer photovoltaic devices
2010_NanoToday_Controlling nanoscale morphology in polymer photovoltaic devices
Ximin He
 
Influence_of_UV_light_intensity_on_dielectric_beha.pdf
Influence_of_UV_light_intensity_on_dielectric_beha.pdfInfluence_of_UV_light_intensity_on_dielectric_beha.pdf
Influence_of_UV_light_intensity_on_dielectric_beha.pdf
Jaya Lakshmi
 
Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...
Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...
Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...
Scientific Review SR
 
PACI assignment.pptx
PACI assignment.pptxPACI assignment.pptx
PACI assignment.pptx
Ravikasera3
 
Ag (5 8) bonacickoutecky2001
Ag (5 8) bonacickoutecky2001Ag (5 8) bonacickoutecky2001
Ag (5 8) bonacickoutecky2001
hong-nguyen
 
Davey l1 macromolec-struc-anlys(1) lec 1
Davey l1 macromolec-struc-anlys(1) lec 1Davey l1 macromolec-struc-anlys(1) lec 1
Davey l1 macromolec-struc-anlys(1) lec 1
RANJANI001
 

Similar to Structural Studies on Pyrene-Based Blue-Light Emitting Molecules and Ion-Sensing Calixarenes - Project Report (20)

Brent Godau Chem 499 Thesis
Brent Godau Chem 499 ThesisBrent Godau Chem 499 Thesis
Brent Godau Chem 499 Thesis
 
Recent Advances in Photovoltaic Technology based on Perovskite Solar Cell- A ...
Recent Advances in Photovoltaic Technology based on Perovskite Solar Cell- A ...Recent Advances in Photovoltaic Technology based on Perovskite Solar Cell- A ...
Recent Advances in Photovoltaic Technology based on Perovskite Solar Cell- A ...
 
Alberto Tosato Thesis
Alberto Tosato ThesisAlberto Tosato Thesis
Alberto Tosato Thesis
 
Organic- Inorganic Perovskite Solar Cell
Organic- Inorganic Perovskite Solar CellOrganic- Inorganic Perovskite Solar Cell
Organic- Inorganic Perovskite Solar Cell
 
A simple photochemical_method_to_synthes
A simple photochemical_method_to_synthesA simple photochemical_method_to_synthes
A simple photochemical_method_to_synthes
 
A simple photochemical_method_to_synthes
A simple photochemical_method_to_synthesA simple photochemical_method_to_synthes
A simple photochemical_method_to_synthes
 
A simple photochemical_method_to_synthes
A simple photochemical_method_to_synthesA simple photochemical_method_to_synthes
A simple photochemical_method_to_synthes
 
First Paper
First PaperFirst Paper
First Paper
 
Pin mặt trời chất màu nhạy quang www.mientayvn.com
Pin mặt trời chất màu nhạy quang www.mientayvn.comPin mặt trời chất màu nhạy quang www.mientayvn.com
Pin mặt trời chất màu nhạy quang www.mientayvn.com
 
Organic solar cells the exciting interplay of excitons and nano-morphology
Organic solar cells the exciting interplay of excitons and nano-morphologyOrganic solar cells the exciting interplay of excitons and nano-morphology
Organic solar cells the exciting interplay of excitons and nano-morphology
 
Panushka UROP (1)
Panushka UROP (1)Panushka UROP (1)
Panushka UROP (1)
 
2010_NanoToday_Controlling nanoscale morphology in polymer photovoltaic devices
2010_NanoToday_Controlling nanoscale morphology in polymer photovoltaic devices2010_NanoToday_Controlling nanoscale morphology in polymer photovoltaic devices
2010_NanoToday_Controlling nanoscale morphology in polymer photovoltaic devices
 
Influence_of_UV_light_intensity_on_dielectric_beha.pdf
Influence_of_UV_light_intensity_on_dielectric_beha.pdfInfluence_of_UV_light_intensity_on_dielectric_beha.pdf
Influence_of_UV_light_intensity_on_dielectric_beha.pdf
 
Catalysis In Chemistry
Catalysis In ChemistryCatalysis In Chemistry
Catalysis In Chemistry
 
Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...
Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...
Study of Microstructural, Electrical and Dielectric Properties of La0.9Pb0.1M...
 
PACI assignment.pptx
PACI assignment.pptxPACI assignment.pptx
PACI assignment.pptx
 
Ag (5 8) bonacickoutecky2001
Ag (5 8) bonacickoutecky2001Ag (5 8) bonacickoutecky2001
Ag (5 8) bonacickoutecky2001
 
New Material:Perovskites presentation
New Material:Perovskites presentationNew Material:Perovskites presentation
New Material:Perovskites presentation
 
Davey l1 macromolec-struc-anlys(1) lec 1
Davey l1 macromolec-struc-anlys(1) lec 1Davey l1 macromolec-struc-anlys(1) lec 1
Davey l1 macromolec-struc-anlys(1) lec 1
 
Jackson,ChemMat,v15p1820
Jackson,ChemMat,v15p1820Jackson,ChemMat,v15p1820
Jackson,ChemMat,v15p1820
 

Structural Studies on Pyrene-Based Blue-Light Emitting Molecules and Ion-Sensing Calixarenes - Project Report

  • 1. This journal © The Royal of Dalton r Dalton Transactions REPORT Received 5th June 2015 www.rsc.org/dalton Structural Studies on Pyrene-Based Blue-Light Emitting Molecules and Ion-Sensing Calixarenes Thomas G. Warwick* and Dr. Mark R. J. Elsegood The awarding of the physics Nobel Prize to Shuji Nakamura and colleagues for their role in the development of the blue LED has drawn attention to the field of organic LED, or OLED design, where there is a keenness to develop blue-light emitting organic substrates with desirable emissive properties. Calixarenes meanwhile have been an important family of hosts in host-guest chemistry in recent decades, with well-established synthetic procedures and nomenclature. This wealth of literature makes calixarenes appealing compounds when considering solutions to new ion-sensing/molecular recognition-type problems. Single crystal X-ray diffraction experiments can provide absolute structure determinations of crystalline solids, and is therefore an invaluable technique when the properties of a material are highly dependent upon their structure, as with the above. Single crystal X-ray diffraction experiments were performed on six calixarenes and two pyrene-based blue-light emitting materials, submitted by a collaborator in Japan, and ten additional structures. The results include both structures which appear to be highly, and poorly optimized for their intended purpose, the reasons for which are discussed here. Most saliently, irrefutable proof of 2:1 host-to-guest Cl - binding behavior for a thiacalix[4]arene has been discovered, and these results have been included in a 2015 research paper. Introduction Calixarenes The chemistry of calix[n]arenes has been thoroughly explored subsequent to the work of Gutsche and co-workers, who were the first to tabulate assorted general synthetic routes, their rudimentary chemical properties, and nomenclature in the late 70s.1 Success in ―mapping‖ out calixarene chemistry met with an abundance of prospective applications for these new materials and spawned a research area in which Gutsche remains a seminal figure, with collaborative papers published over a period spanning four decades,2a~b and more recently, three books collating the key concepts resulting from this work.3a~c In the broadest terms then, calixarenes are cyclic oligomers formed by *Student Number B126328, Department of Chemistry, Loughborough, Leicester, LE11 3UN. Email: t.warwick-11@student.lboro.ac.uk. Supervisor Email: m.r.j.elsegood@lboro.ac.uk †Department of Applied Chemistry, Faculty of Science and Engineering, Saga University, Honjo-machi 1, Saga 840-8502, Japan. E-mail: yamatot@cc.saga-u.ac.jp †Herein we use R? to denote compounds external to our experiments that are available in the literature. CCDC deposition numbers etc. are available in appendix ii †All figures generated by the author in XP interactive molecular graphics, available as part of the SHELXTL suite released under license by BRUKER 21 †Non-standard structure codes, TY?, are used at the request of the supervisor for the sake of cross-referencing and internal consistency
  • 2. This journal © The Royal of Dalton r acid- or base-catalysed condensation of a finite number of simple aromatic alcohols, such as phenol or resorcinol, with an aldehyde, wherein water is eliminated. [n] is a variable denoting the number of aromatic alcohol monomers incorporated into the oligomer. The aldehyde, which effectively forms a bridging unit between aromatic alcohols, may be switched to (say) elemental sulfur, to afford sulphur-bridged thiacalix[n]arenes,4 or to a suitable dielectrophile to afford oxacalix[n]arenes.5 Additionally, more complex derivatives of the starting aromatic alcohols may be chosen to yield functionalized calixarene products, with the conformation, as well as the connectivity of these starting materials affecting the ultimate properties of the products. Calixarenes therefore are highly tuneable molecules, with the interplay of these controllable parameters offering a means to synthesize calixarenes which are highly optimized for their purpose. These purposes include various types of catalysis, such as α-olefin polymerization and lactone ring opening,6 in which calixarenes with novel bridging units, e.g. dimethyleneoxa (-H2COCH2-), coordinated to vanadium metal centers are employed. Calixarenes bearing Lewis base moieties can access the d orbitals of transition metal ions/centers,7 whilst the size of the central cavity of the calixarene may be fine-tuned to optimize specificity towards species of a particular size, and so calixarenes are often employed as ion- sensors/purification agents.8 These molecular recognition-type properties of calixarenes of course have biomedical applications too, and recently, phosphonate-derivatized thiacalix[4]arenes were demonstrated to successfully inhibit Protein Tyrosine Phosphatase, an enzyme which catalyzes the dephosphorylation of phosphotyrosine protein residues, in bubonic plague-causing Yersinia bacteria, in which these Protein Tyrosine Phosphatases are the primary cause of the plague symptoms.9 Pyrene-Based Blue Light Emitters The first evidence of the photovoltaic, or PV, effect occurring in an organic material was published in 1906 by Pochettino who, inspired by Edmond Becquerel, was performing similar experiments on anthracene crystals.10 The conversion efficiency was however poor and so the field of photovoltaics remained focused on exploiting the PV properties of inorganic materials, culminating in the first solar cell in 1954.11 In the mid-80s, Tang reported an organic PV cell with a relatively high (ca. 1%) conversion efficiency, which was a consequence of the cell‘s unique, two-layer design.12 Here, two organic layers were placed in-between two electrodes of different work functions, and excitons, produced by the absorption of photons by one of the layers, dissociated towards the low-work function cathode, generating a potential difference.13 Of course, Organic Light Emitting Diodes, or OLEDs, operate via the exact reverse principle; electrons and electron holes are introduced to the organic layers via application of an electric current, and Coulombic forces bring these together. Once both electrons and electron holes exist within a single layer (the emissive layer), recombination occurs to produce light with a wavelength inversely proportional to the energy separation of that material‘s Frontier Molecular Orbitals (its band gap). With the 2014 Nobel Prize for physics awarded to Shuji Nakamura for his role in the development of the blue LED,14 research into the development of organic molecules which emit blue light upon application of an electric
  • 3. This journal © The Royal of Dalton r field has been brought sharply into focus. Necessarily, blue emission bands are hard to access due to; the large band gap required to produce high energy light; low-energy excimer emission occurring when π-clouds on adjacent molecules approach each other at a distance less than ca. 4.5 Å resulting from π···π stacking;15 and homolysis of excited species whose recombination pathways include non-radiative processes. The latter issue presents the further condition of having to avoid, where possible, inclusion of weaker carbon-heteroatom bonds in the material‘s structure. Pyrene is a highly stable Poly Aromatic Hydrocarbon, PAH, with a long fluorescence lifetime of >100ns16 and high quantum yield, Φf, of 0.32 in cyclohexane17 whose emission from the S1 state is red, or bathochromically, shifted into the blue/visible region of the electromagnetic spectrum by rational substitution. Choice of substituent, and its location on the pyrene ring can effectively suppress π···π stacking/excimer emission in the solid state, and so pyrene derivatives are a promising class of compounds for use as emissive materials in blue OLEDs. These OLEDs, once fully optimized and in sufficient demand, may replace current Liquid Crystal Display technology, reducing the energy consumption of televisions, monitors and such like by circumventing the need for backlighting. Furthermore, displays may be printable18 and flexible,19 complimenting research into dye-sensitized solar cells. Aims Clearly, success or failure of both classes of materials is intimately connected to their solid state structure, which in turn affects their solid state properties. Determination of these structures can, in the case of the blue-light emitters, provide arguments as to why a prospective material emits (say) green or yellow light, by quantifying the distance between excimer-forming π···π stacked units. Similarly, unanticipated charge-transfer stabilization between proximal non-hydrocarbon substituents may be simultaneously proven and quantified. For calixarenes, solid state structure determination is a companion to spectrophotometric and NMR studies in assessing whether a host complex has, or has not captured a guest species. Moreover, a fully solved structure can yield precisely where host-guest interactions occur in multi-functional host complexes such as prospective catalysts. Precise data of this kind is essential for maintaining oversight on the frequent serendipitous results in these fields of research, and the ability to tabulate unexpected/chance results is arguably the single most beneficial contribution provided by a solid state structure determination. As both classes of materials readily form crystalline solids, single crystal X-ray diffraction experiments were to be performed on a variety of samples representing both classes of materials, provided by Takehiko Yamato† and co-workers. Each sample was to be subjected to data collection, correction, structure solution, modeling, refinement and validation in accordance with established crystallographic procedures. Parameters and variables governing these procedures were to be tabulated and presented as part of a results ―package‖. The crystal structures, and key structural features, were to be presented diagrammatically, textually and quantitatively, as part of this package. Adherence to established procedures and in-house styles were to be rigorously upheld to a) provide high-quality
  • 4. This journal © The Royal of Dalton r results, and b) provide a benchmark against which improvements in crystallographic knowledge and skill of the author could be compared. Experimental A general experimental procedure representing the most common choice of experimental conditions is given. Deviations from this procedure for each sample will then be presented sample-wise. The reader is directed to tables fully detailing experimental conditions for each sample provided in appendix i. General X-ray Crystallography Procedure Crystal data were collected at the ALS, station 11.3.1 on a Bruker APEX 2 and CCD diffractometer using silicon 111 monochromated X-radiation with ω-scans.20 Data were corrected for Lorentz and polarization effects and absorption, based on repeated symmetry-equivalent reflections, and structures solved by dual-space iterative methods.21,22 Structures were refined by full matrix least squares on F2 .21,22 H atoms were geometrically constrained (unless stated otherwise below) using a riding model, in which all hydrogen atom Uiso values set at 120% of that of the carrier atom, except for methyl and hydroxyl H atoms (set as 150%). Observed disorder was modeled as multiple sets of point atoms distributed over discreet locations with an occupancy sum of 1. Restraints were placed on geometries and Uaniso values of atoms in disordered moieties as appropriate. Sample-Specific Procedures In TY44 the chloroform solvent carbon and its adjoining H atom were modeled as fully disordered over two sets of positions without employing any restraints. In TY49 badly disordered solvent molecules were modeled as diffuse clouds of electron density using the Platon ―Squeeze‖ procedure.23 This gave a solution of three methanol molecules and one water molecule. The o-pyridyl group C(53) > N(6) and the t butyl group C(7) > C(10) were modelled as fully disordered over two sets of positions for the Me groups. Geometric and Uaniso restraints were used. H(1) and H(3) were freely refined, whilst H(5) was modelled as two partial point atoms, with each one belonging to one of the o- pyridyl disorder components. For TY51 the t butyl group containing C(17) on one pyrene ring was modelled as fully disordered over two sets of positions for the Me groups using geometric and Uaniso restraints. The badly disordered methanol molecule was modelled as a diffuse cloud of electron density using the Platon ―Squeeze‖ procedure.23 In TY53 diffraction data were collected on a Bruker APEX 2 CCD diffractometer using graphite monochromated Mo- Kα X-radiation with narrow slice, 0.3° ω-scans. The structure was solved by direct methods.21,22 The structure was heavily disordered overall. Two of the ethanol solvent molecules were modelled as point atoms, both of which are disordered over two positions. The third ethanol molecule was modelled as a diffuse cloud of electron density using the Platon ―Squeeze‖ procedure.23 The water molecule was modelled partially as a single, dehydrated point O partial
  • 5. This journal © The Royal of Dalton r point atom, and partially as a diffuse cloud of electron density using the squeeze procedure. The point atom constitutes the major disorder component, with occupancy of ca. 63%. The terminal ethyl ester carbon atom C(49) also exhibits two-part disorder. Geometric and Uaniso restraints were heavily applied. In TY54 diffraction data were collected on a Bruker D8 diffractometer, equipped with a PHOTON 100 CMOS detector, with shutterless ω-scans. Full matrix least squares refinement on F2 was performed as a two-component non-merohedral twin, on ‗twin5‘ data,21,22 with the major component at 53.5(2)%. Rotation is about axis [0 1 -1]. For TY55 diffraction data were collected on a Bruker D8 diffractometer, equipped with a PHOTON 100 CMOS detector, with shutterless ω-scans. Again, the structure was badly disordered. The benzyl group containing C(58) > C(64) was modeled as fully disordered over two positions. The 1-methylimidazole group attached to C(53), and C(53) itself, were also modeled as fully disordered over two positions. The ethanol molecule containing O(5) was modeled as fully disordered over three positions, with an occupancy sum restraint employed. Geometric and Uaniso restraints were heavily applied. For TY57 diffraction data were collected on a Bruker APEX 2 CCD diffractometer using graphite monochromated Mo- Kα X-radiation with narrow slice, 0.3° ω-scans. The structure was solved by direct methods.21,22 NH atoms were freely refined using geometric restraints. In TY59 diffraction data were collected on a Bruker D8 diffractometer, equipped with a PHOTON 100 CMOS detector, with shutterless ω-scans. Two t butyl groups on calixarenes were modelled as disordered over two sets of positions for the Me groups. Two n butyl chains in the cations exhibit some signs of disorder, but this was not modelled. The chloroform solvent molecule was modelled as fully disordered over two sets of positions. NH atoms were freely refined using geometric restraints. For GW114 diffraction data were collected on a Bruker APEX 2 CCD diffractometer using graphite monochromated Mo-Kα X-radiation with narrow slice, 0.3° ω-scans. The carbazole ring N(3) > C(29) was modeled as disordered over two sets of positions. Both substituted pyridine molecules in the asymmetric were also modeled as disordered over two sets of positions. Geometric and Uaniso restraints were used. In GW123 diffraction data were collected on a Bruker APEX 2 CCD diffractometer using graphite monochromated Mo-Kα X-radiation with 0.5° ω- scans. A single t butyl group C(29) > C(30) was modelled as fully disordered over two sets of positions for the Me groups. For GW115 diffraction data were collected on a Bruker APEX 2 CCD diffractometer using graphite monochromated Mo-Kα X-radiation with narrow slice, 0.3° ω-scans. The structure was solved by direct methods.21,22 The chloroform hydrogen H(35) and NH hydrogen atoms H(6) and H(8), were freely refined using geometric constraints. In GWALS11 diffraction data were collected on a Bruker D8 diffractometer, equipped with a PHOTON 100 CMOS detector, with shutterless ω-scans. For GWALS12 diffraction data were collected on a Bruker D8 diffractometer, equipped with a PHOTON 100 CMOS detector, with shutterless ω-scans. No solvent of crystallization was present in the asymmetric unit, which exhibited full molecule, two-fold disorder. This was modelled using geometric restraints, and the major disorder component refined to have an occupation of ca. 66%. For GWALS13, GWALS14, GWALS15 and GWALS16, diffraction data were collected on a Bruker D8 diffractometer, equipped with a PHOTON 100 CMOS detector, with shutterless ω-scans. No disorder or solvent of
  • 6. This journal © The Royal of Dalton r crystallization was present in the asymmetric unit and NH atoms were freely refined where present. For RSH29 diffraction data were collected on a Bruker APEX 2 CCD diffractometer using graphite monochromated Mo-Kα X- radiation with narrow slice, 0.3° ω-scans. Hydroxyl and water H atoms were located by inspection of the difference maps and electron density peaks. H atoms H(2), H(4), H(10), H(18), H(20), H(25), H(25A), H(26), H(26A), H(27) and H(27A) were freely refined with geometric restraints. Programs used were Bruker APEX 2 (data collection), Bruker SAINT (integration and cell refinement), SHELXTL suite (solution, refinement and graphics), SHELXLe (refinement), and local programs. These data, along with abundant supplementary figures and descriptions of each sample generated by the author are available at ws8.lboro.ac.ukcm-xray-datawarwick or by request. Crystal data will become available free of charge at www.ccdc.cam.ac.uk/data_request/cif when and if structures are published. All of these data have been separately archived within the university and on the accompanying DVD ROM. Results and Discussion TY57 & TY59 Crystal structures of calixarene samples were determined in order to asses their ability to bind guests through hydrogen bonding. For TY59 the asymmetric unit (Fig. 1) contains four calixarenes, two chloride anions, two tetra- n-butylammonium cations, two acetonitrile and one chloroform, molecules of crystallisation. Within each of the four calixarenes are pairs of N—H···O H-bonds between urea moieties to a single carbonyl O atom. Looking down on the square-shaped S4 planes of the four unique calixarenes, three are approximately geometrically aligned, whist one, containing S(1A), is slightly twisted. Most importantly, pairs of calixarene molecules are linked via four H-bonds between both urea NH moieties on each calixarene and a chloride ion. The structure provides unequivocal proof of 2:1 host-to-guest binding properties of a 1,3-alternate thiacalix[4]arene bearing two phenyl-ureido moieties bound, through ether O atoms, to the p-position of two opposing aryl rings which are part of the central calixarene cavity (herein referred to as oligomeric rings). A full host-guest unit is shown in Fig. 2a. These results corroborate UV-vis analyses performed by Kumar and co-workers, who demonstrated 2:1 host-to-guest binding behaviour in an analogous calix[4]arene.24a Structural differences between this analogous calix[4]arene, R1†, and TY59 include CH2 instead of S bridges; two n propoxy instead of benzyloxy substituents bound to oligomeric rings; and most saliently, p-nitrophenyl-ureido, instead of phenyl-ureido, moieties bound through ether O atoms. Structurally, TY57 (Fig. 2b) differs from TY59 only through bearing p-tolyl-ureido, not phenyl-ureido, moieties. Its crystal structure is however absent of any chloride anions and solvent of crystallisation. Such was the significance of our X-ray diffraction studies on these samples, they have been included in a paper due for publication in the New Journal of Chemistry later in 2015.25 It builds the work of Kumar24a~c and a study by Lhoták and co- workers which demonstrates the necessity of both aryl-ureido moieties to be bound to the p-, and not m-, or both
  • 7. This journal © The Royal of Dalton r p- and m- positions of the oligomeric rings.26 Our study assesses the effect of electron donating or withdrawing groups bound to the p- position of the aryl-ureido moieties, on the anion binding ability of a series of 1,3-alternate thiacalix[4]arene-based receptors, of which TY57 and TY59 and R3 are a part. Association constant data (Table 1) for these receptors clearly demonstrate the enhancing effect of electron withdrawing substituents attached to the p- position of the R-p-aryl-ureido moieties, on the acidity of the urea hydrogen atoms. R3, which was not subject to X-ray analysis, shares the same ‗scaffold‘ as TY57 and TY59, and so these data show an increasing binding ability of 1,3-alternate thiacalix[4]arene receptors towards Cl- with increasing electronegativity of the R substituent on the R-p-aryl-ureido chains. This occurs as more electronegative substituents increase the acidity of the urea hydrogen Fig. 1 The asymmetric unit of sample TY59 highlighting its anion-binding properties. Non-NH hydrogen atoms are omitted for clarity. H- bonds are shown as dashed lines. Host TY57 TY59 R3 R1a R2b Guest Cl- Cl- Cl- Cl- Cl- R-p-aryl-ureido; R= tolyl none trifluoromethyl nitro nitro Association constant 1,500±110c 2,940±210c 6,590±460c 3,470,000±90,000d 270±70e Table 1 Selected binding constant data from24a,25,26 highlighting the electronic effects of p- substituents on binding ability. a Receptor 1 from24a as described above. b Receptor 9b from26 with p-nitrophenyl-ureido moieties substituted at the m- position (i.e. both leaning off to one side) of oligomeric rings. c Calculated from Δδ of NH(x) protons by titration of host at 4.0 x 10-3 M with the guest, in CDCl3–CD3CN (10:1, v/v) at 298K, 300MHz.25 d Calculated from Δabsmax, at 300 & 342nm, of host at 5 x 10-5 M upon incremental addition of 0–100 mol. equivalents of guest at 298K, using non-linear regression analysis program SPECFIT.24a,27 e Calculated from Δδ of NH(x) protons by titration of host at 0.5–2.0mM with guest to afford host:guest ratios 1- 20:1, in DMSO-d6 at 298K, 300MHz.26 Guests added as tetra-n-butylammonium (TBA) salts.24a,25,26
  • 8. This journal © The Royal of Dalton r Fig. 2a (left) The full bifurcated H-bonding environment of Cl(1) is shown for TY59 with the tetra-n-butylammonium cation observed approaching from behind. 2b (right) The asymmetric unit of TY57 is shown from the other side, highlighting the continuous parallel arrangement of urea moieties which runs continuously along the crystallographic b axis to give an extended, single helical H-bond array. Non- H-bonding H atoms are omitted for clarity. atoms, strengthening the H-bonds between the host and the guest through stabilization of the conjugate base. This is verified by H-bond length data obtained via our diffraction experiments. For the TY59 fragment shown, the chloride anion is held in place by H-bonds: N(1)—H(1)···Cl(1), N(2)—H(2)···Cl(1), N(3A)—H(3A)···Cl(1) and N(4A)— H(4A)···Cl(1) of A···H length 2.26(2), 2.81(3), 2.87(3) and 2.30(2) Å respectively. TY57 meanwhile has intermolecular N(3)—H(3)···O(5A) and N(4)—H(4)···O(5A) H-bonds of A···H length 2.13(3) and 2.35(7) Å respectively. These slightly shorter distances are likely a consequence of the diminished donor capability of urea N atoms due to the presence of the immediately adjacent, electron donating tolyl, instead of phenyl substituents. For completion, the low association constant for receptor R2 is a consequence of the nitrophenyl-ureido moieties being attached to the m- positions of oligomeric rings, forbidding any cooperative H-bonding within each calixarene by lengthening the A—D distances. This in turn allows only facile 1:1 host-to-guest binding behavior for this receptor. The 1000-fold increase in selectivity towards Cl- of R1 compared to TY57 and TY59 meanwhile, is most likely due to the presence of n propoxy groups. These would have less severe steric impositions than the benzyloxy groups of both TY samples, perhaps therefore allowing a more favourable enclave for the captured Cl- anion. Without diffraction data for both R1 and R2 however, these notions remain debatable.
  • 9. This journal © The Royal of Dalton r TY53 & TY55 Both samples are 1,3-alternate thiacalix[4] arenes bearing two 2-(1-methyl-1H-imidazolyl)methoxy groups bound to the p- positions of oligomeric rings, differing by the remaining oligomeric rings being bound to benzyloxy groups for TY55 (Fig. 3a) and 2-(ethyl ethanoyl)oxy groups for TY53 (Fig. 3b). Both samples crystallize in space group P1 , and exhibit remarkably similar disorder. Firstly, TY55 exhibits two-part disorder about a single C(58) > C(64) benzyl substituent (major occupancy 50%), whilst TY53 exhibits two-part disorder about a single C(49) atom in an ethanoyl moiety (major occupancy 56%). A single C(53) > C(57) 1-methyl-1H-imidazole ring in TY55 also exhibits two-part disorder. As this 1-methyl-1H-imidazole ring moves between the two positions, cavities are formed in the positions temporarily left unoccupied by the 1-methyl-1H-imidazole ring. It is in these two cavities that the disorder components of one of the ethanol molecules, which is three-fold disordered, are found. The two ethanol disorder components containing O(5X) and O(5Y) are permitted when the minor component C(54X) > N(4X) 1-methyl-1H- imidazole ring is present. The ethanol disorder component containing O(5) is permitted when the major component C(54) > N(4) 1-methyl-1H-imidazole ring is present. This arrangement of disorder components in turn permits a single O(5)—H(5)···N(4) bond of length 2.75(2) Å. The connection between 1-methyl-1H-imidazole disorder and ethanol disorder was however not modelled. Instead this ethanol molecule was modelled as fully disordered over all three positions irrespective of 1-methyl-1H-imidazole ring site occupation using an occupancy sum restraint. The major component occupancy for the 1-methyl-1H-imidazole ring is 56% whilst the restrained major component occupancy of the ethanol molecule is 42%. There is a remaining non-disordered ethanol molecule within the asymmetric unit, and all disorder components are shown (Fig. 3c). Two of the three ethanol molecules in sample TY53 were modelled as two-fold disordered with both major and minor disorder components of both ethanol molecules H-bonding, albeit through different H atoms, to N atoms in the immediately adjacent 1-methyl-1H-imidazole rings C(42) > N(2) and C(51) > N(4). A single disordered water molecule is also present in the asymmetric unit through which ethanol disorder components also H-bond to form centrosymmetric dimers, in which calixarenes approach one another head-to-head (Fig. 3d). The remaining ethanol molecule was modelled as a diffuse cloud of electron density using the Platon ―Squeeze‖ procedure.23 This disorder is surprising, given that an analogous 1,3-alternate thiacalix[4]arene, R4, bound at the p- position of all four oligomeric rings by 2-(1-methyl-1H-imidazolyl)methoxy groups, refined to completion with an R-factor of 4.01% without the use of restrains, and showed no signs of disorder.28 By comparison, R-factors of TY53 and TY55 are less good, at 6.84 and 9.18% respectively. In R4 all four 1-methyl-1H-imidazole rings lie away from the central calixarene cavity, whereas in both TY samples, the 1-methyl-1H-imidazole ring C(42) > N(2) nestles within the central calixarene cavity, parallel to the two flanking oligomeric rings. Remarkably, this has little effect on the dimensions of the S4 units which remain square-shaped, and have areas 30.714 Å2 for R428 and 31.147 A2 for TY53. Inclusion of 2-(ethyl ethanoyl)oxy and benzyloxy groups does seem to affect the shape of the overall
  • 10. This journal © The Royal of Dalton r calixarene structure however with both TY53 and TY55 having a more conical shape than R4 , whose cuboid shape and tetrahomosubstitution with lone pair-donating moieties is essential in allowing 1:2 host-to-guest binding of transition metals.28 TY49 Oxacalix[3]arene derivatives whose electron-accepting/donating substituents all reside on one side of the O3 cavity can selectively bind anionic/cationic guests with C3 symmetry,29 and TY49 (Fig. 4a) is an example of such a calixarene. The three oligomeric rings are functionalized through the p- position with [N-(2-Pyridinyl)acetamyl]oxy Fig. 3a (top left) The asymmetric unit of TY55 with only major disorder components shown. 3b (top right) The asymmetric unit of TY53 with only major disorder components shown. 3c (bottom left) Sample TY55 with all disorder shown. 3d (bottom right) The H-bonding array observed in sample TY53. Non-H-bonding H atoms are omitted for clarity.
  • 11. This journal © The Royal of Dalton r groups, two of which reside within the central calixarene cavity, whilst the cavity itself adopts a collapsed, or squashed conformation to facilitate a single intramolecular N(3)—H(3)···O(8) H-bond, of length 2.29(2) Å. So distorted is the cavity that a π···π interaction is observed, between oligomeric rings C(13) > C(18) and C(25) > C(30), with a distance of 3.87 Å. The o-pyridyl groups C(46) > N(4) and C(53) > N(6) meanwhile also align face-to- face, but at a distance of 4.80 Å which is too far for π···π stacking. This conformation appears to give rise to a very favourable self-assembly of molecules in the solid state, with a further intermolecular C—H···O H-bond, bond of length 2.28 Å, occurring between the m-pyridyl H(56) atom and a carbonyl O(7) atom (Fig. 4b & c). The benzyloxy analogue, R5,30 of TY49 adopts much more open conformation by comparison. For example, the narrowest angle between the three cavity-located O atoms is ∠O(1)-O(2)-O(3) = 47.71° for TY49 whilst the analogous angle for R5 is 52.81°.30 This wider angle necessitates greater π···π distances between oligomeric rings, with the closest distance between any two being 4.51 Å. Moreover, the angle between the two planes of the π···π-stacked C(13) > C(18) and C(25) > C(30) rings in TY49 is 23.97°, whilst the two closest oligomeric rings in R5 are far less parallel in their alignment, with the angle between their respective planes being 41.35°.30 A tentative conclusion therefore is that the cation binding ability of TY49 may be diminished by competitive intramolecular H-bonding between potential binding sites, and poorly optimized cavity geometry. Fig. 4a (left) Receptor TY49 showing the labelling scheme and the distorted central cavity. π···π stacking occurs vertically in this view. 4b (middle) The intra- and intermolecular H-bonding of receptor molecules, highlighting the ‗dog-legged‘ [N-(2-Pyridinyl)acetamyl]oxy substituent. 4c (right) A packing plot for receptor TY49. Non-H-bonding H atoms are omitted for clarity.
  • 12. This journal © The Royal of Dalton r TY44 This sample is a 1,3-alternate thiacalix[4]arene substituted with two pyridyl-4-methoxy groups, and two benzyloxy groups (Fig. 5a). This receptor is one of a family or receptors, which were studied elsewhere, in which the pyridyl-4- methoxy groups are replaced by pyridyl-3- or pyridyl-2-methoxy groups.31 i.e. the position of the pyridyl N atom changes with each member of this family, resulting in some measurable change in binding properties of the members. The TY44 thiacalixarenes align in columns running parallel to the crystallographic c axis, ca. 45° from the b axis, with chloroform solvent molecules of crystallization residing in voids between any four thiacalixarenes, generated by the observed vertex-to-vertex packing of the square-shaped S4 units, which all lie in a common plane. This packing affords S···S interactions of length 3.32–3.63 Å between sulfur atoms on adjacent thiacalixarenes (Fig. 5b). In contrast, the pyridyl-2-methoxy-substiuted analogue, R6, packs with offset S4 planes, occluding any S···S interactions. The S···S distances in R6 are all 4.25 Å.31 Both TY44 and R6 however pack such that adjacent thiacalixarene molecules have their pyridyl substituents aligned in opposite directions along the normal to the S4 plane, probably to minimise lone pair interactions between N atoms on adjacent molecules. This, in the case of TY44, allows molecules to pack more closely, permitting the S···S interactions. Some C—H···π interactions are observed between chloroform molecules and the oligomeric rings of TY44. In particular, the non-disordered chloroform molecule exhibits a C(67)—H(67)···π{C(21) > C(26)} interaction of distance 3.10 Å. The disordered chloroform meanwhile makes one of two mutually exclusive Fig. 5a (left) A thermal ellipsoid plot of receptor TY44, highlighting the substituent conformation about the calixarene cavity. H atoms and solvent of crystallization are omitted for clarity. 5b (right) A packing plot of TY44 with S···S contacts drawn. Non-chloroform H atoms are omitted for clarity. Thermal ellipsoids are drawn at the 50% probability level.
  • 13. This journal © The Royal of Dalton r C—H···π interactions, with the major component giving C(68)—H(68)···π{C(11A) > C(16A)} and the minor component giving C(68X)—(H68X)···π{C(31) > C(36)} (Fig. 5b). These have distances of 3.02 and 3.09 Å respectively. The distance between N atoms in TY44 is long, at 8.29 Å, due to the pyridine rings lying outside of the central thiacalixarene cavity. This distance is comparable to the analogous distance in R6, which is 9.08 Å.31 The binding capability of TY44 is likely therefore to be poor, given that the upward-pointing N atoms would only be able to interact with host species through weak, singular N···Guest contacts, and not strong N,N‘···Guest···S,S‘ contacts demonstrated for R6,31 bearing in mind that, for pyridyl-2-methoxy substituents, the N occupies the o- position. Rotation of the pyridine ring about the methoxy bridge, which is possible in solution, would create therefore an extremely electron-rich, S-, O- and N-containing enclave, suitable for small cationic species with accessible orbitals for electron donation. TY51 Moving on to the blue-light emitting samples, TY51 is formally 1,2-di-(7-t-butylpyrene)-3,4,5,6-tetraphenylbenzene (Fig. 6a). The two 7-t-butylpyrenyl and four phenyl substituents align in a propeller-like fashion around the central arene ring C(41) > C(46), with the two pyrene rings aligning without overlap to avoid unfavourable steric interactions between their respective t butyl groups. The torsion angles of the arene substituents relative to the central ring are summarized in Table 2. Ring Torsion angle, in° C(1) > C(16) 80.11(8) C(21) > C(36) 86.97(8) C(47) > C(52) 75.96(12) C(53) > C(58) 65.19(13) C(59) > C(64) 67.61(15) C(65) > C(70) 70.34(16) Table 2 Torsion angles of arene substituents relative to the central arene ring C(41) > C(46) of TY51. The central arene rings all align with their C6 axes parallel to one another, with the C(47) > C(52) ring nestling between the two pyrene rings on adjacent molecules (Fig. 6b). This permits multiple C–H···π interactions which lie in the range of 2.56–3.39 Å. No π···π stacking is present in the crystal structure, however, due to the steric impositions of the propeller-like topology of the individual molecules. Only the central C(47) > C(52) rings are aligned in such a way as to allow π···π stacking between adjacent molecules, however the centroid···centroid separation of these rings is too large, at 7.54 Å, to allow π···π stacking to occur.15 It is well known32a~c that π···π stacking promotes excimer formation between proximal molecules, and that these excimers, by definition, emit lower energy photons than the original monomeric species. As such, excimer formation between prospective blue-light emitting molecules
  • 14. This journal © The Royal of Dalton r undermines any optimization techniques hitherto employed, as emissive wavelengths are bathochromically shifted towards greener wavelengths, making the solid-state structure of these materials somewhat critical to their success. Thin films of the prospective blue-light emitter 1,3,5,9-tetra(p-methoxyphenyl)-7-t-butylpyrene, R7, with the comparable pyrene centroid···centroid separation to that of TY51, 8.57 Å, was shown to emit blue light of wavelength 471 nm upon application of an electric current.33 Moreover, these emissions occurred with a high Φf of 0.48 in DCM at 298K, and a small stokes shift of 2 nm, indicating that π···π stacking was neither happening in the solid or solution state.33 It may be concluded therefore that, assuming substituent electronic effects are minimal, TY51 is a promising candidate material for blue OLEDs. TY54 In contrast to TY51, TY54 does show significant π···π stacking between adjacent molecules. TY54 itself is a pyrene- derivatized molecule, bearing two phenyl substituents at the 1, and 3 positions, a t butyl group at the 7 position, and naphthalene rings bridged by imine groups to the 4, & 5 positions, and the 9, & 10 positions (Fig. 7a). The large torsion angles between the phenyl rings and the central pyrene, which are of the order of those observed in TY51, and the t butyl group, are not able to supress close approach of the flat naphthyl ‗spines‘. Hence adjacent molecules pack with their pyrene cores co-aligned in parallel planes (Fig. 7b). The pyrene core of each molecule is approached from above by the ring C(1) > C(6) on one immediately adjacent molecule, and from below by the ring C(11) > (16) on a second, immediately adjacent, molecule. The centroid···centroid distances of these two π···π interactions are 3.82 and 3.75 Å respectively. Furthermore, this π···π bonding array is observed at each molecule, running continuously Fig. 6a (left) Emitter TY51 with the full labelling scheme shown. 6b (right) A packing plot of TY51 highlighting the steric impositions of the propeller-like topology.
  • 15. This journal © The Royal of Dalton r Fig. 7a (top) emitter TY54 with a full labelling scheme shown. 7b (bottom left) A packing plot highlighting the π···π stacking in TY54. 7c (bottom right) a side-on view of TY54 emphasizing the twisted nature of the imine-bridged naphthalene rings. through each plane. Molecules in adjacent planes meanwhile pack with their t butyl groups approaching one another, and their two phenyl substituents approaching one another. Where molecules in two adjacent planes‘ phenyl groups approach one another, a single, very weak π···π interaction of distance 4.52 Å is observed between the two C(41) > C(46) rings on the neighbouring molecules. Interestingly the two imine-bridged naphthyl substituents twist away from one another, about the axis which is the perpendicular bisector of the C(2)—C(3) and C(12)—C(13) bonds. The angle between the arene rings C(1) > C(6) and C(11) > C(16) about this axis is 3.9(2)°. This behavior is radically different from that observed in an analogous compound, R8, which bears two p-formylphenyl substituents instead of phenyl substituents, and two imine-bridged benzene substituents instead of naphthalene substituents. In R8, the pyrene core is bent in on itself about the same axis, giving the molecule a characteristic banana shape. For comparison, using the labeling scheme of
  • 16. This journal © The Royal of Dalton r TY54, the acute angle between the planes of the C(22) > C(27) and C(29) > C(34) rings in R8 is 13.6°.34 This difference in conformation is not reflected however in the π···π distances, which remain short for R8, at 3.58 Å.34 This is shorter than the π···π distances observed in TY54, and so the p-formylphenyl substituents are more co-planar with the pyrene core, with torsion angles of 51 and 58°,34 whilst the torsion angles of the phenyl substituents of TY54 are 75.77(11) and 80.92(13)°. Conclusion Single crystal X-ray diffraction techniques have been employed to determine the solid state structure of eight samples, representing two different areas of chemistry, in collaboration with Takehiko Yamato. The collaborative nature of this work necessitated completion to a publishable standard, and indeed experimental results for two of these eight determinations have been included in an advanced draft of a research paper. The most significant finding is irrefutable evidence of 2:1 host-to-guest binding behavior of a 1,3-alternate thiacalix[4]arene bearing phenyl- ureido and benzyloxy substituents, corroborating the results of a separate study which did not contain X-ray data.24a The remainder of our structural findings add to the current structure-property investigation literature available both for classes of compounds studied here. The most salient experimental results for each structure were abstracted from the abundance of quantitative data collected, and presented as a package tailored to the research areas of interest to our collaborator. Upon future publication of our X-ray work, Crystallographic Information Files, or CIFs, for all of the structures studied will be deposited with the Cambridge Structural Database and so become freely accessible through moiety and key-word etc. database interrogations. Whilst the different classes and sub-classes of compounds studied here undermine the validity of direct comparisons between them, comparisons to reference compounds highlight their currency. Researchers involved in calixarene host-guest chemistry, and OLED engineering are encouraged to engage with crystallographic collaborators as fully as possible, to reduce the amount of structure-property research published without X-ray structures, which seems oxymoronic. In particular, if more dedication as given to growing crystals of a host material with the guest entrapped, variable-temperature X-ray diffraction experiments could elucidate how a guest is released from the host. From this, something might be learnt of what subtle structural changes need to take place for a guest to successfully bind to a host. This may have particular significance for less rigid calix[3]arenes, and involve such things as reconfiguration or deformation of the calixarene ‗scaffold‘, repositioning of functional groups, and antagonistic steric effects imposed by a crystalline structures. Should such experiments afford any useful data, materials may be optimized during synthesis for better function in either the solution state, where (say) hosts are added to complex with chromate, or other guests, which are harmful to aquatic life. Solid state optimization meanwhile could allow an improvement in the selectivity of ion-selective electrodes or synthetic membranes.
  • 17. This journal © The Royal of Dalton r Additional Work Structures Upon completion of all relevant samples submitted by the Takehiko Yamato group, refinement, modelling, validation and description were performed on/for ten additional samples. GW114, GW115, GW123, GWALS11, GWALS12, GWALS13 GWALS14, GWALS15, GWALS16 and RSH29 were studied as the proficiency of the author in these steps allowed increasingly rapid turnaround of X-ray results. These samples possess radically different chemistry to the ‗TY’ samples, and that suggested by the title of this report however. They are therefore excluded from the main discussion. A considerable investment of time (which is not represented by this disproportionately small discussion) and development of the crystallographic skill of the author resulted from these additional studies. The diverse range of structures presented more challenging, or even entirely new ‗classes‘ of crystallographic disorder, a full discussion of which cannot be presented here for reasons of brevity. For example, GW114 contained two novel disorder scenarios which required modelling. Firstly, full- molecule disorder about a carbazole substituent mandated the splitting of the entirety of this substituent into two restrained disorder components occupying separate regions in space. A before-and-after for the modelling of this disorder is shown in Figs. 8a & b. Fig 8a (top left) a thermal ellipsoid plot of GW114 showing the undesirable prolate displacement ellipsoids of the carbazole carbons. These are indicative of two-fold disorder. Fig 8b (top right) the carbazole substituent has now been modelled split over two sets of positions for the constituent atoms, which possess much more satisfactory displacement ellipsoids. Fig 8c (bottom left) The novel disorder corresponding to overlapping pyridyl N and C atoms is emphasized. Thermal ellipsoids are drawn at 50% probability and H atoms are omitted for clarity.
  • 18. This journal © The Royal of Dalton r The cause of this disorder was related to the second class of disorder present, in which the entire molecule could adopt one of two orientations during the crystal growth phase. For this particular case, the result was an electron density map in which the atomic position of N(1) is partially overlapped by C(4) and vice versa. Looking at Fig 8c it can be seen that these positions correspond to an imaginary C2 axis along the C(2)—F(2) & C(5)—F(5) bonds, in one molecule, and the C(2A)—F(2A) & C(5A)—F(5A) bonds in the second molecule (with both of the two molecules in the asymmetric unit experiencing this disorder, where ―?A‖ is used to distinguish between molecules). 2-fold rotation about this axis affords a major disorder component, where, focusing on only one of the molecules, the atomic position of N(1) is partially occupied by C(4X), and a minor disorder component, where the atomic position of C(4) is partially occupied by N(1X). Necessarily, the F(4) atom bound to C(4) is also disordered about this axis. It was found that the disorder observed in the carbazole substituent was due to two favorable, mutually exclusive C—H···F interactions between the disordered carbazole substituent, and either F(4A) in one adjacent asymmetric unit, or F(2) in another. Obviously, F(4A) is only present at the site which permits this C—H···F interaction when the entire F4A-containing molecule adopts the appropriate orientation, hence the connection. GW115 is another example of a complex modelling procedure, due to its extensive H-bonding array. Here, each macromolecule (Fig 9a) was H-bonded to three other macromolecules via three unique sets of H-bonds. These are as follows: The first and second macromolecules are H-bonded to one another by two bifurcated H-bonds between hydrogen atoms H(25B) and H(27B) on one macromolecule, to F(4) on the other macromolecule, and vice versa. Between the first and third is a single N—H···N H-bond between H(6) on the first macromolecule and N(2) on the third macromolecule. Fig 9a (left) GW115 is shown. Each macromolecule is H-bonded to a single chloroform atom. Fig 9b (right) the complex H-bonding environment of a single macromolecule. Here H atoms are omitted for clarity if not involved in H-bonding.
  • 19. This journal © The Royal of Dalton r Lastly, between the first and the fourth is a pair of parallel H-bonds. These are between O(2) on the first macromolecule to H(14) on the fourth macromolecule, and H(8) on the first macromolecule to N(3) on the fourth macromolecule. This H-bonding (Fig 9b) array is observed at each macromolecule, generating a nexus of intermolecular H-bonds. Work on these samples was of course performed to the same standard as the TY samples, and results were delivered their respective submitters in exactly the same package. These data have too been archived, and crystal data is included in appendix i. Guides Three documents were written by the author, for the intended use of future X-ray project students. The first, is entitled ―Basic SHELXL Commands for Small Molecule Refinement‖ is a tabulation of SHELX (structure modelling) commands learned throughout the entirety of the project. Commands are presented with their appropriate syntax and function explained in short, digestible paragraphs, occasionally taking the form of short tutorials for the commonest tasks. Examples of simplified instruction files with the necessary entries required for a command to work are also provided where appropriate. All of the commands are presented with useful tips for efficient, tidy structure modelling as learned by the author, and some salient in-house protocols are presented for the sake of internal consistency. An introduction is also provided explaining how to download and set up SHELXL and its GUI SHELXLe, include these executables in a computer‘s environment variables for command prompt loading, and some basic commands for navigating the command prompt. It is hoped that such a section, and the document as a whole, will equip the reader with essential computational skills and program-specific knowledge to set up, manage, and operate this software with confidence. This guide is ca. 6,000 words in length. The second guide, entitled ―Commands for Probing Materials in XP Interactive Molecular Graphics‖ has the same basic layout as the above, but this time with XP (structure probing and graphics) commands described. Again, an intro detailing program setup is provided, as well as some essential information on file extensions. Due to the diverse combination of XP commands, tutorial-type descriptions are kept to a minimum. The most salient diagram generation commands are presented as short tutorial. This guide is ca. 9,000 words in length. The final guide is entitled ―Tutorial For Mapping the K Drive to an Off-Campus Machine‖. The K: drive is the university‘s X-ray network drive, which is usually accessed via a clumsy VPN client which limits users to uploading/downloading files one-at-a-time amongst other productivity-limiting frustrations. This guide instructs the reader on how to map the K: drive to their network-connected PC or laptop, from where they may access their work as if it were stored locally on their machine, and therefore ‗work from‘ the K: drive. Additionally, instruction is given on mapping the students U: drive, provided to all Loughborough University students.
  • 20. This journal © The Royal of Dalton r Dalton Transactions Template The template for this report was generated by retroconversion of a generic Dalton Transactions manuscript into .docx format, and has been made available for those who wish to use it. All additional work is available by request from the author or Dr. Mark R. J. Elsegood, and is available at ws8.lboro.ac.ukcm-xray-datawarwick where they are provided with their raw text versions for future additions, corrections and improvements. The attached DVD ROM also contains all work, supplementary and otherwise, completed during this project. Acknowledgement The Advanced Light Source is supported by the Director, Office of Science, Office of Basic Energy Sciences, of the U.S. Department of Energy under Contract No. DE-AC02-05CH11231. Dr. Mark Elsegood is thanked for his guidance on matters both practical and literary.
  • 21. This journal © The Royal of Dalton r Appendix i Experimental and Crystal Data Compound TY44 TY49 TY51 Formula C66H70N2O4S4·CHCl3 C57H66N6O9·3(COH4)·H2O C70H54·CH4O Mr [gmol-1 ] 1202.84 1093.30 927.17 Crystal system Triclinic Monoclinic Orthorhombic Space group 1P C2/c Pna21 Unit cell dimensions a [Å] 16.3706 (10) 26.1641 (11) 11.6362 (4) b [Å] 19.2771 (12) 15.4995 (6) 26.6850 (8) c [Å] 20.9368 (13) 28.5153 (11) 17.6959 (6) α [°] 83.800 (4) 90 90 β [°] 82.809 (4) 94.063 (3) 90 γ [°] 71.799 (4) 90 90 V [Å3 ] 6210.6 (7) 11534.8 (8) 5494.8 (3) Z 4 8 4 T [K] 100 100 100 Dx(calc) [g cm-3 ] 1.286 1.259 1.121 Radiation type Synchrotron Synchrotron Synchrotron Radiation λ [Å] 0.7293 0.7085 0.7749 Crystal size [mm3 ] 0.10 × 0.08 × 0.04 0.30 × 0.17 × 0.03 0.11 × 0.06 × 0.04 Extinction Coeff. N/A 0.0008 (2) 0.0066 (10) 2θmax [°] 75.0 66.2 55.2 Reflections measured 236971 113193 43351 Unique reflections 59903 22109 9808 reflections with I > 2σ(I) 45699 16974 8404 Transmission factors 0.966, 0.986 0.974, 0.997 0.992, 0.997 Rint 0.062 0.074 0.072 Number of parameters 1475 747 669 Number of restraints 0 407 51 R[F2 > 2σ(F2 )] 0.068 0.067 0.058 wR(F2 ) 0.201 0.205 0.156
  • 22. This journal © The Royal of Dalton r Compound TY53 TY54 TY55 Formula C58H72N4O8S4·3(C2H6O)·O C52H34N4 C68.30H84.45Cl0N4O6.03S4 Mr [gmol-1 ] 1253.63 714.83 1186.24 Crystal system Triclinic Triclinic Triclinic Space group 1P 1P 1P Unit cell dimensions a [Å] 13.9134 (18) 10.2435 (4) 14.6911 (6) b [Å] 15.4414 (19) 13.5620 (6) 14.8673 (6) c [Å] 17.086 (2) 13.7266 (6) 15.0303 (7) α [°] 103.616 (2)° 106.560 (3) 85.407 (3) β [°] 104.584 92.709 (3) 86.152 (3) γ [°] 96.412 (2)° 102.490 (3) 79.641 (3) V [Å3 ] 3395.2 (7) 1772.45 (13) 3214.3 (2) Z 2 2 2 T [K] 150 100 100 Dx(calc) [g cm-3 ] 1.209 1.339 1.226 Radiation type Mo K radiation Synchrotron Synchrotron Radiation λ [Å] 0.71073 0.7749 0.7749 Crystal size [mm3 ] 0.88 × 0.42 × 0.20 0.30 × 0.17 × 0.05 0.11 × 0.09 × 0.01 Extinction Coeff. 0.0043(6) 0.017 (4) N/A 2θmax [°] 48.0 48.8 29.2 Reflections measured 37604 10339 22331 Unique reflections 12465 3836 8081 reflections with I > 2σ(I) 8007 2960 5492 Transmission factors 0.844, 0.961 0.973, 0.995 0.973, 0998 Rint 0.062 0.041 N/A Number of parameters 818 510 947 Number of restraints 103 0 784 R[F2 > 2σ(F2 )] 0.068 0.056 0.092 wR(F2 ) 0.190 0.172 0.277
  • 23. This journal © The Royal of Dalton r Compound TY57 TY59 Formula C74H84N4O6S4·CHCl3·0.5(C2H3N) C144H160N8O12S8·C16H36N+ ·Cl·0.5(CHCL3)·C2H3N Mr [gmol-1 ] 1393.58 2829.91 Crystal system Monoclinic Triclinic Space group P21/c 1P Unit cell dimensions a [Å] 20.1570 (13) 15.1315 (5) b [Å] 26.8064 (17) 28.8618 (11) c [Å] 27.0262 (17) 35.9491 (13) α [°] 90 104.778 (2) β [°] 93.8414 (12) 94.065 γ [°] 90 92.353 (2) V [Å3 ] 14570.4 (16) 15113.8 (9) Z 8 4 T [K] 150 100 Dx(calc) [g cm-3 ] 1.271 1.244 Radiation type Mo K radiation Synchrotron Radiation λ [Å] 0.71073 0.7749 Crystal size [mm3 ] 0.26 × 0.24 × 0.08 0.16 × 0.13 × 0.04 Extinction Coeff. N/A N/A 2θmax [°] 46.2 50.6 Reflections measured 127772 116001 Unique reflections 28608 42128 reflections with I > 2σ(I) 17146 27707 Transmission factors 0.927, 0.977 0.956, 0.989 Rint 0.093 0.062 Number of parameters 1737 3705 Number of restraints 109 256 R[F2 > 2σ(F2 )] 0.092 0.068 wR(F2 ) 0.301 0.185
  • 24. This journal © The Royal of Dalton r Compound GW114 GW115 GWALS11 Formula C29.02H16.02F3N3 C34H30F6N8O2·CHCl3 C41H24F2N4 Mr [gmol-1 ] 463.77 816.03 610.64 Crystal system Orthorhombic Monoclinic Orthorhombic Space group P212121 P21/n P212121 Unit cell dimensions a [Å] 8.1301 (5) 13.8932 (16) 7.8011 (4) b [Å] 18.1470 (12) 14.1669 (16) 13.4885 (6) c [Å] 29.9929 (19) 18.231 (2) 27.7849 (13) α [°] 90 90 90 β [°] 90 99.7139 (16) 90 γ [°] 90 90 90 V [Å3 ] 4425.1 (5) 3536.8 (7) 2923.7 (2) Z 8 4 4 T [K] 150 150 100 Dx(calc) [g cm-3 ] 1.392 1.532 1.387 Radiation type Mo K radiation Mo K radiation Synchrotron Radiation λ [Å] 0.71073 0.71073 0.7749 Crystal size [mm3 ] 0.40 × 0.21 × 0.05 0.54 × 0.54 × 0.36 0.30 × 0.02 × 0.02 Extinction Coeff. N/A N/A 0.0077(9) 2θmax [°] 44.0 56.8 51.6 Reflections measured 30333 35615 20833 Unique reflections 9168 8798 4347 reflections with I > 2σ(I) 6291 5995 3904 Transmission factors 0.961, 0.995 0.839, 0.888 0.968, 0.998 Rint 0.057 0.051 0.053, Number of parameters 760 499 425 Number of restraints 546 2 0 R[F2 > 2σ(F2 )] 0.045 0.064 0.035 wR(F2 ) 0.089 0.209 0.083
  • 25. This journal © The Royal of Dalton r Compound GWALS13 GWALS14 GWALS15 Formula C34H42S4 C34H24F8N6O4 C60H42F30N6O12 Mr [gmol-1 ] 578.91 732.59 1608.99 Crystal system Triclinic Monoclinic Monoclinic Space group 1P P21/n P21/c Unit cell dimensions a [Å] 4.8575 (2) 12.7854 (5) 17.4161 (6) b [Å] 9.7907 (4) 8.9381 (4) 14.3221 (5) c [Å] 16.7849 (6) 14.1393 (5) 25.7502 (9) α [°] 101.411 90 90 β [°] 94.356 96.872 (2) 101.207 γ [°] 96.746 (2) 90 90 V [Å3 ] 773.02 (5) 1604.19 (11) 6300.5 (4) Z 1 2 4 T [K] 150 150 150 Dx(calc) [g cm-3 ] 1.244 1.517 1.696 Radiation type Synchrotron Synchrotron Synchrotron Radiation λ [Å] 0.7749 0.7749 0.7749 Crystal size [mm3 ] 0.25 × 0.07 × 0.03 0.35 × 0.18 × 0.02 0.05 × 0.03 × 0.01 Extinction Coeff. N/A N/A 0.00050(8) 2θmax [°] 80.6 67.2 51.2 Reflections measured 14583 20949 44141 Unique reflections 7360 4910 9083 reflections with I > 2σ(I) 6299 3877 6863 Transmission factors 0.904, 0.988 0.959, 0.998 0.992, 0.998 Rint 0.024 0.030 0.056 Number of parameters 173 239 997 Number of restraints 0 0 0 R[F2 > 2σ(F2 )] 0.036 0.043 0.055 wR(F2 ) 0.104 0.121 0
  • 26. This journal © The Royal of Dalton r Compound GWALS16 RSH29 GWALS12 Formula C38H50S4 C21H13HoN3O12·1.5(H2O) C12H6BrF4NO Mr [gmol-1 ] 635.02 691.30 336.09 Crystal system Triclinic Triclinic Monoclinic Space group 1P 1P P21/n Unit cell dimensions a [Å] 4.8430 (2) 11.6545 (3) 7.0989 (4) b [Å] 9.7919 (5) 13.5819 (3) 13.8009 (7) c [Å] 18.6253 (8) 15.1150 (4) 11.7747 (7) α [°] 81.716 (2) 90.78 90 β [°] 82.641 (2) 95.76 91.175 (4) γ [°] 82.641 (2) 103.15 90 V [Å3 ] 862.11 (7) 2316.43 (10) 1153.34 (11) Z 1 4 4 T [K] 150 150 150 Dx(calc) [g cm-3 ] 1.223 1.982 1.936 Radiation type Synchrotron Mo K radiation Synchrotron Radiation λ [Å] 0.7749 0.71073 0.6199 Crystal size [mm3 ] 0.30 × 0.12 × 0.01 0.20 × 0.18 × 0.06 0.30 × 0.10 × 0.06 Extinction Coeff. N/A N/A 0.0020(6) 2θmax [°] 80.6 61.2 50.4 Reflections measured 16243 37412 15540 Unique reflections 8185 14098 3109 reflections with I > 2σ(I) 6239 12120 2128 Transmission factors 0.895, 0.996 0.541, 0.818 0.518, 0.863 Rint 0.029 0.024 0.044 Number of parameters 191 779 344 Number of restraints 0 14 174 R[F2 > 2σ(F2 )] 0.053 0.024 0.031 wR(F2 ) 0.0147 0.057 0.079
  • 27. This journal © The Royal of Dalton r Compound GW123 Formula C32H26F8O2 Mr [gmol-1 ] 594.53 Crystal system Triclinic Space group 1P Unit cell dimensions a [Å] 11.674 (3) b [Å] 21.478 (5) c [Å] 22.335 (5) α [°] 88.008 (3) β [°] 85.529 (3)° γ [°] 86.292 (3)° V [Å3 ] 5569 (2) Z 8 T [K] 150 Dx(calc) [g cm-3 ] 1.418 Radiation type Mo K radiation Radiation λ [Å] 0.71073 Crystal size [mm3 ] 0.91 × 0.15 × 0.05 Extinction Coeff. 0.0023(2) 2θmax [°] 45.0 Reflections measured 462780 Unique reflections 14560 reflections with I > 2σ(I) 9130 Transmission factors 0.896, 0.994 Rint 0.078 Number of parameters 1565 Number of restraints 50 R[F2 > 2σ(F2 )] 0.050 wR(F2 ) 00.128 Appendix ii CIF Locations for Reference Structures Discussed Compound CIF location [Hyperlink; click to follow] R1 Not Available R2 Not Available R3 Not Available R4 https://summary.ccdc.cam.ac.uk/structure-summary-form 997001 R5 https://summary.ccdc.cam.ac.uk/structure-summary-form 841827 R6 https://summary.ccdc.cam.ac.uk/structure-summary-form 245644 R7 http://pubs.acs.org DOI: 10.1021/ol4002653 R8 http://pubs.acs.org DOI: 10.1021/ol401438a
  • 28. This journal © The Royal of Dalton r References 1. R, Muthukrishnan, J. Org. Chem., 1978, 43, 4905. 2. (a) S. Fleming, J. M. Harrowfield, M. I Ogden, B. W. Skelton, D. F. Stewart, and A. F. White, Dalton Trans., 2003, 17, 3319; (b) https://sites.google.com/site/cdgutsche/home/bibliography (accessed April 2015). 3. (a) C. D. Gutsche, Monographs in Supramolecular Chemistry: Calixarenes, Royal Society of Chemistry, London, 1989; (b) C. D. Gutsche, Monographs in Supramolecular Chemistry: Calixarenes Revisited, Royal Society of Chemistry, London, 1998; (c) C. D. Gutsche, Monographs in Supramolecular Chemistry: Calixarenes, an Introduction, Royal Society of Chemistry, London, 2008. 4. H. Kumagai, M. Hasegawa, S. Miyanari, Y. Sugawa, Y. Sato, T. Hori, S. Ueda, H. Kamiyama, and S. Miyano, Tet. Lett.,1997, 38, 3971. 5. C. Zhang, Z. Wang, S. Song, X. Meng, Y.-S. Zheng, X.-L. Yang, and H.-Bi. Xu, J. Org. Chem., 2014, 79, 2729. 6. C. Redshaw, M. A. Rowan, L. Warford, D. M. Homden, A. Arbaoui, M. R. J. Elsegood, S. H. Dale, T. Yamato, C. Pérez Casas, S. Matsui, and S. Matsuura, Chem. Eur. J., 2007, 13, 1090. 7. J.-L. Zhao, H. Tomiyasu, X.-L. Ni, X. Zeng, M. R. J. Elsegood, C. Redshaw, S. Rahman, P. E. Georghiou, and Takehiko Yamato, New J. Chem., 2014, 38, 6041. 8. A. Yilmaz, S. Memon, and M. Yilmaz, Tetrahedron, 2002, 58, 7735. 9. A. I. Vovk, L. A. Kononets, V. Y. Tanchuk, S. O. Cherenok, A. B. Drapailo, V. I. Kalchenko, and V. P. Kukhar, Bioorg. Med. Chem. Lett., 2010, 20, 483. 10. A. Pochettino, Atti. Accad. Rend., 1906, 15, 355. 11. D. M. Chapin, C. S. Fuller, G. L. Pearson, J. Appl. Phys., 1954, 25, 676. 12. C. W. Tang, Appl. Phys. Lett., 1986, 48, 183. 13. H. Spanggaard, F. C. Krebs, Sol. Energy Mat. & Sol. Cells, 2004, 83, 125. 14. http://www.nobelprize.org/nobel_prizes/physics/laureates/2014/press.html (accessed April 2015). 15. G. E. Bacon. Acta Crystallogr., 1951, 4, 558. 16. F. M. Winnik, Chem. Rev., 1993, 93, 587. 17. I. B. Berlman, Handbook of fluorescence spectra of aromatic molecules, Academic Press, New York 2nd ed., 1971. 18. A. Dino, G. E. Jabbour, and N. Peyghambarian, Advanced Mat., 2000, 12, 1249. 19. B. Park, C. H. Park, Y. Yim, and J. Park. J. Appl. Phys., 2010, 108, 84508. 20. SAINT and APEX 2 Software for CCD Diffractometers, Bruker AXS: Madison, WI, USA, 2008. 21. G. M. Sheldrick, SHELXTL User Manual, version 6.10, Bruker AXS: Madison, WI, USA, 2000. 22. G. M. Sheldrick, Acta Crystallogr., 2008, A64, 112. 23. A. L. Spek, Acta Crystallogr., 1990, Sect. A, 46, C34. 24. (a) J. N. Babu, V. Bhalla, M. Kumar, R. K. Mahajan, and R. K. Puri, Tet. lett., 2008, 49, 2772; (b) M. Kumar, R. K. Mahajan, K. V. Bhalla, N. Singh, H. N. Sharma, and I. Kaur, Tet. Lett., 2001, 42, 5315; (c) V. Bhalla, R. Kumar, M. Kumar, and A. Dhir, Tetrahedron, 2007, 63, 11153. 25. S. Rahman, H. Tomiyasu, H. Kawozie, J.-l. Zhao, H. Cong, X.-L. Ni, X. Zeng, M. R. J. Elsegood, T. G. Warwick, S. J. Teat, C. Redshaw, P. E. Georghiou, and T. Yamato. N. J. Chem., 2015. 26. O. Kundrat, V. Eigner, P. C. Rínova, J. Kroupa, and Pavel Lhotak, Tetrahedron., 2011, 67, 8367. 27. H. Gampp, M. Maeder, C. J. Meyer, and A. D. Zuberbulher, Talanta, 1985, 32, 95. 28. J.-L. Zhao, H. Tomiyasu, X.-L. Ni, X. Zeng, M. R. J. Elsegood, C. Redshaw, S. Rahman, P. E. Georghiou, and T. Yamato, New J. Chem., 2014, 38, 6041. 29. A. Casnati, P. Minari, A. Pochini and R. Ungaro, J. Chem. Soc. Chem. Commun., 1991, 19, 1413. 30. X.-L. Ni, S. Rahman, S. Wang, C.-C. Jin, X. Zeng, D. L. Hughes, C. Redshaw, and T. Yamato, Org. Biomol. Chem., 2012, 10, 4618. 31. J.-L. Zhao, H. Tomiyasu, X.-L. Ni, X. Zeng, M. R. J. Elsegood, C. Redshaw, S. Rahman, P. E. Georghiou, S. J. Teat, and T. Yamato, Org. Biomol. Chem., 2015, 13, 3476. 32. (a) F. Camerel, S. Diring, B. Donnio, T. Dintzer, S. Toffanin, R. Capelli, M. Muccini, and R. Ziessel, J. Am. Chem. Soc., 2009, 131, 18177; (b) Z. J.-H. Zhao, X. Chen, X. Wang, P. Lu, and Y. Yang, J. Org. Chem., 2009, 74, 383; (c) K. R. J. Thomas, M. Velusamy, J. T. Lin, C. H. Chuen, and Y.-T. Tao, J. Mater. Chem., 2005, 15, 4453. 33. X. Feng, J.-Y. Hu, F. Iwanaga, N. Seto, C. Redshaw, M. R. J. Elsegood, and T. Yamato, Org. Lett., 2013, 15(6), 1318. 34. X. Feng, F. Iwanaga, J.-Y. Hu, H. Tomiyasu, M. Nakano, C. Redshaw, M. R. J. Elsegood, and Takehiko Yamato Org. Lett., 2013, 15(14), 3594.