ArticlePDF Available

Abstract and Figures

The synergy of aromatic gain and hydrogen bonding in a supramolecular polymer is explored. Partially aromatic bis(squaramide) bolaamphiphiles were designed to self-assemble through a combination of hydrophobic, hydrogen-bonding, and aromatic effects into stiff, high-aspect-ratio fibers. UV and IR spectroscopy show electron delocalization and geometric changes within the squaramide ring indicative of strong hydrogen bonding and aromatic gain of the monomer units. The aromatic contribution to the interaction energy was further supported computationally by nucleus-independent chemical shift (NICS) and harmonic oscillator model of aromaticity (HOMA) indices, demonstrating greater aromatic character upon polymerization: at least 30 % in a pentamer. The aromatic gain-hydrogen bonding synergy results in a significant increase in thermodynamic stability and a striking difference in aggregate morphology of the bis(squaramide) bolamphiphile compared to isosteres that cannot engage in this effect. © 2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Content may be subject to copyright.
German Edition:DOI:10.1002/ange.201503905
Non-Covalent Interactions International Edition:DOI:10.1002/anie.201503905
Aromatic Gain in aSupramolecular Polymer**
Victorio Saez Talens,Pablo Englebienne,Thuat T. Trinh, Willem E. M. Noteborn, Ilja K. Voets,
and Roxanne E. Kieltyka*
Abstract: The synergy of aromatic gain and hydrogen bonding
in asupramolecular polymer is explored. Partially aromatic
bis(squaramide) bolaamphiphiles were designed to self-assem-
ble through acombination of hydrophobic,hydrogen-bond-
ing, and aromatic effects into stiff,high-aspect-ratio fibers.UV
and IR spectroscopyshowelectron delocalization and geo-
metric changes within the squaramide ring indicative of strong
hydrogen bonding and aromatic gain of the monomer units.
The aromatic contribution to the interaction energy was further
supported computationally by nucleus-independent chemical
shift (NICS) and harmonic oscillator model of aromaticity
(HOMA) indices,demonstrating greater aromatic character
upon polymerization:atleast 30 %inapentamer.The
aromatic gain–hydrogen bonding synergy results in asignifi-
cant increase in thermodynamic stability and astriking differ-
ence in aggregate morphology of the bis(squaramide) bolam-
phiphile compared to isosteres that cannot engage in this effect.
Aromatic gain is considered to be athermodynamic driving
force in several organic reactions.Inaromatic substitutions,
Bergman cyclizations, aromatic Cope rearrangements,[1,5]H
sigmatropic shifts,and [4++2] cyclizations,among others,the
restoration of aromaticity helps to explain their exergonic
character and increased efficiency.[1] In supramolecular poly-
mers,[2] where monomers are held together by non-covalent
interactions resulting in higher-order aggregates with various
topologies,the concept of aromatic gain in the construction of
such systems is unexplored.
Aromaticity has captivated chemists since its introduction
as aconcept 150 years ago by Kekul¦.[3] In contrast to other
chemical concepts,such as chemical bonding,electronegativ-
ity and acidity/basicity,aromaticity is not adirect physical
observable and its exact definition is the subject of much
debate.[4] Classically,cyclic p-conjugated compounds are
aromatic when they show differences in geometric, energetic,
and magnetic criteria relative to their acyclic analogues.[5] In
recent years,computational methods such as nucleus-inde-
pendent chemical shift (NICS),[5c,6] harmonic oscillator model
of aromaticity (HOMA),[7] and aromatic stabilization ener-
gies (ASE),[5b] have grown in use to describe aromaticity.In
the case of NICS,excellent correlation has been reported with
experimental nuclear magnetic resonance data as well as
other descriptors of aromaticity,[4a] thus opening the door to
predict the aromaticity of new compounds and structures
apriori. Ve ry recently,NICS calculations demonstrated
reciprocal hydrogen-bonding–aromaticity relationships that
can have important consequences on the strength of hydro-
gen-bonding interactions.[8]
Squaramides,[9] which are composed of two NH hydrogen-
bond donors opposite two carbonyl hydrogen-bond acceptors
on aconformationally rigid cyclobutene ring, are predicted to
show partial aromatic character.[9] This character arises from
the delocalization of the nitrogen lone pair into the cyclo-
butenedione ring system (HîckelÏs rule:(4n+2) pelectrons,
n=0).[10] In the solid state,catemers of disecondary squar-
amides arranged in ahead-to-tail motif have been reported[11]
and they may benefit from strong resonance-assisted hydro-
gen bonding (RAHB) interactions similar to squaric acids.[12]
Applications of the squaramide unit have been found in
medicinal chemistry,catalysis,and anion recognition.[13] The
capacity of squaramides to form strong hydrogen bonds that
simultaneously influence their aromatic character is highly
appealing to guide the formation of increasingly stable
supramolecular polymers.Herein, we incorporate the squar-
amide synthon into abolaamphiphilic construct that self-
assembles into stiff fibers in water, and we explore the
coupling of hydrogen-bonding and aromatic gain using
experiment and computation.
Compound 1consists of two oligo(ethylene glycol) methyl
ether chains opposite acentral hydrophobic core with two
embedded squaramide units (Figure 1). 1HNMR spectra of
1in D2Owere suggestive of strong aggregation that is
resistant to thermal denaturation up to 6588C. Only 1HNMR
spectra recorded in CDCl3or [D2]HFIP were well-resolved
and suggestive of various degrees of depolymerization.
Theeffect of the squaramide synthon on the self-assembly
of 1in water was evaluated by cryo-transmission electron
[*] V. Saez Talens, W. E. M. Noteborn, Dr.R.E.Kieltyka
Department of Supramolecular and Biomaterials Chemistry
Leiden Institute of Chemistry,Leiden University
P.O. Box 9502, 2300 RA Leiden (The Netherlands)
E-mail:r.e.kieltyka@chem.leidenuniv.nl
Dr.P.Englebienne
Process &Energy Laboratory
Delft University of Technology
Leeghwaterstraat 39, 2628 CB Delft (The Netherlands)
Dr.T.T.Trinh
Department of Chemistry
NorwegianUniversity of Science and Te chnology
7491 Trondheim (Norway)
Dr.I.K.Voets
Department of Chemical Engineering and Chemistryand
Institute for Complex Molecular Systems
Eindhoven University of Technology
P.O. Box 513, 5600 MB Eindhoven (The Netherlands)
[**] We thank R.I. Koning (TEM), B. Koster (TEM), F. Galli (AFM), M.
Rabe (IR), A. J. M. Sweere (CULGI), R. Matadeen (TEM, NeCEN), K.
Pieterse (ICMS Animation Studio), and A. Kros for essential
discussions. This work is funded by aVENI grant (to R.E.K.) from
NWO.
Supporting information for this article is available on the WWW
under http://dx.doi.org/10.1002/anie.201503905.
..
Angewandte
Communications
10502 Ó2015 Wiley-VCH Ve rlag GmbH &Co. KGaA, We inheim Angew.Chem. Int. Ed. 2015,54,10502 –10506
microscopy (cryo-TEM) and atomic force microscopy
(AFM). Cryo-TEM images of 1(1 wt%) displayed stiff,
micrometer-long fibrils with auniform diameter. Short, rod-
like structures on the order of 12.6 2.4 nm in length were
found upon sonication (Figure 2a)and slowly progressed into
micrometer-long fibers.Fibers of 1were 6.4 1.2 nm in
diameter, on par with the length of the hydrophobic region of
the bolaamphiphile (Figure 2b). By small-angle X-ray scat-
tering measurements (SAXS,Figure 2c), across-sectional
radius (rcs)of3.5 nm and across-sectional mass per unit
length (ML)of2.5 ×10
20–6.0 ×10
20 gnm¢1was determined for
fibers of 1,indicating that approximately 10–30 squaramide
bolaamphiphiles per nm can be found along the fiber axis.
These results suggest that hydrogen bonds parallel to the fiber
axis drive the formation of highly anisotropic fibers,mean-
while the combination of hydrophobic and p-interactions
between squaramide moieties facilitate the assembly of
several bolaamphiphiles in the lateral direction (Figure 1).
To better understand the consequence of self-assembly on
the squaramide synthon, spectroscopy at the molecular level
was pursued. UV spectroscopy of 1in water showed maxima
at 255 and 329 nm, and ashoulder around 310 nm (Figure 3a).
Disruption of the polymerized state was achieved using both
temperature and various solvents.More specifically,hexa-
fluoroisopropanol (HFIP), apotent hydrogen bond disruptor,
promoted depolymerization resulting in the gradual loss of
the red-shifted hydrogen-bonded squaramide N¢Hproton-
donor pp*bands (329 nm) and the blue-shifted C=Oproton-
acceptor n–p*bands (255 nm), concomitant with the growth
of the non-hydrogen bonded monomer band (310 nm).[14]
These experimental trends are in agreement with TD-DFT
calculations,where two superimposed absorption bands of
similar intensity corresponding to the HOMO–LUMO and
HOMO–(LUMO +1) transitions are predicted for the mo-
nomer;inoligomers,the high wavelength band is progres-
sively red-shifted while the other appears blue-shifted. Self-
assembly of 1through strong hydrogen bonding interactions
results in increased orbital overlap between squaramide units
and further electron delocalization within the individual
squaramide rings,enabling aromatic gain to occur.
Geometric changes to the squaramides upon self-assem-
bly were examined by IR spectroscopy.Solutions of
1(2 wt%) in D2Owere measured at room temperature.
Above the amide Iregion, asmall broad band at 1796 cm¢1,
consistent with squaramide ring breathing, was found exper-
imentally and confirmed by modeling (Figure 3b). In the
amide Iregion, asymmetric and symmetric C=Ostretches
(1687, 1676, and 1642 cm¢1)ofthe squaramide and carbamate
moieties were recorded. Strong hydrogen bonding of the
squaramide units was observed through the N¢Hstretch at
3162 cm¢1(inset in Figure 3b). In [D2]HFIP,the blue-shifting
of several bands such as the ring breathing (13 cm¢1)and
symmetric C=Ostretch (14 cm¢1)modes were observed and
suggestive of depolymerization. Owing to lack of transpar-
encyof[D
2]HFIP in the N¢Hregion, an approximation for
free N¢Hstretch (3452 cm¢1)was made for 1in CDCl3.The
experimental data correlated well with ab initio calculations.
These results revealed that bond lengths in the squaramide
are systematically altered as afunction of oligomer length
(Figure 4a): double bonds become longer, whereas single
bonds shorten, resulting in aring with less bond length
alternation. With these bond lengths,wecomputed HOMA
values of ¢0.015 and 0.516 for the isolated monomer and the
central monomer in apentamer,respectively,while avalue of
Figure 1. a) Structure of the squaramide-based bolaamphiphile 1.b)Self-assembly of 1into fibrillar structures, and depolymerization by
hexafluoroisopropanol(HFIP). Within the fibrillar structure, hydrogen bonds are proposed to occur parallel to the fiber axis while p-interactions
between squaramidebolaamphphiles occur in the lateral direction, as depicted. c) Proposed hydrogen-bonding interactions between squaramide
monomers.
Angewandte
Chemie
10503Angew.Chem. Int.Ed. 2015,54,10502 –10506 Ó2015 Wiley-VCH Ve rlag GmbH &Co. KGaA, Weinheim www.angewandte.org
one is defined for aromatic compounds.Experiments point to
strong hydrogen bonding and computed geometric consid-
erations demonstrate an increase in aromatic character within
the squaramide unit due to supramolecular polymerization.
NICS-scan profiles,[15] ameasurement of the magnetic
shielding above and at the center of the ring, were computed
on an axis passing though the center of the squaramide ring
for monomers to pentamers to quantify the aromatic charac-
ter upon oligomerization. Theprofiles for the individual
squaramide units were negative overall and exhibited amini-
mum around 0.6 è, consistent with an aromatic ring. Upon
increasing oligomer length, the NICS values became more
negative without achange in the shape of the curves,
suggestive of increased aromaticity.Inparticular,the change
in NICS at 0.6 èfrom the ring plane (Figure 4b)when going
from amonomer (¢6.8) to the central monomer of apentamer
(¢8.4), is in line with previous reports.[16] Additionally,the
aromatic stabilization energy accounts for at least 30 %ofthe
total interaction energy in asquaramide pentamer
(¢85.6 kJmol¢1out of ¢271.7 kJmol¢1including BSSE cor-
rection) using aheterodimer of vinylogous amides that cannot
exhibit aromaticity as areference.
We further investigated the thermodynamic consequence
of aromatic gain experimentally by measuring the critical
aggregation concentration (CAC)using static light scattering
(SLS). An order of magnitude lower CAC, corresponding to
afree energy difference DDGagg =¢5.25 kJmol¢1,was
obtained for molecule 1(7.94 ×10
¢6m)incomparison to
urea-based analogue 5(7.41 ×10
¢5m;see the Supporting
Information) (Figure 5). These results are further supported
Figure 2. Cryo-TEM images of 1in aqueoussolution (1 wt %) after
sonication: a) t=0, and b) t=2weeks. Inset :Histogramsoflength (a)
and width (b). Scale bar :a)50nm ;b)100 nm. c) Small-angleX-ray
scattering profiles of squaramidefibers collected at aconcentration of
4and 5mgmL¢1.
Figure 3. a) UV/Vis spectrum of 1in water (0.005 wt %) as afunction
of HFIP concentration. b) IR spectrum recorded in the amide Iregion
and amide II in D2Oand [D2]HFIP. Inset:N
¢Hand C¢Hstretch
region.
..
Angewandte
Communications
10504 www.angewandte.org Ó2015 Wiley-VCH VerlagGmbH &Co. KGaA,Weinheim Angew.Chem. Int. Ed. 2015,54,10502 –10506
by DFT calculations,where the interaction energy computed
per hydrogen bond of urea oligomers was found to be smaller
(¢23.9 vs. ¢35.6 kJmol¢1for pentamers) and does not
increase as steeply with oligomer length (+18%vs.
+30%). Intriguingly,astriking difference in the fiber
morphology was found above the CACofboth molecules.
Whereas 1consistently formed long and stiff micron-sized
fibers,short worm-like or spherical aggregates were obtained
for 5.Given the similarity of the hydrophilic and hydrophobic
blocks,these results suggest that the coupling of aromatic gain
and hydrogen-bonding in addition to the structural rigidity of
the squaramide units act collectively to lower the critical
aggregation concentration, and propagate the formation of
high-aspect-ratio fibers in water.
We find that the capacity of squaramides to couple
hydrogen bonding and aromaticity facilitates the formation of
robust supramolecular polymers.The gain in aromatic
character upon assembly is demonstrated through bond
length equalization, decreased NICS values,high aromatic
stabilization (ASE) values,and increased thermodynamic
stability of the resultant aggregates.These changes are in
accordance with the geometric,magnetic,and energetic
criteria used to describe aromaticity.Moreover,the aromatic
gain is asignificant component of the total interaction energy
of squaramide-based supramolecular polymers,explaining
the observed increase in thermodynamic stability relative to
the monomers and to their urea counterparts.Insummary,
this self-tuning behavior between hydrogen bonding and
aromaticity within the squaramide ring system cannot be
achieved by other simple ditopic synthons,such as ureas or
amides,commonly used to construct supramolecular poly-
mers.Therefore,weanticipate that the information gained
here can enrich the palette of hydrogen-bonding monomers
used for supramolecular polymer assembly by implementing
aromaticity as adesign consideration.
Keywords: aromaticity ·non-covalent interactions ·
self-assembly ·squaramides ·supramolecular polymers
Howtocite: Angew.Chem. Int. Ed. 2015,54,10502–10506
Angew.Chem. 2015,127,10648 –10652
[1] a) I. V. Alabugin, M. Manoharan, B. Breiner,F.D.Lewis, J. Am.
Chem. Soc. 2003,125,9329 –9342 ;b)C.F.Bernasconi, M. L.
Ragains,S.Bhattacharya, J. Am. Chem. Soc. 2003,125,12328
12336;c)T.Bekele,M.H.Shah, J. Wolfer,C.J.Abraham,A.
Weatherwax, T. Lectka, J. Am. Chem. Soc. 2006,128,1810
1811;d)D.J.Babinski, X. G. Bao,M.ElArba, B. Chen, D. A.
Hrovat,W.T.Borden, D. E. Frantz, J. Am. Chem. Soc. 2012,134,
16139 –16142 ;e)M.R.Manaa, D. W. Sprehn, H. A. Ichord, J.
Am. Chem. Soc. 2002,124,13990 –13991;f)M.Manoharan, F.
de Proft, P. Geerlings, J. Org.Chem. 2000,65,6132 –6137.
[2] a) T. F. A. de Greef,M.M.J.Smulders,M.Wolffs,A.P.H.J.
Schenning, R. P. Sijbesma,E.W.Meijer, Chem. Rev. 2009,109,
5687 –5754 ;b)T.Aida, E. W. Meijer,S.I.Stupp, Science 2012,
335,813 –817;c)N.Chebotareva, P. H. H. Bomans,P.M.
Frederik, N. A. J. M. Sommerdijk, R. P. Sijbesma, Chem.
Commun. 2005,4967 –4969;d)C.M.A.Leenders,L.Alber-
tazzi, T. Mes,M.M.E.Koenigs,A.R.A.Palmans, E. W. Meijer,
Chem. Commun. 2013,49,1963 –1965;e)G.Borzsonyi, R. L.
Beingessner,T.Yamazaki, J. Y. Cho,A.J.Myles,M.Malac,R.
Egerton, M. Kawasaki,K.Ishizuka, A. Kovalenko, H. Fenniri, J.
Am. Chem. Soc. 2010,132,15136 –15139;f)J.D.Hartgerink, E.
Beniash, S. I. Stupp, Science 2001,294,1684 –1688;g)E.Obert,
M. Bellot, L. Bouteiller,F.Andrioletti, C. Lehen-Ferrenbach, F.
Boue, J. Am. Chem. Soc. 2007,129,15601 –15605;h)B.
Rybtchinski, ACSNano 2011,5,6791 –6818;i)D.Gçrl, X.
Zhang,F.Wîrthner, Angew.Chem. Int. Ed. 2012,51,6328
6348; Angew.Chem. 2012,124,6434 –6455;j)J.van Esch, S.
de Feyter,R.M.Kellogg,F.deSchryver,B.L.Feringa, Chem.
Eur.J.1997,3,1238 –1243;k)J.van Esch, R. M. Kellogg,B.L.
Feringa, Tetrahedron Lett. 1997,38,281 –284.
Figure 4. a) Geometric changes (computed at the M06-2X/6-
311++G(d,p) level of theory) in N-methyl squaramide between an
isolated monomer (normal text) and the central monomer in apen-
tamer (in bold italics). b) NICS values at apoint 0.6 çfrom the ring
plane for an isolated monomer and each monomer in oligomers of
length 2–5 (GIAO-M06-2X/6-311++G(d,p)).
Figure 5. Scattered intensity (1000 s¢1)asafunction of concentration,
log[C], of 1and 5using static light scattering (SLS) experiments to
determine the critical aggregation concentration (CAC). Representative
cryo-TEM images:the image on the left belongs to 1and the image
on the right to 5,showing fibrils in the case of 1and worm-likeor
spherical aggregates in the case of 5.Scale bar:50nm.
Angewandte
Chemie
10505Angew.Chem. Int.Ed. 2015,54,10502 –10506 Ó2015 Wiley-VCH Ve rlag GmbH &Co. KGaA, Weinheim www.angewandte.org
[3] A. Kekul¦, Bull. Soc.Chem. Fr. 1865,3,98.
[4] a) M. K. Cyran
´ski, T. M. Krygowski, A. R. Katritzky,P.v.R.
Schleyer, J. Org.Chem. 2002,67,1333 –1338;b)K.K.Baldridge,
J. S. Siegel, J. Phys.Org.Chem. 2004,17,740 –742;c)R.M.
Gomila,D.Quinonero,C.Rotger,C.Garau, A. Frontera, P.
Ballester,A.Costa, P. M. Deya, Org.Lett. 2002,4,399 –401;
d) A. R. Katritzky,M.Karelson, S. Sild, T. M. Krygowski, K. Jug,
J. Org.Chem. 1998,63,5228 –5231.
[5] a) T. M. Krygowski,H.Szatylowicz, O. A. Stasyuk, J. Domini-
kowska,M.Palusiak, Chem. Rev. 2014,114,6383 –6422;
b) M. K. Cyran
´ski, Chem. Rev. 2005,105,3773 –3811;c)Z.F.
Chen, C. S. Wannere,C.Corminboeuf,R.Puchta, P. v. R.
Schleyer, Chem. Rev. 2005,105,3842 –3888.
[6] P. v. R. Schleyer,C.Maerker,A.Dransfeld, H. J. Jiao,N.J.R.V.
Hommes, J. Am. Chem. Soc. 1996,118,6317 –6318.
[7] T. M. Krygowski, J. Chem. Inf.Comput. Sci. 1993,33,70–78.
[8] J. I. Wu,J.E.Jackson, P. v. R. Schleyer, J. Am. Chem. Soc. 2014,
136,13526 –13529.
[9] a) F. R. Wurm, H. A. Klok, Chem. Soc.Rev. 2013,42,8220
8236;b)R.I. Storer,C.Aciro,L.H.Jones, Chem. Soc.Rev. 2011,
40,2330 –2346.
[10] M. C. Rotger,M.N.Pina, A. Frontera, G. Martorell, P. Ballester,
P. M. Deya, A. Costa, J. Org. Chem. 2004,69,2302 –2308.
[11] R. Prohens,A.Portell, C. Puigjaner, R. Barbas,X.Alcobe,M.
Font-Bardia, S. Tomas, CrystEngComm 2012,14,5745 –5748.
[12] G. Gilli, V. Bertolasi, P. Gilli, V. Ferretti, Acta Crystallogr.Sect. B
2001,57,859 –865.
[13] a) G. Ambrosi, M. Formica, V. Fusi, L. Giorgi, A. Guerri, M.
Micheloni,P.Paoli, R. Pontellini, P. Rossi, Chem. Eur.J.2007,13,
702 –712 ;b)N.Busschaert, I. L. Kirby,S.Young,S.J.Coles,P.N.
Horton, M. E. Light, P. A. Gale, Angew.Chem. Int. Ed. 2012,51,
4426 –4430 ; Angew.Chem. 2012,124,4502 –4506;c)A.Ros-
tami, G. Guerin, M. S. Ta ylor, Macromolecules 2013,46,6439
6450.
[14] L. Sobczyk, S. J. Grabowski, T. M. Krygowski, Chem. Rev. 2005,
105,3513 –3560.
[15] A. Stanger, J. Org.Chem. 2006,71,883 –893.
[16] D. QuiÇonero,R.Prohens,C.Garau, A. Frontera, P. Ballester,
A. Costa, P. M. Dey, Chem. Phys.Lett. 2002,351,115 –120.
Received:April 28, 2015
Published online: July 14, 2015
..
Angewandte
Communications
10506 www.angewandte.org Ó2015 Wiley-VCH VerlagGmbH &Co. KGaA,Weinheim Angew.Chem. Int. Ed. 2015,54,10502 –10506
... [8][9][10][11][12][13][14][15][16] Looking at materials utilizing these intermolecular forces one can divide them mainly by motifs based on aromatic motifs or ones employing directional hydrogen bonds. [17][18][19][20][21][22][23][24][25] In the last decade, the benzene trisamide (BTA) motif has become a focus point for multiple investigations. [26][27][28][29][30][31][32][33][34] It comprises a benzene core functionalized with three amide groups which allow individual modification with further pendant groups. ...
Article
Full-text available
Hydrogen bonds are a versatile tool for creating fibrous, bottlebrush‐like assemblies of polymeric building blocks. However, a delicate balance of forces exists between the steric repulsion of the polymer chains and these directed supramolecular forces. In this work we have systematically investigated the influence of structural parameters of the attached polymers on the assembly behaviour of benzene trisurea (BTU) and benzene tris(phenylalanine) (BTP) conjugates in water. Polymers with increasing main chain lengths and different side chain sizes were prepared by reversible addition‐fragmentation chain‐transfer (RAFT) polymerization of hydroxyethyl acrylate (HEA), tri(ethylene glycol) methyl ether acrylate (TEGA) and oligo(ethylene glycol) methyl ether acrylate (OEGA). The resulting structures were analyzed using small angle X‐ray scattering (SAXS) and transmission electron microscopy (TEM). Both BTU and BTP formed fibres with PHEA attached, but a transition to spherical morphologies was observed at degrees of polymerisation (DP) of 70 and above. Overall, the main chain length appeared to be a dominating factor in inducing morphology transitions. Increasing the side chain size generally had a similar effect but mainly impeded any aggregation as is the case of POEGA. Interestingly, BTP conjugates still formed fibres, suggesting that the stronger intermolecular interactions can compensate partially for the steric repulsion.
... [40][41][42][43] Squaramides are particularly amenable to this purpose because of their strong hydrogen-bond-donating ability, planar structure, and the observed increase in aromaticity upon guest binding. 44,45 Moreover, their potential uses across a range of applications in the chemical sciences have led to their extensive utilization in recent years. 46 Herein, we describe a new class of squaramides that take advantage of both NH and CH bonding to recognize anions and, with the benefit of an incorporated indoline moiety, display bright green fluorescence. ...
... 241 Furthermore, the ability of squaramides to form strong hydrogen bonds that simultaneously improve the aromaticity of the four-membered ring is extremely beneficial, as the molecular recognition and self-assembly processes can take advantage of the thermodynamic stability caused by the aromatic gain. 244,245 Squaramides are predominantly known to be favourable in asymmetric synthesis as chiral ligands and H-binding catalysts, [246][247][248][249] but are also very useful as building blocks in the field of medicinal chemistry. ...
... Squaramides have attracted a significant amount of research attention in recent years due to several physical and chemical characteristics that render them an ideal scaffold on which to build charge neutral anion receptors [1,2]. Their ability to partake in bi-directional H-bonding that increases the aromatic character of the cyclobutenedione ring [3,4] has been seen to be of particular benefit for the recognition of halides [5][6][7][8]. Indeed, Amendola et al. showed that a simple squaramide-based receptor forms 1:1 complexes with halide ions that are up to 2 orders of magnitude more stable than the corresponding urea-based receptor [9]. ...
Article
Full-text available
The syntheses of two squaramide–naphthalimide conjugates (SN1 and SN2) are reported; the structures of SN1 and SN2 differ by the attachment of a squaramide—either at the ‘head’ or the ‘tail’ of the naphthalimide fluorophore. Both compounds displayed weak fluorescence due to the inclusion of a nitro-aromatic squaramide which efficiently quenches the emission of the naphthalimide. Both compounds were also shown to undergo self-aggregation as studied by 1H NMR and scanning electron microscopy (SEM). Furthermore, SN1 and SN2 gave rise to stark colourimetric changes in response to basic anions such as AcO−, SO42− HPO42−, and F−. The observed colour changes are thought to be due to deprotonation of a squaramide NH. The same basic anions also result in a further quenching of the naphthalimide emission. No colour change or emission modulations were observed in the presence of Cl−; however, 1H NMR studies suggest that moderate H-bonding occurs between this anion and both SN1 and SN2.
Article
The synthesis of three squaramide-containing bis-naphthalimide conjugates (1-3) are reported where the structures differ by the length of the carbon chain between the squaramide core and the 1,8-naphthalimide fluorophores. All compounds displayed weak fluorescence properties and self-aggregation as studied by UV/Vis, ¹H NMR spectroscopy and Scanning Electron Microscopy (SEM), however it was shown that aggregation can be disrupted by increased temperature. Anion binding experiments displayed colourimetric and spectroscopic modulations in response to a variety of anions where more basic anions resulted in receptor deprotonation. However, selectivity was observed in the case of H-bonding to OAc⁻ with binding measured by ¹H NMR titration experiments. No fluorometric changes were observed for any of the anions tested and we propose that the squaramide is capable of disrupting eximer formation resulting in a reduction of emission from these systems.
Article
Full-text available
Hydrogen‐bonded squaramide (SQ) supramolecular polymers exhibit uncommon thermoreversible polymorph transitions between particle‐ and fiber‐like nanostructures. SQs 1–3, with different steric bulk, self‐assemble in solution into particles (AggI) upon cooling to 298 K, and SQs 1 and 2, with only one dendronic group, show a reversible transformation into fibers (AggII) by further decreasing the temperature to 288 K. Nano‐DSC and UV/Vis studies on SQ 1 reveal a concentration‐dependent transition temperature and ΔH for the AggI‐to‐AggII conversion, while the kinetic studies on SQ 2 indicate the on‐pathway nature of the polymorph transition. Spectroscopic and theoretical studies reveal that these transitions are triggered by the molecular reorganization of the SQ units changing from slipped to head‐to‐tail hydrogen bonding patterns. This work unveils the thermodynamic and kinetic aspects of reversible polymorph transitions that are of interest to develop stimuli‐responsive systems.
Article
Hydrogen‐bonded squaramide (SQ) supramolecular polymers exhibit uncommon thermoreversible polymorph transitions between particle‐ and fiber‐like nanostructures. SQs 1‐3 , with different steric bulk, self‐assemble in solution into particles ( AggI ) upon cooling to 298 K, and SQs 1 and 2 , with only one dendronic group, show a reversible transformation into fibers ( AggII ) by further decreasing the temperature to 288 K. Nano‐DSC and UV/Vis studies on SQ 1 reveal a concentration‐dependent transition temperature and Δ H for the AggI ‐to‐ AggII conversion, while the kinetic studies on SQ 2 indicate the on‐pathway nature of the polymorph transition. Spectroscopic and theoretical studies reveal that these transitions are triggered by the molecular reorganization of the SQs units changing from slipped to head‐to‐tail hydrogen bonding patterns. This work unveils the thermodynamic and kinetic aspects of reversible polymorph transitions that are of interest to develop stimuli‐responsive systems.
Article
Carrier-free delivery of therapeutically relevant small molecules and functional oligonucleotides is extremely challenging and is one of the major hurdles in cancer therapy. Herein, we report a non-covalent approach for...
Article
Full-text available
Supramolecular materials provide unique opportunities to mimic both the structure and mechanics of the biopolymer networks that compose the extracellular matrix. However, strategies to modify their filamentous structures in space and time in 3D cell culture to study cell behavior as encountered in development and disease are lacking. We herein disclose a multicomponent squaramide-based supramolecular material whose mechanics and bioactivity can be controlled by light through co-assembly of a 1,2-dithiolane (DT) monomer that forms disulfide cross-links. Remarkably, increases in storage modulus from ∼200 Pa to >10 kPa after stepwise photo-cross-linking can be realized without an initiator while retaining colorlessness and clarity. Moreover, viscoelasticity and plasticity of the supramolecular networks decrease upon photo-irradiation, reducing cellular protrusion formation and motility when performed at the onset of cell culture. When applied during 3D cell culture, force-mediated manipulation is impeded and cells move primarily along earlier formed channels in the materials. Additionally, we show photopatterning of peptide cues in 3D using either a photomask or direct laser writing. We demonstrate that these squaramide-based filamentous materials can be applied to the development of synthetic and biomimetic 3D in vitro cell and disease models, where their secondary cross-linking enables mechanical heterogeneity and shaping at multiple length scales.
Article
Full-text available
The formation of supramolecular polymers in water through rational design of a benzene-1,3,5-tricarboxamide (BTA) motif is presented. Intermolecular hydrogen bonding and hydrophobic effects cooperate in the self-assembly into long fibrillar aggregates. Minimal changes in molecular structure significantly affect the internal packing of the aggregates.
Article
Computed association energies and dissected nucleus-independent chemical shifts (NICS) document the mutual enhancement (or reduction) of intermolecular interactions and the aromaticity of H-bonded substrates. H-bonding interactions that in-crease cyclic 4n+2 π-electron delocalization boost aromaticity. Conversely, such interactions are weak-ened when aromaticity is decreased due to more lo-calized quinoidal π-character. Representative exam-ples of the tautomeric equilibria of π-conjugated het-erocyclic compounds in protic solvents and other H-bonding environments also illustrate such H-bonding/aromaticity interplay.
Article
Perylenbisimide zählen zu den bedeutendsten funktionellen Farbstoffen mit vielfältigen Anwendungsmöglichkeiten. Aufgrund ihrer chemischen Robustheit, Photostabilität und herausragenden optischen und elektronischen Eigenschaften wurden sie als Pigmentfarbstoffe, Fluoreszenzsonden sowie n‐Halbleiter in der organischen Elektronik und Photovoltaik eingesetzt. Das ausgedehnte quadrupolare π‐System dieser Farbstoffklasse ermöglichte außerdem die Herstellung zahlreicher supramolekularer Architekturen mit faszinierenden photophysikalischen Eigenschaften. Jedoch blieb der supramolekulare Ansatz zur Bildung von Perylenbisimid‐Molekülverbänden größtenteils organischen Medien vorbehalten. Erfreulicherweise gab es in den letzten Jahren beachtliche Bemühungen zur Entwicklung von wasserlöslichen Perylenbisimiden und deren Anwendungen in wässrigen Medien. Dieser Aufsatz gibt einen aktuellen Überblick über die Selbstorganisation von Perylenbisimiden durch π‐π‐Wechselwirkung im wässrigen Medium. Synthetische Strategien zur Herstellung von wasserlöslichen Perylenbisimiden und der Einfluss des Wassers auf die π‐π‐Stapelung von Perylenbisimiden, sowie daraus resultierende Anwendungen werden eingehend diskutiert.
Article
Molecular geometry is of fundamental importance for any profound interpretation of molecular properties. This was elegantly pointed out by Roald Hoffmann and triggered intensive research in this direction. Experimental and theoretical methods enabling the determination of molecular geometry, developed since the mid-20th century, have provided numerical confirmation of the above-mentioned statement. Geometry-based aspects of aromaticity are the main subject of this Review. The theoretical method and the basis set used for estimation of HOMA constants should be the same as that used for geometry optimization of a given system. Their analysis was carried out for 25 systems and various advanced quantum-chemical models. The reference systems for the double CC, CN, and CO bond lengths are ethene, methylimine, and formaldehyde, respectively.
Article
Poly(squaramides) are a novel class of anion-responsive macromolecules that incorporate the diaminocyclobutenedione hydrogen bond donor group into the polymer backbone. Herein, the synthesis and properties of a series of fluorene-based poly(squaramides) varying in conformational rigidity, squaramide content, and propensity for aggregation are described. Structure–activity relationships for the anion sensory behavior of these polymers (as probed by fluorescence titrations, dynamic light scattering, confocal fluorescence microscopy, and transmission electron microscopy) indicate that anion-induced polymer aggregation leads to a cooperative response with enhanced levels of sensitivity and selectivity. These observations are consistent with a mechanism involving noncovalent cross-linking of polymer chains through squaramide–anion hydrogen-bonding interactions and point toward new applications of polyamides as stimulus-responsive materials.
Article
Two different types of monodimensional crystalline aggregates formed by squaramides, chains and ribbons, have been studied both in the solid state and ab initio with four model compounds. Although ribbons are a priori possible, only covalently forced syn conformations have been observed. Cooperative induction is postulated to explain the preference for chains instead of ribbons in squaramide crystals.
Article
Squaric acid diesters can be applied as reagents to couple two amino-functional compounds. Consecutive coupling of two amines allows the synthesis of asymmetric squaric acid bisamides with either low molecular weight compounds but also biomolecules or polymers. The key feature of the squaric acid diester mediated coupling is the reduced reactivity of the resulting ester-amide after the first amidation step of the diester. This allows the sequential amidation and generation of asymmetric squaramides with high selectivity and in high yields. This article gives an overview of the well-established squaric acid diester mediated coupling reactions for glycoconjugates and presents recent advances that aim to expand this very versatile reaction protocol to the modification of peptides and proteins.
Article
Recent claims that linear relationships exist between energetic, geometric, and magnetic criteria of aromaticity are shown to be invalid for any representative set of heteroaromatics in which the number of heteroatoms varies.
Article
We report herein experimental and theoretical evidence for an aromatic Cope rearrangement. Along with several successful examples, our data include the first isolation and full characterization of the putative intermediate that is formed immediately after the initial [3,3] sigmatropic rearrangement. Calculations at the B3LYP/6-31G(d) level of theory predict reaction energy barriers in the range 22-23 kcal/mol for the [3,3]-rearrangement consistent with the exceptionally mild reaction conditions for these reactions. The experimental and computational results support a significant enthalpic contribution of the concomitant pyrazole ring formation that serves as both a kinetic and thermodynamic driving force for the aromatic Cope rearrangement.