ArticlePDF Available

Energetically Demanding Transport in a Supramolecular Assembly

Authors:

Abstract and Figures

A challenge in contemporary chemistry is the realization of artificial molecular machines that can perform work in solution on their environments. Here, we report on the design and production of a supramolecular flashing energy ratchet capable of processing chemical fuel generated by redox changes to drive a ring in one direction relative to a dumbbell toward an energetically uphill state. The kinetics of the reaction pathway juxtapose a low energy [2]pseudorotaxane that forms under equilibrium conditions with a high energy, metastable [2]pseudorotaxane which resides away from equilibrium.
Content may be subject to copyright.
Energetically Demanding Transport in a Supramolecular Assembly
Chuyang Cheng,
Paul R. McGonigal,
Wei-Guang Liu,
Hao Li,
Nicolaas A. Vermeulen,
Chenfeng Ke,
Marco Frasconi,
Charlotte L. Stern,
William A. Goddard III,
and J. Fraser Stoddart*
,
Department of Chemistry, Northwestern University, Evanston, Illinois 60208, United States
Materials and Process Simulation Center, California Institute of Technology, Pasadena, California 91125, United States
*
SSupporting Information
ABSTRACT: A challenge in contemporary chemistry is
the realization of articial molecular machines that can
perform work in solution on their environments. Here, we
report on the design and production of a supramolecular
ashing energy ratchet capable of processing chemical fuel
generated by redox changes to drive a ring in one direction
relative to a dumbbell toward an energetically uphill state.
The kinetics of the reaction pathway juxtapose a low
energy [2]pseudorotaxane that forms under equilibrium
conditions with a high energy, metastable [2]-
pseudorotaxane which resides away from equilibrium.
The active transport of ions and small molecules across cell
membranes, driven by bespoke biochemical machinery,
plays a pivotal role in the operation of cells.
1
Natures
transmembrane protein pumps create concentration gradients
of ions, such as Na+,K
+and Ca2+, in addition to protons,
2
whose
stored potential may then be used as a secondary energy
resource
3
in metabolic processes, e.g., ATP synthesis.
4
During
billions of years of evolution, these proteins have developed
nely tuned secondary and tertiary structures capable of
harnessing external fuel to exert precise control over the
noncovalent forcesthe potential energy landscape of energy
barriers and wellsexperienced by their cargos in order to drive
them energetically uphill, temporarily away from thermodynamic
equilibrium.
5
Over the past 30 years, chemists have taken advantage of the
restricted degrees of freedom available to the components of
rotaxanes
6
and pseudorotaxanes
7
to construct and control
increasingly sophisticated articial molecular switches that
exhibit large amplitude relative motions between their
constituent parts.
8
In the majority of cases, however, the
application of a stimulus to these articial systems induces
molecular motions toward a low-energy equilibrium state. The
development of articial molecular assemblies
9
that can drive a
cargo away from its equilibrium position by manipulating kinetic
pathwaysthereby mimicking the function of their biochemical
counterpartsis still very much in its infancy.
10
Here, we report a
family of synthetic supramolecular pumps that exploit chemical
energy to drive a ring away from its initial state in solution to form
metastable, energetically demanding products that can then
gradually decay back toward equilibrium. Control of this
supramolecular transport is exercised by inuencing the kinetics
of the reaction pathway using only reversible noncovalent
bonding interactions and without the need to make or break
covalent bonds. The end result is that the ring is ensnared on an
oligomethylene chain
11
for which it has little to no binding
anity at any stage of the pumping process.
We designed (Figure 1) a homologous series of dumbbells
DB193+ bearing 4,4-bipyridinium (BIPY2+) units as radical
recognition sites
12
for cyclobis(paraquat-p-phenylene)
13
(CBPQT4+). A 3,5-dimethylpyridinium (PY+) unit is attached
to one side of the BIPY2+ unit by a short oligomethylene chain,
while a bulky stopper is attached to the other side by a longer
chain. Before testing the dumbbells, a symmetrical derivative
DB04+ (Figure 1) was prepared as a model compound to conrm
that CBPQT2(+)can thread across the PY+barrier in order to
form the trisradical complex DB03+CBPQT2(+)under the
reducing conditions that can bring about radicalradical
interactions. The absorption spectra of the radical cation
DB03+and diradical dication CBPQT2(+)were rst of all
recorded separately in MeCN at concentrations of 104M. No
absorption indicative of BIPY+dimerization was observed in the
near-IR (NIR) region under these conditions. By contrast, when
Received: August 21, 2014
Published: September 25, 2014
Figure 1. Structural formulas and simplied graphical representations of
the CBPQT4+ ring, the symmetrical dumbbell DB04+ and the nine one-
stroke supramolecular pumpsDB193+ employed in this structure
function investigation. Charges are balanced by PF6
counterions which
are omitted for the sake of clarity.
Communication
pubs.acs.org/JACS
© 2014 American Chemical Society 14702 dx.doi.org/10.1021/ja508615f |J. Am. Chem. Soc. 2014, 136, 1470214705
Downloaded via NORTHWESTERN UNIV on June 21, 2018 at 20:53:59 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
DB04+ and CBPQT4+ were reduced together, a distinctive purple
solution was obtained within seconds and UVvisNIR
spectroscopy revealed (Figure 2a) features characteristic of
BIPY+radicalradical interactions, including a broad NIR band
around 1100 nm that is indicative
12
of the formation of the
trisradical inclusion complex DB03+CBPQT2(+).Single
crystals of (DB0CBPQT)·6PF6suitable for X-ray diraction
(XRD)
14
were grown in a glovebox.
15
The solid-state super-
structure conrmed (Figure 2b) that the ring encircles the
dumbbell.
With the viability of complex formation between CBPQT2(+)
and DB03+veried, we hypothesized that subjecting a mixture of
CBPQT4+ and a dumbbell DB193+ (Figure 1) to a cycle of
reduction and oxidation would result in the formation of a high-
energy [2]pseudorotaxane. To test our hypothesis, activated zinc
dust was added to a 1:1 mixture of DB13+and CBPQT4+ in
CD3CN, giving a purple solution indicative of trisradical complex
formation within a few seconds. After ltration to remove the
excess of zinc, tris(4-bromophenyl)aminium hexachloridoantim-
onate (magic blue) was added at 0 °C to reoxidize the radical
species to their fully charged states, and a 1H NMR spectrum was
recorded. Comparison of the 1H NMR spectra of the dumbbell,
the reaction mixture, and the ring revealed (Figure 3a,b,d)
formation of a product species in 80% yield that exhibited
broadening (Hα/β) and splitting (HCH2) of resonances associated
with the ring as well as signicant upeld shifts of the methylene
proton resonances H14of the dumbbell component, indicating
the formation of DB13+CBPQT4+ in which the ring encircles
the hexamethylene chain.
12b,d
This [2]pseudorotaxane was
found to be metastable and to dissociate entirely within half an
hour at room temperature or within 4 h at 279 K (Figure 3c).
Changes in the relative proportions of DB13+CBPQT4+ and its
separated components were monitored (Figure S4)
15
by 1H
NMR spectroscopy at a constant temperature (279 K) which
revealed (Figure 3e) that the dethreading displays rst-order
kinetics with a rate constant k= (8.4 ±0.6) ×105s1. Based
upon these data, the transition-state energy barrier (ΔG) for the
CBPQT4+ ring to dissociate from DB13+ by slipping over its
positively charged BIPY2+ and PY+units was found to be 21.5
kcal mol1.
It appears, therefore, that the design of dumbbell DB14+ allows
it to pumpa ring actively onto its hexamethylene chain
following potential energy surfaces resembling those represented
as curves in Scheme 1. Initially, the tetracationic CBPQT4+ ring is
repelled (Scheme 1a) by both the PY+and BIPY2+ units.
Reduction, however, simultaneously lowers the kinetic barrier to
threading
7d
that comes from Coulombic repulsion between the
PY+unit and the ring, while also bringing radicalradical
interactions into play to create (Scheme 1b) a new global
minimum in which the dumbbell is encircled by the ring. As a
result, the ring slips over the PY+unit to form (Scheme 1c) a
thermodynamically stable trisradical inclusion complex
12
BIPY+CBPQT2(+). A subsequent oxidation step to restore
the repulsion between the highly charged ring and dumbbell
components
12d
then imparts a driving force for the CBPQT4+
ring to undergo translational motion. In thermodynamic terms,
the global energy minimum would be reached if the ring were to
Figure 2. UVvisNIR absorption spectroscopic and single crystal
XRD evidence that CBPQT2(+)threads across the Coulombic PY+
barrier. (a) Spectra of DB03+(black trace), CBPQT2(+)(blue trace),
and a 1:1 mixture of DB03+and CBPQT2(+)(purple trace). The broad
peak around 1100 nm indicates formation of the trisradical inclusion
complex DB03+CBPQT2(+).[DB4+] = 1.0 ×104M, [CBPQT4+]=
1.0 ×104M. Solvent, MeCN; T= 298 K. (b) A tubular representation
of the solid-state superstructure of (DB0CBPQT)·6PF6.
14,15
Solvent
molecules, hydrogen atoms, and counterions are omitted for the sake of
clarity.
Figure 3. 1H NMR spectroscopic analysis of [2]pseudorotaxane formation and the kinetic barrier to disocciation. Partial 1H NMR spectra (600 MHz,
CD3CN, 279 K) of (a) the dumbbell DB13+, (b) a reaction mixture containing DB13+CBPQT4+ after performing a cycle of reduction and oxidation,
(c) after standing at 279 K for 4 h, and (d) CBPQT4+. Peaks are highlighted in red and blue for DB13+ and CBPQT4+, respectively. The changes in
chemical shifts (blue and black dashed lines) are indicative of [2]pseudorotaxane formation. (e) Plot of the change in molar fraction over time at 279 K
based on integration of the 1H NMR spectra. Inset: linear relationship between ln(c/c0) and time, indicating that the dethreading process obeys rst-
order kinetics. The activation barrier ΔGwas calculated using the Eyring equation.
Journal of the American Chemical Society Communication
dx.doi.org/10.1021/ja508615f |J. Am. Chem. Soc. 2014, 136, 147021470514703
retrace its path and dissociate from the dumbbell. The
electrostatic barrier to translation originating from the PY+
group, however, is restored upon oxidation, kinetically
disfavoring the backward pathway and forcing the ring to shuttle
away (Scheme 1d) in the opposite direction. Consequently, the
CBPQT4+ ring is ensnared (Scheme 1e) as part of a metastable
[2]pseudorotaxane that does not benetfromstabilizing
noncovalent bonding interactions. Indeed, the enforced
proximity of the three charged viologen units dictates that the
[2]pseudorotaxane resides in an energetically demanding state
compared to its separated components.
By performing the same reduction and oxidation protocol with
the entire homologous series of dumbbells DB283+, we were
able to gain some insight into the relationship between the
structure and energetics of the supramolecular pumps. Extension
of the oligomethylene chain between the BIPY2+ and stopper
units resulted in two pseudorotaxanes, namely
DB23+CBPQT4+ and DB33+CBPQT4+, that exhibited
prolonged half-lives compared to DB13+CBPQT4+.The
dissociations of DB23+CBPQT4+ and DB33+CBPQT4+
were monitored by 1H NMR spectroscopy at 331 and 365 K,
respectively, which allowed the kinetic barriers to dethreading to
be determined (Figures S5 and S6)
15
as 25.3 and 29.0 kcal mol1.
Although the intrinsic barrier to association ΔG0(Scheme 1a),
which stems mainly from Coulombic repulsion between
positively charged BIPY2+/PY+units and the CBPQT4+ ring, is
similar for dumbbells DB133+, the large variation in the kinetic
barriers to dethreading can be rationalized by considering the
dierence in the Gibbs free energy ΔGof the metastable
[2]pseudorotaxanes. A shorter oligomethylene chain connes
the like-charged CBPQT4+ ring and BIPY2+ unit in closer
proximity to one another, destabilizing the pseudorotaxane and
resulting in a higher ΔG.
12d
Density functional theory (DFT)
calculations support this reasoning. We calculated (Table S3)
15
increases in enthalpy (ΔH) of 14.9, 8.8, and 7.0 kcal mol1for the
[2]pseudorotaxanes based on DB13+,DB23+,andDB33+,
respectively, compared to their separated components.
The oligomethylene portion of the dumbbell can be replaced
by an oligoethylene glycol chain without diminishing the
eciency of the pumping process. Dumbbells DB43+ and
DB53+, bearing triethylene glycol and tetraethylene glycol tails,
respectively, gave rise to [2]pseudorotaxanes in 85% yield, as
estimated by 1H NMR spectrocopy.
15
While the dissociation of
DB43+CBPQT4+ follows (Figure S7)
15
rst-order kinetics in a
manner consistent with the oligomethylene derivatives, our
observations indicate that the dissociation (Figure S8)
15
of
DB53+CBPQT4+ is autocatalytic. The most probable explan-
ation for this phenomenon is that hydrogen bonding between the
tetraethylene glycol chain and CBPQT4+ ring, which is a well-
established interaction,
16
stabilizes the transition state to
dethreading.
7d,17
Indeed, the addition of 1.0 equiv of dumbbell
DB53+ to a solution of DB53+CBPQT4+ accelerated the
dissociation of the [2]pseudorotaxane.
15
In order to avoid this
complication in subsequent investigations, we decided to limit
dumbbell design to those containing oligomethylene chains.
Having explored the eect of altering the length and nature of
the chain between the BIPY2+ unit and the stopper, we turned
our attention to assessing the inuence of the short oligo-
methylene linker connecting the Coulombic barrier
18
PY+to the
rest of the dumbbell. Taking DB23+ as a point of reference, two
dumbbells were synthesized bearing a butylene (DB63+) and a
bismethylene (DB73+) linker in place of the propylene one in
DB23+. Surprisingly, this seemingly minor alteration in molecular
structure was found to have a dramatic eect on the outcomes of
redox cycling. Subjecting a mixture of CBPQT4+ and DB63+ to
the standard protocol gave only a physical mixture after
reduction and oxidation (Figure S9),
15
suggesting that
elongating the linker diminishes considerably the ability of the
BIPY2+ and PY+units to act together in creating a highly charged
end group that retains the ring on the dumbbell on account of
Coloumbic repulsion. In contrast, when CBPQT4+ and DB73+
were reduced and oxidized under the same conditions, a bench
stable [2]rotaxane was produced that could be isolated (Figure
S1)
15
and fully characterized by NMR spectroscopy and mass
spectrometry. Indeed, the stability of the mechanically
interlocked product precluded measurement of any barrier to
dethreading; no dissociation could be observed (Figure S10),
15
even after reuxing in MeCN for 18 h! In order to quantify the
change in the energy barrier that occurs upon shortening the
propylene linker, we prepared DB83+ with a bismethylene linker
and a hexamethylene chain on which to collect the ring, rather
than the octamethylene one in DB73+. The [2]pseudorotaxane
DB83+CBPQT4+ dissociates at 355 K with a rate constant k=
(2.2 ±0.1) ×105s1, corresponding to a dethreading barrier of
28.5 kcal mol1(Figure S11).
15
It is quite remarkable that, with
the removal of a single methylene group from DB13+ (propylene
linker) to give DB83+ (bismethylene linker), we observe an
increase of 7.0 kcal mol1in the barrier to dissociation, from 21.5
to 28.5 kcal mol1. This phenomenon can be accounted for to a
certain extent by considering that the magnitude of Coloumbic
forces are inversely proportional to the square of the distance
between two charged species and that the longer the linker, the
further the ring resides away from the PY+unit. DFT calculations
indicate, however, that the interplay between the PY+and
BIPY2+ units also plays a role. The calculated potential energy
prole for dissociation of DB83+CBPQT4+ is characterized
(Figure S12)
15
by one major energy barrier of 18.7 kcal mol1.
The prole of a model [2]pseudorotaxane analog that contains a
butylene linker, DB93+CBPQT4+, exhibits (Figure S13)
15
two
smaller barriers, the highest of which is only 12.4 kcal/mol. It
Scheme 1. Graphical Representation of the Pumping
Mechanism and Corresponding Energy Proles
Journal of the American Chemical Society Communication
dx.doi.org/10.1021/ja508615f |J. Am. Chem. Soc. 2014, 136, 147021470514704
appears, therefore, that the dense buildup of charge between
closely tethered BIPY2+ and PY+units (e.g., in DB783+) acts in
an additive manner in repelling the slippage of the CBPQT4+
ring, although this eect is diminished if a longer spacer is
employed.
19
In summary, we have designed, synthesized, and operated a
family of articial supramolecular pumps which couple the
dissipation of redox chemical energy to the formation of a non-
equilibrium state. The alternation between the asymmetric
potential energy surfaces that they create shapes the kinetically
favored pathway followed by a positively charged ring, leading to
a metastable [2]pseudorotaxane in which the ring is trapped on a
dumbbell in the absence of any stabilizing interactions. By
studying a homologous series of supramolecular pumps and their
respective [2]pseudorotaxanes, we have elucidated the delicate
balance between molecular structure and the energetics of the
pumping process. Judicious choice of molecular structure allows
for both high-eciency [2]pseudorotaxane formation and the
regulation of its subsequent dissociation over a Coloumbic
barrier, which can be tuned within a range >10 kcal mol1. The
supramolecular pumps we have described exhibit some of the
qualities of the active transport
1,2
of ions and small molecules
that takes place in Naturei.e., they harness an energy input in
order to oscillate between asymmetric potential energy surfaces
and, thus, transport a species to a high-energy state. In their
current form, however, these synthetic supramolecular pumps
are unable to act in the repetitive manner reminiscent of their
natural counterparts in order to push a system incrementally
further and further from equilibrium, as occurs, e.g., when carrier
proteins create concentration gradients. By bringing the
knowledge gained in this structurefunction study to bear on
the design of synthetic supramolecular pumps, it may be possible
to devise more sophisticated systems capable of this kind of
repetitive and progressive action.
ASSOCIATED CONTENT
*
SSupporting Information
Detailed synthesis procedures and characterization data (NMR
spectroscopy and mass spectrometry) for all compounds. UV
visNIR spectroscopic and single crystal XRD characterization
of the [2]pseudorotaxane (DB0CBPQT)·6PF6. This material
is available free of charge via the Internet at http://pubs.acs.org.
AUTHOR INFORMATION
Corresponding Author
stoddart@northwestern.edu
Notes
The authors declare no competing nancial interest.
ACKNOWLEDGMENTS
This material is based on work supported by the National Science
Foundation (NSF) under CHE-1308107. W.-G.L. and W.A.G.
thank NSF (CMMI-1120890 and EFRI-1332411) for nancial
support and to ONR-DURIP and NSF-CSEM for computing
resources.
REFERENCES
(1) Higgins, C. F. Annu. Rev. Cell. Biol. 1992,8, 67.
(2) (a) Skou, J. C. Angew. Chem., Int. Ed. 1998,37, 2321. (b) Kaplan, J.
H. Annu. Rev. Biochem. 2002,71, 511. (c) MacLennan, D. H.; Rice, W. J.;
Green, N. M. J. Biol. Chem. 1997,272, 28815. (d) Lozier, R. H.;
Bogomolni, R. A.; Stoeckenius, W. Biophys. J. 1975,15, 955.
(3) (a) Wright, E. M.; Loo, D. D. F.; Hirayama, B. A. Physiol. Rev. 2011,
91, 733. (b) West, I. C. Biochim. Biophys. Acta 1980,604, 91.
(4) (a) Walker, J. E. Angew. Chem., Int. Ed. 1998,37, 2309. (b) Elston,
T.; Wang, H. Y.; Oster, G. Nature 1998,391, 510. (c) Boyer, P. D.
Angew. Chem., Int. Ed. 1998,37, 2297.
(5) (a) Gai, F.; Hasson, K. C.; McDonald, J. C.; Anfinrud, P. A. Science
1998,279, 1886. (b) Subramaniam, S.; Henderson, R. Nature 2000,406,
653. (c) Astumian, R. D. Nat. Nanotechnol. 2012,7, 684.
(6) Stoddart, J. F. Chem. Soc. Rev. 2009,38, 1802.
(7) (a) Nygaard, S.; Laursen, B. W.; Flood, A. H.; Hansen, C. N.;
Jeppesen, J. O.; Stoddart, J. F. Chem. Commun. 2006, 144. (b) Arduini,
A.; Bussolati, R.; Credi, A.; Monaco, S.; Secchi, A.; Silvi, S.; Venturi, M.
Chem.-Eur. J. 2012,18, 16203. (c) Baroncini, M.; Silvi, S.; Venturi, M.;
Credi, A. Angew. Chem., Int. Ed. 2012,51, 4223. (d) Li, H.; Cheng, C. Y.;
McGonigal, P. R.; Fahrenbach, A. C.; Frasconi, M.; Liu, W. G.; Zhu, Z.
X.; Zhao, Y. L.; Ke, C. F.; Lei, J. Y.; Young, R. M.; Dyar, S. M.; Co, D. T.;
Yang, Y. W.; Botros, Y. Y.; Goddard, W. A., III; Wasielewski, M. R.;
Astumian, R. D.; Stoddart, J. F. J. Am. Chem. Soc. 2013,135, 18609.
(8) (a) Balzani, V.; Credi, A.; Raymo, F. M.; Stoddart, J. F. Angew.
Chem., Int. Ed. 2000,39, 3349. (b) Collin, J.-P.; Heitz, V.; Sauvage, J.-P.
Top. Curr. Chem. 2005,262, 29.
(9) (a) Feringa, B. L. J. Org. Chem. 2007,72, 6635. (b) Michl, J.; Sykes,
E. C. H. ACS Nano 2009,3, 1042. (c) Vogelsberg, C. S.; Garcia-Garibay,
M. A. Chem. Soc. Rev. 2012,41, 1892.
(10) (a) Serreli, V.; Lee, C. F.; Kay, E. R.; Leigh, D. A. Nature 2007,
445, 523. (b) Kay, E. R.; Leigh, D. A.; Zerbetto, F. Angew. Chem., Int. Ed.
2007,46, 72. (c) Coskun, A.; Banaszak, M.; Astumian, R. D.; Stoddart, J.
F.; Grzybowski, B. A. Chem. Soc. Rev. 2012,41, 19.
(11) (a) Goldup, S. M.; Leigh, D. A.; McBurney, R. T.; McGonigal, P.
R.; Plant, A. Chem. Sci. 2010,1, 383. (b) Sevick, E. M.; Williams, D. R. M.
Molecules 2013,18, 13398.
(12) (a) Trabolsi, A.; Khashab, N.; Fahrenbach, A. C.; Friedman, D. C.;
Colvin, M. T.; Coti, K. K.; Benítez, D.; Tkatchouk, E.; Olsen, J. C.;
Belowich, M. E.; Carmielli, R.; Khatib, H. A.; Goddard, W. A., III;
Wasielewski, M. R.; Stoddart, J. F. Nat. Chem. 2010,2, 42. (b) Li, H.;
Fahrenbach, A. C.; Dey, S. K.; Basu, S.; Trabolsi, A.; Zhu, Z. X.; Botros,
Y. Y.; Stoddart, J. F. Angew. Chem., Int. Ed. 2010,49, 8260.
(c) Fahrenbach, A. C.; Barnes, J. C.; Lanfranchi, D. A.; Li, H.;
Coskun, A.; Gassensmith, J. J.; Liu, Z. C.; Benítez, D.; Trabolsi, A.;
Goddard, W. A., III; Elhabiri, M.; Stoddart, J. F. J. Am. Chem. Soc. 2012,
134, 3061. (d) Li, H.; Zhu, Z. X.; Fahrenbach, A. C.; Savoie, B. M.; Ke, C.
F.; Barnes, J. C.; Lei, J. Y.; Zhao, Y. L.; Lilley, L. M.; Marks, T. J.; Ratner,
M. A.; Stoddart, J. F. J. Am. Chem. Soc. 2013,135, 456.
(13) Odell, B.; Reddington, M. V.; Slawin, A. M. Z.; Spencer, N.;
Stoddart, J. F.; Williams, D. J. Angew. Chem., Int. Ed. 1988,27, 1547.
(14) Crystals were grown by slow vapor diusion of iPr2O into a
MeCN solution at 0 °C. Crystal data for C82H94F36N16P6,(M=
2173.55): monoclinic, space group P21/c(no. 14), a= 18.8758(10), b=
20.3394(13), c= 13.1067(8) Å, β= 104.883(3)°,V= 4863.1(5) Å3,T=
99.98 K, Z=2,μ(CuKα) = 2.114, 30255 reections measured, 8210
unique (Rint = 0.0327) which were used in all calculations. wR(F2)=
0.1898. CCDC 1019889.
(15) See the SI.
(16) Zhu, Z. X.; Li, H.; Liu, Z. C.; Lei, J. Y.; Zhang, H. C.; Botros, Y. Y.;
Stern, C. L.; Sarjeant, A. A.; Stoddart, J. F.; Colquhoun, H. M. Angew.
Chem., Int. Ed. 2012,51, 7231.
(17) Andersen, S. S.; Share, A. I.; Poulsen, B. L.; Korner, M.; Duedal,
T.; Benson, C. R.; Hansen, S. W.; Jeppesen, J. O.; Flood, A. H. J. Am.
Chem. Soc. 2014,136, 6373.
(18) (a) Li, H.; Zhao, Y. L.; Fahrenbach, A. C.; Kim, S. Y.; Paxton, W.
F.; Stoddart, J. F. Org. Biomol. Chem. 2011,9, 2240. (b) Hmadeh, M.;
Fahrenbach, A. C.; Basu, S.; Trabolsi, A.; Benítez, D.; Li, H.; Albrecht-
Gary, A. M.; Elhabiri, M.; Stoddart, J. F. Chem.Eur. J. 2011,17, 6076.
(19) Taking all of the calculated and experimentally measured
thermodynamic data together (Table S18), the fraction of the redox
energy input that is stored in the high-energy pseudorotaxanes can be
calculated to be between 5 and 10%.
Journal of the American Chemical Society Communication
dx.doi.org/10.1021/ja508615f |J. Am. Chem. Soc. 2014, 136, 147021470514705
... To illustrate their operation, we introduce the endergonic assembly of a pseudorotaxane driven by the alternation of redox stimuli. [16] The system is composed of macrocycle 1 4 + and dumbbell 2 3 + (Figure 2), which comprise redox-active viologen moieties in their structure. Due to charge repulsion, the formation of corresponding pseudorotaxane [2 3 + �1 4 + ] is negligible at millimolar concentration in an organic solvent-the typical experimental conditions. ...
... [26] In principle, alternation of any condition can afford a non-autonomous energy ratchet mechanism, provided that the thermodynamic and kinetic conditions outlined above are respected. For example, not only alternating redox [16,17] or acid/base conditions, [27,28] but also temperature, humidity, magnetic field, or other parameters could result in conceptually analogous effects. In fact, some of the first purposely-designed energy ratchets featured a combination of orthogonal stimuli. ...
... For example, when a dumbbell as simple as 2 3 + bears a glycol collecting chain instead of an alkyl one, the dethreading reaction becomes autocatalytic. [16] Considering how much complexity can be hidden in a simple dethreading reaction, where large supramolecular assemblies are investigated, kinetic schemes can only serve as a rough guide. Yet, they are useful even when seeking emergent phenomena, such as oscillations. ...
Article
Full-text available
Non‐equilibrium chemical systems underpin multiple domains of contemporary interest, including supramolecular chemistry, molecular machines, systems chemistry, prebiotic chemistry, and energy transduction. Experimental chemists are now pioneering the realization of artificial systems that can harvest energy away from equilibrium. In this tutorial Review, we provide an overview of artificial molecular ratchets: the chemical mechanisms enabling energy absorption from the environment. By focusing on the mechanism type—rather than the application domain or energy source—we offer a unifying picture of seemingly disparate phenomena, which we hope will foster progress in this fascinating domain of science.
... [4,19] As a general example, we consider a square reaction scheme for pseudorotaxane assembly that can be traveled unidirectionally upon alternating acid-base stimuli, leading to the relocation of a macrocycle to an energetically-unfavored location along a molecular axle ( Figure 1). [20][21][22][23] This choice stems from the simplicity of pH changes as stimuli to power molecular machinery and the importance of simplicity for the progress of the field. [24] When the axle is deprotonated -state i -the separated components are thermodynamically more stable than the pseudorotaxane; however, protonation of the axle forms a recognition site for the ring -in state ii -leading to thermodynamically favored assembly in state iii. ...
... [38,39] So far, studies that approached self-assembly kinetics investigated the threading processes affording thermodynamically stable complexes, whereas the dethreading kinetics from high-energy states remained essentially unexplored. [22] To advance the understanding of minimal molecular ratchets, we focused on a versatile molecular pump module developed by Leigh and coworkers, capable of accumulating macrocycles in a high-energy state upon a sequence of acidbase stimuli. [20,40] The operating mechanism in action is based on the self-assembly of a crown ether macrocycle onto a dibenzyl ammonium ion. ...
Article
Full-text available
The operation of nanomachines is fundamentally different from that of their macroscopic counterparts. In particular, the role of solvent is critical yet rarely associated with machine functionality. Here, we study a minimal model of one of the most advanced molecular machines to gain control of its operation by engineering components and the employed solvent. Operation kinetics were changed over more than four orders of magnitude and could be modulated by solvent. Leveraging solvent properties, it was possible to monitor the relaxation of the molecular machine towards equilibrium and measure the heat exchanged in the process. Our work expands the capabilities of acid‐base powered molecular machines, confirming experimentally that such systems have a dominant entropy content.
... Additional references are cited within the Supporting Information. [39][40][41][42][43] ...
Article
Full-text available
Cyclobis(paraquat‐p‐phenylene), also known as “blue box”, is a highly electron‐deficient macrocycle, widely used as a molecular receptor for small electron‐rich molecules. Inserting a reactive functional group onto the molecular structure of this cyclophane is paramount for its inclusion into complex architectures. To this aim, including an alkyne moiety would be ideal, because it can participate in click reactions. However, the synthesis of such alkyne‐functionalized cyclophane suffers from several drawbacks: the use of toxic and expensive CCl4, the need for high‐pressure reactors, and overall low yield. We have revised the existing synthesis of this cyclophane derivative bearing an alkyne moiety, to overcome all these limitations. In particular, photochemical radical bromination is adopted to obtain a sensitive intermediate. We demonstrated that the synthesized host molecule can be functionalized via click reactions and take part in radical‐radical interactions. Our work makes a key functionalized paraquat macrocycle more accessible, facilitating the development of novel redox‐responsive systems.
... Building on this concept, a one-stroke molecular pump, D2 3 + , was synthesized [226] by connecting a 3,5-DMPy + Coulombic barrier and a long collecting chain, terminated by a 2,6-diisopropylphenyl stopper, to a viologen recognition site. Under reducing conditions, the attraction between the CBPQT 2(* +) ring and the V * + recognition site creates (Figure 18b) a deep thermodynamic well for the ring to pass over the low kinetic barrier imposed by 3,5-DMPy + and encircle the dumbbell. ...
Article
Full-text available
The tetracationic cyclophane, cyclobis(paraquat‐p‐phenylene), also known as the little blue box, constitutes a modular receptor that has facilitated the discovery of many host–guest complexes and mechanically interlocked molecules during the past 35 years. Its versatility in binding small π‐donors in its tetracationic state, as well as forming trisradical tricationic complexes with viologen radical cations in its doubly reduced bisradical dicationic state, renders it valuable for the construction of various stimuli‐responsive materials. Since the first reports in 1988, the little blue box has been featured in over 500 publications in the literature. All this research activity would not have been possible without the seminal contributions carried out by Siegfried Hünig, who not only pioneered the syntheses of viologen‐containing cyclophanes, but also revealed their rich redox chemistry in addition to their ability to undergo intramolecular π‐dimerization. This Review describes how his pioneering research led to the design and synthesis of the little blue box, and how this redox‐active host evolved into the key component of molecular shuttles, switches, and machines.
Article
In recent years ratchet mechanisms have transformed the understanding and design of stochastic molecular systems—biological, chemical and physical—in a move away from the mechanical macroscopic analogies that dominated thinking regarding molecular dynamics in the 1990s and early 2000s (e.g. pistons, springs, etc), to the more scale‐relevant concepts that underpin out‐of‐equilibrium research in the molecular sciences today. Ratcheting has established molecular nanotechnology as a research frontier for energy transduction and metabolism, and has enabled the reverse engineering of biomolecular machinery, delivering insights into how molecules ‘walk’ and track‐based synthesizers operate, how the acceleration of chemical reactions enables energy to be transduced by catalysts, and how dynamic (supra)molecular systems can be driven away from equilibrium. The recognition of molecular ratchet mechanisms in biology, and their invention in synthetic systems, is proving significant in areas as diverse as supramolecular chemistry, systems chemistry, dynamic covalent chemistry, DNA nanotechnology, polymer and materials science, molecular biology, heterogeneous catalysis, endergonic synthesis, and many other branches of chemical science. Put simply, ratchet mechanisms give chemistry direction. Kinetic asymmetry, quantified by the outcome of ratcheting, is the dynamic counterpart of structural asymmetry (i.e. chirality). Given the ubiquity of ratchet mechanisms, it is surely just as fundamentally important.
Article
Full-text available
Over the last two decades ratchet mechanisms have transformed the understanding and design of stochastic molecular systems—biological, chemical and physical—in a move away from the mechanical macroscopic analogies that dominated thinking regarding molecular dynamics in the 1990s and early 2000s (e.g. pistons, springs, etc), to the more scale‐relevant concepts that underpin out‐of‐equilibrium research in the molecular sciences today. Ratcheting has established molecular nanotechnology as a research frontier for energy transduction and metabolism, and has enabled the reverse engineering of biomolecular machinery, delivering insights into how molecules ‘walk’ and track‐based synthesisers operate, how the acceleration of chemical reactions enables energy to be transduced by catalysts (both motor proteins and synthetic catalysts), and how dynamic systems can be driven away from equilibrium through catalysis. The recognition of molecular ratchet mechanisms in biology, and their invention in synthetic systems, is proving significant in areas as diverse as supramolecular chemistry, systems chemistry, dynamic covalent chemistry, DNA nanotechnology, polymer and materials science, molecular biology, heterogeneous catalysis, endergonic synthesis, the origin of life, and many other branches of chemical science. Put simply, ratchet mechanisms give chemistry direction. Kinetic asymmetry, the key feature of ratcheting, is the dynamic counterpart of structural asymmetry (i.e. chirality). Given the ubiquity of ratchet mechanisms in endergonic chemical processes in biology, and their significance for behaviour and function from systems to synthesis, it is surely just as fundamentally important. This Review charts the recognition, invention and development of molecular ratchets, focussing particularly on the role for which they were originally envisaged in chemistry, as design elements for molecular machinery. Different kinetically asymmetric systems are compared, and the consequences of their dynamic behaviour discussed. These archetypal examples demonstrate how chemical systems can be driven inexorably away from equilibrium, rather than relax towards it.
Article
Non‐equilibrium chemical systems underpin multiple domains of contemporary interest, including supramolecular chemistry, molecular machines, systems chemistry, prebiotic chemistry, and energy transduction. Experimental chemists are now pioneering the realization of artificial systems that can harvest energy away from equilibrium. In this tutorial review, we provide an overview of artificial molecular ratchets: the chemical mechanisms enabling energy absorption from the environment. By focusing on the mechanism type – rather than the application domain or energy source – we offer a unifying picture of seemingly disparate phenomena, which we hope will foster progress in this fascinating domain of science.
Article
The tetracationic cyclophane, cyclobis(paraquat‐p‐phenylene), also known as the little blue box, constitutes a modular receptor that has facilitated the discovery of many host‐gust complexes and mechanically interlocked molecules during the past 35 years. Its versatility in binding small π‐donors in its tetracationic state, as well as forming trisradical tricationic complexes with viologen radical cations in its doubly reduced bisradical dicationic state, renders it valuable for the construction of various stimuli‐responsive materials. Since the first reports in 1988, the little blue box has been featured in over 500 publications in the literature. All this research activity would not have been possible without the seminal contributions carried out by Siegfried Hünig, who, not only pioneered the syntheses of viologen‐containing cyclophanes, but also revealed their rich redox chemistry in addition to their ability to undergo intramolecular π‐dimerization. This review describes how his pioneering research led to the design and synthesis of the little blue box, and how this redox‐active host evolved into the key component of molecular shuttles, switches and machines.
Article
A helical humidity resistive sensor was fabricated by using solution processed supramolecular self-assembly of viologen substituted perylene diimide (PDI-DV). The nanohelices obtained in iPr2O / MeCN under −20 ℃, as...
Article
Full-text available
We examine a rod piston-rotaxane system, where the positions of several mobile rings on the axle are controlled by an external force acting on one of the rings. This allows us to access the translational entropy of the rings. For a simple rotaxane molecule with an axle that has uniform ring-axle interactions along its length, the molecule behaves like a miniature piston filled with a one-dimensional ideal gas. We then examine the effect of two stripes on the axle, having different ring-axle interactions with the mobile rings, so that one section is of high energy (repulsive) for the rings and another section is of lower energy (or attractive). This kind of rotaxane can exhibit rapid changes in displacement or force, and in particular, this molecule can exhibit a yield stress in which the piston suddenly compresses under a small increase in the applied force.
Article
The miniaturization of components used in the construction of working devices is being pursued currently by the large-downward (top-down) fabrication. This approach, however, which obliges solid-state physicists and electronic engineers to manipulate progressively smaller and smaller pieces of matter, has its intrinsic limitations. An alternative approach is a small-upward (bottom-up) one, starting from the smallest compositions of matter that have distinct shapes and unique properties—namely molecules. In the context of this particular challenge, chemists have been extending the concept of a macroscopic machine to the molecular level. A molecular-level machine can be defined as an assembly of a distinct number of molecular components that are designed to perform machinelike movements (output) as a result of an appropriate external stimulation (input). In common with their macroscopic counterparts, a molecular machine is characterized by 1) the kind of energy input supplied to make it work, 2) the nature of the movements of its component parts, 3) the way in which its operation can be monitored and controlled, 4) the ability to make it repeat its operation in a cyclic fashion, 5) the timescale needed to complete a full cycle of movements, and 6) the purpose of its operation. Undoubtedly, the best energy inputs to make molecular machines work are photons or electrons. Indeed, with appropriately chosen photochemically and electrochemically driven reactions, it is possible to design and synthesize molecular machines that do work. Moreover, the dramatic increase in our fundamental understanding of self-assembly and self-organizational processes in chemical synthesis has aided and abetted the construction of artificial molecular machines through the development of new methods of noncovalent synthesis and the emergence of supramolecular assistance to covalent synthesis as a uniquely powerful synthetic tool. The aim of this review is to present a unified view of the field of molecular machines by focusing on past achievements, present limitations, and future perspectives. After analyzing a few important examples of natural molecular machines, the most significant developments in the field of artificial molecular machines are highlighted. The systems reviewed include 1) chemical rotors, 2) photochemically and electrochemically induced molecular (conformational) rearrangements, and 3) chemically, photochemically, and electrochemically controllable (co-conformational) motions in interlocked molecules (catenanes and rotaxanes), as well as in coordination and supramolecular complexes, including pseudorotaxanes. Artificial molecular machines based on biomolecules and interfacing artificial molecular machines with surfaces and solid supports are amongst some of the cutting-edge topics featured in this review. The extension of the concept of a machine to the molecular level is of interest not only for the sake of basic research, but also for the growth of nanoscience and the subsequent development of nanotechnology.
Article
The puzzling results of ¹⁸O-exchange experiments churned in Paul Boyer's mind, and he realized that the proton-motive force generated upon oxidative phosphorylation is not used primarily for the synthesis of an ATP molecule, but instead its release. The concept of the binding change mechanism was born. For the formation of ATP from ADP and inorganic phosphate—one of the most important reactions in nature—catalysis by ATP synthase requires sequential conformational changes and a rotary mechanism that drives these changes; this enzyme is truly a remarkable molecular machine.
Article
The cyclic modulation of nucleotide-binding properties of the three catalytic β subunits by a series of conformational changes was an attractive explanation for the postulated binding change mechanism of ATP synthase. In the crystal structure of the catalytic F1 domain of this enzyme there is indeed a complex made up of three α subunits and three β subunits arranged in alternation around a central α-helical segment of the γ subunit. This complex is asymmetric owing to the different conformations of the β subunits. The change in conformation is brought about by rotation of the rigid yet curved segment, which has meanwhile been proven experimentally.
Article
The miniaturization of components used in the construction of working devices is being pursued currently by the large-downward (top-down) fabrication. This approach, however, which obliges solid-state physicists and electronic engineers to manipulate progressively smaller and smaller pieces of matter, has its intrinsic limitations. An alternative approach is a small-upward (bottom-up) one, starting from the smallest compositions of matter that have distinct shapes and unique properties—namely molecules. In the context of this particular challenge, chemists have been extending the concept of a macroscopic machine to the molecular level. A molecular-level machine can be defined as an assembly of a distinct number of molecular components that are designed to perform machinelike movements (output) as a result of an appropriate external stimulation (input). In common with their macroscopic counterparts, a molecular machine is characterized by 1) the kind of energy input supplied to make it work, 2) the nature of the movements of its component parts, 3) the way in which its operation can be monitored and controlled, 4) the ability to make it repeat its operation in a cyclic fashion, 5) the timescale needed to complete a full cycle of movements, and 6) the purpose of its operation. Undoubtedly, the best energy inputs to make molecular machines work are photons or electrons. Indeed, with appropriately chosen photochemically and electrochemically driven reactions, it is possible to design and synthesize molecular machines that do work. Moreover, the dramatic increase in our fundamental understanding of self-assembly and self-organizational processes in chemical synthesis has aided and abetted the construction of artificial molecular machines through the development of new methods of noncovalent synthesis and the emergence of supramolecular assistance to covalent synthesis as a uniquely powerful synthetic tool. The aim of this review is to present a unified view of the field of molecular machines by focusing on past achievements, present limitations, and future perspectives. After analyzing a few important examples of natural molecular machines, the most significant developments in the field of artificial molecular machines are highlighted. The systems reviewed include 1) chemical rotors, 2) photochemically and electrochemically induced molecular (conformational) rearrangements, and 3) chemically, photochemically, and electrochemically controllable (co-conformational) motions in interlocked molecules (catenanes and rotaxanes), as well as in coordination and supramolecular complexes, including pseudorotaxanes. Artificial molecular machines based on biomolecules and interfacing artificial molecular machines with surfaces and solid supports are amongst some of the cutting-edge topics featured in this review. The extension of the concept of a machine to the molecular level is of interest not only for the sake of basic research, but also for the growth of nanoscience and the subsequent development of nanotechnology.
Article
The cyclic modulation of nucleotide-binding properties of the three catalytic β subunits by a series of conformational changes was an attractive explanation for the postulated binding change mechanism of ATP synthase. In the crystal structure of the catalytic F1 domain of this enzyme there is indeed a complex made up of three α subunits and three β subunits arranged in alternation around a central α-helical segment of the γ subunit. This complex is asymmetric owing to the different conformations of the β subunits. The change in conformation is brought about by rotation of the rigid yet curved segment, which has meanwhile been proven experimentally.
Article
Mechanistic understanding of the translational movements in molecular switches is essential for designing machine-like prototypes capable of following set pathways of motion. To this end, we demonstrated that increasing the station-to-station distance will speed up the linear movements forward and slow down the movements backward in a homologous series of bistable rotaxanes. Four redox-active rotaxanes, which drove a cyclobis(paraquat-p-phenylene) (CBPQT(4+)) mobile ring between a tetrathiafulvalene (TTF) station and an oxyphenylene station, were synthesized with only variations to the lengths of the glycol linker connecting the two stations (n = 5, 8, 11, and 23 atoms). We undertook the first mechanistic study of the full cycle of motion in this class of molecular switch using cyclic voltammetry. The kinetics parameters (k, ΔG(⧧)) of switching were determined at different temperatures to provide activation enthalpies (ΔH(⧧)) and entropies (ΔS(⧧)). Longer glycol linkers led to modest increases in the forward escape (t1/2 = 60 to <7 ms). The rate-limiting step involves movement of the tetracationic CBPQT(4+) ring away from the singly oxidized TTF(+) unit by overcoming one of the thiomethyl (SMe) speed bumps before proceeding on to the secondary oxyphenylene station. Upon reduction, however, the return translational movement of the CBPQT(4+) ring from the oxyphenylene station back to the neutral TTF station was slowed considerably by the longer linkers (t1/2 = 1.4 to >69 s); though not because of a diffusive walk. The reduced rate of motion backward depended on folded structures that were only present with longer linkers.
Article
Motor molecules present in nature convert energy inputs, such as a chemical fuel or incident photons of light, into directed motion and force biochemical systems away from thermal equilibrium. The ability, not only to control relative movements of components in molecules, but also and to drive their components preferentially in one direction relative to each other using versatile stimuli, is one of the keys to future technological applications. Herein, we describe a wholly synthetic, small-molecule system which, under the influence of chemical reagents, electrical potential, or visible light, undergoes unidirectional relative translational motion. Altering the redox state of a cyclobis(paraquat-p-phenylene) ring simultaneously (i) inverts the relative heights of kinetic barriers presented by the two termini - one a neutral 2-isopropylphenyl group and the other a positively charged 3,5-dimethylpyridinium unit - of a constitutionally asymmetric dumbbell, which can impair threading/dethreading of a [2]pseudorotaxane, and (ii) controls the ring's affinity for a 1,5-dioxynaphthalene binding site located at the dumbbell's central core. The formation and subsequent dissociation of the [2]pseudorotaxane by passage of the ring over the neutral and positively charged termini of the dumbbell component in one, and only one, direction relatively defined has been demonstrated by (i) spectroscopic (1H NMR and UV/vis) means and cyclic voltammetry, as well as with (ii) DFT calculations and by (iii) comparison with control compounds in the shape of constitutionally symmetrical [2]pseudorotaxanes, one with two positively charged and the other with two neutral ends. Operation of the system relies solely on reversible, yet stable, noncovalent bonding interactions. Moreover, in the presence of a photosensitizer, visible light energy is the only fuel source that is needed to drive the unidirectional molecular translation, making it feasible to repeat the operation numerous times without the buildup of byproducts.
Article
An efficient, mild and operationally simple Ni-catalysed sp3-carbon-to-sp3-carbon homocoupling of unactivated alkyl bromides has been developed and utilised in the active metal template synthesis of an alkyl chain axle [2]rotaxane. The key to the transformation is the use of tridentate nitrogen-donor-atom (terpy or pybox derived) ligands which inhibit competing β-hydride elimination of alkyl-Ni intermediates.