Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Supramolecular Chemistry

ISSN: 1061-0278 (Print) 1029-0478 (Online) Journal homepage: http://www.tandfonline.com/loi/gsch20

Inclusion of neutral guests by water-soluble


macrocyclic hosts – a comparative thermodynamic
investigation with cyclodextrins, calixarenes and
cucurbiturils

Dong-Sheng Guo, Vanya D. Uzunova, Khaleel I. Assaf, Alexandra I. Lazar, Yu


Liu & Werner M. Nau

To cite this article: Dong-Sheng Guo, Vanya D. Uzunova, Khaleel I. Assaf, Alexandra I. Lazar, Yu
Liu & Werner M. Nau (2015): Inclusion of neutral guests by water-soluble macrocyclic hosts –
a comparative thermodynamic investigation with cyclodextrins, calixarenes and cucurbiturils,
Supramolecular Chemistry, DOI: 10.1080/10610278.2015.1105374

To link to this article: http://dx.doi.org/10.1080/10610278.2015.1105374

View supplementary material

Published online: 25 Nov 2015.

Submit your article to this journal

Article views: 28

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=gsch20

Download by: [University of Sussex Library] Date: 05 December 2015, At: 07:44
Supramolecular Chemistry, 2015
http://dx.doi.org/10.1080/10610278.2015.1105374

Inclusion of neutral guests by water-soluble macrocyclic hosts – a comparative


thermodynamic investigation with cyclodextrins, calixarenes and cucurbiturils
Dong-Sheng Guoa,b, Vanya D. Uzunovaa, Khaleel I. Assafa, Alexandra I. Lazara, Yu Liub and Werner M. Naua
a
Department of Life Sciences and Chemistry, Jacobs University Bremen, Bremen, Germany; bState Key Laboratory of Elemento-Organic Chemistry,
Department of Chemistry, Collaborative Innovation Center of Chemical Science and Engineering, Nankai University, Tianjin, P.R. China

ABSTRACT ARTICLE HISTORY


The driving forces of association between three different families of macrocycles as hosts, namely Received 31 August 2015
cyclodextrins (α-, β-, and γ-), p-sulfonatocalix[n]arenes (n = 4–6) as well as cucurbit[n]urils (n = 6–8), Accepted 2 October 2015
Downloaded by [University of Sussex Library] at 07:44 05 December 2015

and three different bicyclic azoalkane homologues as guests, namely 2,3-diazabicyclo[2.2.1] KEYWORDS
hept-2-ene (DBH), 2,3-diazabicyclo[2.2.2]oct-2-ene (DBO) as well as 2,3-diazabicyclo[2.2.3]non-2- Molecular recognition;
ene (DBN), were examined by means of calorimetric titrations, NMR spectroscopy and molecular thermodynamics;
dynamics simulation, all in aqueous solution. The small, spherical and uncharged guests preferably macrocycles; neutral guests;
bind inside the cavities of the medium sized hosts. The inclusion complexation by β-cyclodextrin hydrophobic effect
and p-sulfonatocalix[4]arene shows medium binding affinities (millimolar), while cucurbit[7]uril
macrocycle shows very strong binding (micromolar). For all types of macrocycles, the complex
formation is enthalpically driven (ΔH° < 0), accompanied by slightly unfavourable entropy changes
(ΔS° < 0). The results are discussed in terms of the flexibility of the hosts, the hydrophobic character
of their cavities and the release of high-energy water upon binding, and generalised by including
two additional guests, the ketones cyclopentanone and (+)-camphor.

Introduction guest is augmented for derivatives with protic or anionic


groups (8, 9), which reinforce the complexes on account
Cyclodextrins (α-CD, β-CD, γ-CD), p-sulfonatocalix[n]arenes of hydrogen bonding, ion–ion or ion–dipole interactions
(CX4, CX5, CX6) and cucurbit[n]urils (CB6, CB7, CB8) are for the CDs rims (1, 7), CXs (3) or CBs (10), respectively.
water-soluble macrocycles (see structures in Chart 1) In order to further investigate hydrophobic interactions,
which are well known to form host–guest inclusion com- neutral organic guests are attractive, since puristic, models.
plexes with a large range of compounds (1–4). In many Neutral guest binding by CDs (1, 11, 12), CXs (13–17) and
cases, the complexation is driven, fully or to a large degree, CBs (9, 10, 18–20) has received considerable experimental
by the hydrophobic effect (5, 6), which allows, in particular, attention, especially by means of isothermal titration cal-
non-polar organic residues to be encapsulated inside the orimetry (ITC), which is a powerful tool for determining
host cavities, concomitantly releasing high-energy water host–guest complex interactions; the technique not only
molecules into the bulk (7). Frequently, the affinity of the affords the binding constants (Ka), but also thermodynamic

CONTACT Dong-Sheng Guo dshguo@nankai.edu.cn; Werner M. Nau w.nau@jacobs-university.de


© 2015 Taylor & Francis
2 D.-S. Guo et al.
Downloaded by [University of Sussex Library] at 07:44 05 December 2015

Chart 1. Water-soluble macrocycles used in this study.

parameters (enthalpy and entropy changes, ΔH°, ΔS°). N N N


Systematic comparative studies of the complexation ther- N N N
Hendo' Hsyn
Hendo Hsyn Hendo
modynamics of neutral guests with these three classes of Hendo
Hb Hanti Hb Hb Hanti
macrocycles are not at hand. However, they are particularly Hexo Hexo Hexo
Hexo'

desirable to obtain insights into the hydrophobic effect DBH DBO DBN
that drives the binding to the respective class of macro-
cycles, CDs, CXs and CBs. O H3C CH3
In contrast to the situation for neutral guests, selected H
H
comparative studies on the binding of ionic guests with
the different macrocycles have already been conducted. H
O CH3
Mutihac and co-workers made a survey on thermodynamic
properties for the complexation of some calixarene vs. CB cyclopentanone (+)-camphor
receptors towards amino acids, nucleobases and di- and
Chart 2. Structures of the investigated bicyclic azoalkanes and
tri-peptides (21). García-Río and co-workers compared the the two ketones.
complexation of cationic surfactants with β-CD and CB7
by ITC (22). Whereas the binding affinity of surfactants
with β-CD increases with alkyl chain length, the binding with the three macrocycle classes by fluorescence titra-
constants between CB7 and surfactants remain essentially tions (23). CB7 always caused enhanced fluorescence of the
unchanged, indicating that CB7 exhibits a pronounced dyes, while CX4 always resulted in quenched fluorescence.
preference towards the cationic residues while β-CD pre- Moreover, the pH values, solvent effects and salt effects
fers to bind neutral or anionic guests. Liu and co-workers exert extraordinary influence over the binding ability of
have compared the complexation of organic dye cations the three kinds of macrocycles.
Supramolecular Chemistry  3

Table 1. Binding constants (Ka) and thermodynamic values for the complex formation between the macrocyclic hosts and the bicyclic
azoalkanes and ketone guests.

Host Guest Ka/(M–1)a ΔH°/(kJ mol–1) TΔS°/(kJ mol–1)


α-CD DBH – – –
DBO 57.8 ± 0.4 [50]b −13.4 ± 0.1 −3.4 ± 0.1
DBN – – –
β-CD DBH 241 ± 27 −14.6 ± 0.3 −1.0 ± 0.6
DBO 876 ± 10 −18.4 ± 0.1 −1.6 ± 0.1
DBN 2680 ± 10 −21.4 ± 0.1 −1.9 ± 0.1
γ-CD DBH – – –
DBO [6]b – –
DBN – – –
CX4c DBH 660 −23.3 −7.2
DBO 760 −25.4 −8.88
DBN 860 −26.4 −9.62
CX5 DBH 34.9 ± 2.1 −11.2 ± 0.6 −2.4 ± 0.9
DBO 24.8 ± 1.9 −10.3 ± 0.7 −2.4 ± 0.9
DBN 58.2 ± 2.2 −10.8 ± 0.3 −0.7 ± 0.8
CX6 DBH 55.2 ± 0.3 −15.3 ± 0.8 −5.4 ± 0.2
DBO 82.0 ± 0.1 −16.3 ± 0.4 −5.4 ± 0.4
Downloaded by [University of Sussex Library] at 07:44 05 December 2015

DBN 13.2 ± 0.6 −16.4 ± 0.6 −4.3 ± 0.1


Cyclopentanoned (3520 ± 490) [−27.4 ± 2.9] [−7.1 ± 2.5]
CB6 DBH [1300]e – –
DBO – – –
DBN – – –
(+)-Camphor – – –
CB7 Cyclopentanonef 3.30 × 105 −41 −9
DBH (3.81 ± 0.17) × 105 −38.7 ± 0.1 −6.9 ± 0.2
DBOf 1.34 × 107 −75 −34
DBN (2.08 ± 0.57) × 107 −60.2 ± 1.1 −22.7 ± 2.3
(+)-Camphor (1.94 ± 0.01) × 107 −89.9 ± 3.6 −48.2 ± 3.4
CB8 Cyclopentanonef 1100 −35 −18
DBH 2770 ± 160 −21.0 ± 1.0 −1.3 ± 1.1
DBOf 1.6 × 104 −37 −13
DBN (5.85 ± 0.20) × 105 −39.7 ± 0.3 −6.8 ± 0.3
(+)-Camphor (3.05 ± 0.27) × 107 −36.7 ± 1.9 6.0 ± 1.7
a
Mean values measured from at least three experiments at 25 °C in aqueous solution, pH 6.0; error given as standard deviation (±1σ)
b
Binding constants previously measured in D2O using 1H NMR titration: 50 M–1, cf. Ref. (25)
c
From a previous publication in aqueous solution at pH 6, cf. Ref. (16)
d
Titrations in water showed no detectable heat effect, value in square brackets measured in 200 mM Na2SO4; the Ka value was similar to the literature value of
3.4 × 103 M–1 in 200 mM Na2SO4, cf. Ref. (10)
e
In 200 mM Na2SO4 from Ref. (30)
f
At pH 6 from Ref. (20).

In our present investigation, we have selected a series of study, we have also included two ketones (cyclopentanone
bicyclic azoalkane homologues (Chart 2), namely 2,3-diaza- and (+)-camphor, Chart 2) as guests with cucurbiturils,
bicyclo[2.2.1]hept-2-ene (DBH), 2,3-diazabicyclo[2.2.2]oct- which meet the same criteria with respect to solubility and
2-ene (DBO) and 2,3-diazabicyclo[2.2.2]non-2-ene (DBN), neutrality, as well as non-protic and basic character as our
and investigated their binding to three different macrocy- bicyclic azoalkanes. The combined set of guest molecules,
cles (Chart 1) by ITC, NMR and molecular dynamics (MD) while matching the macrocyclic cavities with respect to
simulation. These azoalkanes, while neutral, are sufficiently their spherical shape, allowed us to explore the size selec-
water soluble (mM range, comparable to the investigated tivity of the macrocycles families.
macrocycles) to also allow for the assessment of weak
host–guest complexation phenomena (down to 10 M−1)
Results
(24, 25). Additionally, they are non-protic and weakly basic,
which precludes hydrogen bonding as the predominant We have structured our experimental study such that
interaction with the host molecules (e.g. CDs). Each of we first examined the binding affinities and thermody-
the macrocyclic host families by itself provides a distinct namic origins of the analytes to the macrocyclic hosts by
chemical environment for complexation with respect to ITC. Additionally, we performed 1H NMR investigations
shape, flexibility, hydrophobicity and high-energy water (facilitated by the millimolar solubilities of most hosts)
content. Therefore, these uncharged spherical guests are to confirm the ITC findings and characterise the mode of
ideally suited to study the inner pockets of the three host host–guest binding. Finally, packing coefficient (PC) cal-
families. To generalise some important conclusions of this culations, together with MD simulations, were carried out
4 D.-S. Guo et al.

to obtain further information on the inclusion complex interactions and the release of more high-energy water
formation between the macrocyclic hosts and the bicyclic molecules (6, 32). The comparably small entropic contribu-
azoalkanes. tions indicate that the resulting inclusion complexes are
not strongly co-conformationally restricted, as expected
from the spherical shape of the guests (31, 33, 34).
ITC measurements
The thermodynamic data and binding constants obtained Calixarenes
by ITC are summarised in Table 1, where some of the Calixarene macrocycles are very versatile hosts, since they
binding constants have been taken from previous inves- not only provide a variety of conformations in solution (26,
tigations done with either 1H NMR or ITC titrations. In all 27, 35, 36), but also change their charge status depend-
cases, the binding is an enthalpically driven process. Within ing on the working pH (37–39). In this respect, although
each macrocyclic family (CDs, CXs and CBs), the binding structurally similar, calixarene macrocycles may provide
constants of the azoalkanes have comparable orders of different environments for complexation, which can com-
magnitude. There is also an obvious trend between the plicate the interpretation of thermodynamic data. As can
differently sized guests: the enthalpy of the reaction (ΔH°), be seen from Table 1, the enthalpies of complexation in
as well as the binding constant (Ka) increases upon going the case of calixarene hosts do not follow any obvious
Downloaded by [University of Sussex Library] at 07:44 05 December 2015

from DBH (smallest), to DBO (intermediate), to DBN (largest trend, CX5 < CX6 < CX4, which can be attributed to the
guest). Among the cyclodextrin family, the β-CD macro- above-mentioned conformational diversity. The rigid cone
cycle is the prime candidate for azoalkane binding due conformation of the CX4 macrocycle (27, 28) provides the
to the favourable size-fit relationship. Among the calix- best shape-fit relationship for the bicyclic azoalkanes,
arene family, CX4 shows the highest binding affinities, with the highest complex stability for the DBN•CX4 case
presumably due to its pronounced structural rigidity (17, (ΔH° = –26.4 kJ mol–1). The pentameric homologue (CX5)
26–29). The most interesting cases, however, are those of has a larger and more flexible cone structure compared
the cucurbituril macrocycles CB7 and CB8, which show to the smaller CX4 one (37, 40), and therefore, its com-
contrasting enthalpic and entropic terms. CB6 presents plexation enthalpies with bicyclic azoalkanes are lower
an exception due to its small cavity size, that is it is only (ca. –10 kJ mol–1). Surprisingly, the largest CX6 homologue
capable of binding the smallest azoalkane, DBH, and the shows slightly higher enthalpies of binding compared to
smaller ketone, cyclopentanone. Our detailed results on CX5. One probable explanation is the ‘alternate’ confor-
the thermodynamics of binding are first outlined sepa- mation composed of two trimeric partial cones, which
rately for each type of macrocycle. is commonly assumed for the CX6 macrocycle (38, 39).
The ‘alternate’ conformation of the CX6 macrocycle can
Cyclodextrins possibly lead to 1:2 binding stoichiometries (41), which,
For cyclodextrins, only the medium sized β-CD shows sig- however, could not be experimentally differentiated in
nificant binding with all three bicyclic azoalkanes, while this study due to the low absolute binding affinities (from
the smaller and larger homologues, α-CD and γ-CD, show 55 to 132 M–1, extracted by assuming a 1:1 binding pat-
insignificant heat effects in our calorimetric experiments, tern). An additional complication lies in the competitive
with the exception of DBO with α-CD (Table 1). Previous 1H binding of Na+ as counterions, which modifies not only
NMR studies had also revealed the low binding constants the absolute binding constants of the organic guests but
of DBO to both, α-CD and γ-CD, with association constants also the measured thermodynamic parameters (42, 43). It
of 50 and 6 M–1, respectively (25, 31). The accessible ther- can furthermore be projected that the binding affinities of
modynamic data for complex formation between the bicy- Na+ with CX4, CX5 and CX6 will be significantly different.1
clic azoalkanes and the cyclodextrin macrocycles indicate As a result, the present thermodynamic data for CXs are
that the binding is enthalpically driven (ΔH° from −13.4 more difficult to rationalise due to conformational flex-
to −21.4 kJ mol−1), accompanied by slightly unfavourable ibility and also due to the competitive Na+ binding. The
entropy changes (TΔS° from −1.0 to −3.4 kJ mol−1). This latter applies particularly for guests with lower binding
pattern is similar to the one observed for the binding of lin- constants, where higher concentrations of CXs need to
ear aliphatic guests to CDs, with the complexation energy be used in the ITC reaction cell, resulting in a higher frac-
being proportional to the number of added methylene tion of CX•Na+ complex and directly related effects on the
groups. The enthalpic unit increment is ca. −3.3 kJ mol−1 apparent binding constant, as documented in detail by
per methylene group (1), going from −14.6 kJ mol−1 for the García-Río and co-workers (43, 44, 45).
DBH•β-CD complex, to −18.4 kJ mol−1 for the DBO•β-CD Keeping the complications in mind, the enthalpic con-
complex and −21.4 kJ mol−1 for the DBN•β-CD complex. tributions in calixarenes are mainly due to dispersion and
This increment goes in line with increasing dispersion C–H⋯π interactions; the release of high-energy water
Supramolecular Chemistry  5

molecules plays no major role, contributing to the compa- Table 2. Complexation-induced 1H NMR chemical shifts of the
rably small overall affinities (6, 16, 46). The negative entropy azoalkanes with 2 equiv. CB7, at pD 6.4.
can be attributed to the loss of degrees of freedom upon Δδ/ppma
complexation. The CX macrocycles show more unfavour- Hendo Hexo
able entropic changes (from –0.7 to –9.62 kJ mol–1) upon Azoalkanes [Hendo] [Hexo] Hsyn/Hanti
complexation than the cyclodextrin macrocycles, point- DBH −0.82 −0.90 −0.84/−0.96
ing to more pronounced conformational restrictions as DBO −0.78 −0.81 –
DBN −0.75 −0.78 –
imposed, for example, by the specific C–H⋯π interactions [−0.75] [−0.75]
with the bicyclic azoalkanes (15). a
NMR assignments according to Ref. (15); signals of the Hb protons overlap
with the CB7 resonances.
Cucurbiturils
Within the cucurbituril series, the CB6 homologue is rela-
tively small and it can only encapsulate cyclopentanone furthermore too poorly water soluble for systematic NMR
and DBH. It is interesting to note that the smaller cyclo- studies, we have concentrated our NMR investigations on
pentanone guest undergoes fast exchange with CB6 on the water-soluble CB7 macrocycle. The inclusion compl-
the NMR time scale, while the complexation with DBH exation with CB7 is strong, but in fast exchange with the
Downloaded by [University of Sussex Library] at 07:44 05 December 2015

undergoes slow exchange due to its large size (30). Both free host and guests on the NMR time scale. Addition of
CB7 and CB8, the larger homologues, display very strong 3 mM CB7 to 1.5 mM azoalkane solution at pD 6.4 (adjusted
binding affinities (from 103 to 107 M–1), but with distinct with NaOD) resulted in large 1H NMR upfield shifts for the
enthalpic and entropic effects: guest binding with CB7 azoalkane protons (up to 0.96 ppm, see Chart 2 and Table
is more strongly enthalpically driven, and there is a less 2), which are characteristics for the formation of inclu-
pronounced entropic loss in the case of CB8. The stronger sion complexes (50). The situation for the symmetric bar-
enthalpic response of CB7 can be traced back to the release rel-shaped CBs is different from that for the asymmetric
of high-energy water, which is more pronounced than for cone-shaped hosts CDs and CXs (15), where the penetra-
CB8 (20, 30, 47, 48); in short, although the CB8 cavity hosts tion depth varies with size of the host, reflected in pro-
more water molecules, they can form a better hydrogen nounced changes of the complexation-induced chemical
bonding network inside the larger cavity, which reduces shifts. Accordingly, the averaged complexation-induced
their net high-energy water content. The less pronounced shifts of all three azoalkanes are rather constant, due to
entropy loss for the formation of the CB8 complexes can, their central co-conformation, with only a small size-­
in turn, be intuitively interpreted in terms of the better dependent decrease from DBH (ca. –0.87 ppm) to DBO
mobility of the guests inside the larger cavity. It is interest- (ca. –0.80 ppm), and DBN (ca. –0.76 ppm). In the case of the
ing to note that the enthalpic contribution of CB8 binding CB7 NMR signals, the azoalkanes do not cause any split-
shows similar values for all guests (ca. –30 kJ mol−1); as a ting of the CB7 methylene signals, with the exception of
consequence, the entropic contributions govern the com- DBN, which induces a small but significant splitting of one
plexation strength. The complexation of (+)-camphor with doublet (data not shown). This suggests that the smaller
CB8 constitutes an exception, because its complexation is azoalkanes, DBH and DBO, do not assume any preferen-
also facilitated by an entropy gain, presumably due to the tial conformation on the NMR timescale, while the largest
entropy gained from desolvation of this putatively more and most tightly bound DBN (Ka = 2.08 × 107 M–1) has a
hydrophobic guest is not entirely counterbalanced by the preferred co-conformation inside CB7. The conformational
entropic loss due to conformational restrictions accompa- fixation is also consistent with the entropy loss being larg-
nying host–guest inclusion. Be this as it may, in terms of est for DBN (Table 1).
the macrocyclic environments, the CB hosts, led by CB7,
provide the strongest binding.
Theoretical calculations
MD simulations with Amber force field were performed
NMR measurements
to obtain detailed information on the inclusion complex
The interaction of the azoalkanes with the CBn macrocy- formation between the different macrocyclic hosts and
cles has been additionally confirmed by NMR measure- bicyclic azoalkane guests; they complement earlier calcu-
ments, whereas NMR studies for the CD and CX families lations and allow for a direct comparison (15, 34, 51). The
are already reported with same homologous guests (15, size selectivity of the azoalkanes towards the chosen hosts
49). For the smallest host, CB6, complexation with the large can be examined by evaluating the PC of their complexes,
guests (DBO, DBN, (+)-camphor) was unobservable either as it was previously successfully applied to predict the sta-
by ITC measurements or by NMR experiments. Since CB8 is bility of other CBn complexes (10). This analysis is possible
6 D.-S. Guo et al.

Figure 1. (Colour online) Side views of the distinct cavities of β-CD, CX4 and CB7; cavity volumes were calculated in analogy to Ref. (10).
Downloaded by [University of Sussex Library] at 07:44 05 December 2015

Figure 2. (Colour online) Top and side views of the MD-average structures of the studied complexes, (A) DBO•β-CD, (B) DBO•CX4 and (C)
DBO•CB7.

for the CD and CB macrocycles, which have geometrically rim (see Figure 1). In general, PCs in the range of 50–60%
well-defined cavities, see Figure 1. For the CX macrocy- provide the best size-fit relationship (9, 10, 52). As can be
cles, even for the CX4 homologue with the best defined seen from Table 3, the PCs of the host–guest complexes
cavity, no definitive cavity volume, and consequently no scale nicely with the experimental binding constants
PC values, could be defined due to the wide-open upper (Table 1). For example, they rationalise the strong binding
Supramolecular Chemistry  7

Table 3. Guest volumes, inner host cavity volumes and associated consistent with the 1H NMR shifts of the guest protons,
PCs with CBsa, β-CD and CX4. e.g. for CB7 (see above). The computed and experimental
PC/% structures reveal the favourable hydrophobic interactions
Guest/host CB6 CB7 CB8 β-CD CX4 between the studied guest molecules and the hydropho-
[volume/Å3] [142] [242] [367] [277] [>102]b bic cavities of both, CB7 and β-CD. The average structure
Cyclopentanone [87] 61 36 24 of the CX4 complex (Figure 2) shows that the guest is
DBH [96] 68 40 26 35 94
DBO [111] 78 46 30 40 109 also included within the cavity, with the azo group (N=N)
DBN [126] 89 52 34 46 124 exposed to the aqueous bulk and the C–H bonds suitably
(+)-Camphor [160] 113 66 44
a
positioned for C–H⋯π interactions; this co-conformation
The data of CBs with azoalkanes are from a previous publication, cf. Ref. (53).
b
The selected cavities are shown in Figure 1; for CX4, only a lower limit could is consistent with previous 2D NMR results (15).
be estimated. The guests’ dynamics inside the host cavity, monitored
by the superposition of six snapshots extracted from the
constant for DBN•CB7 (52%, virtually ideal PC value), and MD trajectories for each complex, are shown in Figure 3.
also account for the lower binding of (+)-camphor to CB7 Rotational motion of the guest molecules inside the cavity
(66%, PC above ideal value). was observed in all complexes, while a restricted transla-
tional motion was most pronounced for the β-CD com-
Downloaded by [University of Sussex Library] at 07:44 05 December 2015

The ‘preferred capacity’ of a concave host can be cal-


culated as 55% of its cavity volume. Accordingly, the pre- plexes. To point to one peculiarity, for DBH and DBO, but
ferred capacity of CB7 is 133 Å3 and that of β-CD is 152 Å3, not for the larger DBN, a flip-flop motion was observed
while that of CX4 has been previously estimated to be inside CB7, facilitated by the looser packing.
130 Å3 (15). Accordingly, these three hosts are comparable We have also looked at the dynamics of the free macro-
in terms of their capacity – as are the smaller and larger cycles (without guests) in explicit water molecules, which
homologues – which justifies the direct comparison con- in fact increases in the sequence of CB7 < β-CD ≈ CX4
ducted in this study. (Figures S3 and S4). It was also observed that, among the
The average structures of the 1:1 complexes of DBO macrocyclic hosts, β-CD showed the highest flexibility after
with the different hosts were obtained from 5-ns MD tra- binding with the neutral azoalkane guests, followed by
jectories, see Figure 2. For β-CD and CB7, the guest mol- CX4, while CB7 formed, as expected, the most rigid host–
ecule is included within the hydrophobic cavity, which is guest complex.

Figure 3. (Colour online) Dynamics of the host–guest complexes shown as clustered molecular display for (A) β-CD, (B) CX4 and (C) CB7
with the homologous guests DBH, DBO and DBN (from left to right). For the hosts, only the average structure is shown for clarity. The
drawings include six representative structures that refer to the time interval of simulation from 0 to 5 ns in 1-ns step.
8 D.-S. Guo et al.

Discussion of forming hydrogen bonds, particularly with


anionic guest molecules, while the CXs donate
CDs, CXs and CBs are three classical water-soluble macro- synergistically negative charge and π-stacking
cycles widely studied with respect to their ability to form interactions for cation binding; CBs offer ion–
host–guest complexes in aqueous media (23, 54–56). CDs dipole interactions with their carbonyl portals,
are cyclic oligomers of anhydroglucopyranose linked by which also promote cation binding (23, 54, 56).
α-1,4 glycosidic bonds. This geometry gives CDs the overall In our study, we selected neutral guests with
shape of a truncated cone or torus rather than that of a poorly basic and hydrogen bonding properties,
perfect cylinder, with the wider side formed by the second- precisely with the aim to bypass the differen-
ary 2- and 3-hydroxyl groups and the narrower side by the tial specific interactions provided by the rims/
primary 6-hydroxyl groups (1, 57). The calixarene skeleton portals, and to focus on their hydrophobic cavi-
ty-driven binding.
is built up by the base-catalysed condensation of 4-substi-
(4) The size/shape-fit always plays a crucial role in
tuted phenols with formaldehyde, with a conformational
governing the binding selectivity for each mac-
variability much larger than CDs (58). While the smaller rocycle, and it depends on its geometry. CBs
homologues CX4 and CX5 are typically represented as a have tighter portals (their diameter is smaller
(flexible) truncated cone or cup, with the upper rim formed than that of the equator (30)), which leads to a
Downloaded by [University of Sussex Library] at 07:44 05 December 2015

by the sulfonate groups and the lower rim by the phenolic more stringent size/shape selectivity. CXs have
hydroxyl groups, the larger CX6 usually adopts an ‘alter- a wide and loose opening, which allows a vari-
nate’ conformation (41, 59–61). CBs are cyclic oligomers ety of guest molecules to be encapsulated with
composed of glycoluril units joined by dual methylene lower selectivity. Taking the example of CX4
bridges, possessing a rigid symmetric barrel or pumpkin and CB7, with comparable preferred capacity
shape (2, 62). All three classes of macrocycles have demon- (ca. 130 Å3, see Results) (15), CB7 shows a high
strated significant biocompatibility and show therefore for affinity to methyl viologen (Ka = ~105 M–1) but
no appreciable binding to lucigenin (69–71),
several biological applications involving molecular recog-
whereas CX4 shows strong binding to both
nition events (56, 63–68). Although the formation of host–
methyl viologen (Ka = ~106 M–1) and lucigenin
guest inclusion complexes is invariably facilitated by their (Ka = ~107 M–1) (16, 40, 72, 73).
hydrophobic inner cavities, the intrinsic characteristics and (5) CDs and CBs can form deep inclusion com-
inclusion properties of the three hosts differ significantly, plexes, forming threaded structures in most
which forms the basis for our present comparative study. cases, while CXs favour a more shallow inclu-
In particular, the following aspects require consideration: sion (64, 74–77). This is because the hydrogen
bonding network formed by the lower rim phe-
(1) The CB cavity is equatorially symmetric, while
nolic hydroxyls leaves the bottom of CX cavity
the CD and CX ones are asymmetric. Both por-
effectively closed, preventing guest threading.
tals/rim of CBs are identical, whereas CDs and
The shape of CXs is therefore better described
CXs are distinguished by a wider side (upper
as that of a bowl or cup, with a wide opening
rim) and a narrower one (lower rim).
and closed bottom, than that of a truncated
(2) Among the three macrocycle families investi-
cone.
gated, the CXs are the most conformationally
variable, with conformations other but the
desirable cone shape possible, including par- To turn to the comparative binding of the three macr-
tial-cone, 1,2-alternate, 1,3-alternate and more ocycles with neutral guests, as can be seen from Table 1,
complicated ones for larger analogues. On the CXs exhibit the lowest binding selectivity for the neutral
contrary, CDs and CBs have robust conforma- guests employed, and also a low host selectivity when the
tions with a better defined inner cavity (3, 10). same guest is compared. With respect to guest selectivity,
While maintaining an overall cone shape, CDs the Ka values are at best double upon going from DBH to
are, however, considerably more flexible than DBN, and with respect to host selectivity, the Ka values
CBs (Figures S3 and S4). of the calixarene hosts do not follow a systematic trend,
(3) The hydrophobic effect is consistently a driv- i.e. CX5 < CX6 < CX4, a fact that can be attributed to the
ing force for all of these three water-soluble
previously mentioned competitive Na+ binding and con-
macrocycle families. In particular, the release
formational diversity of CXs in general, and CX6 in particu-
of high-energy water molecules from the
different macrocyclic cavities (non-classical lar. Even when CX4 and CX5 are compared, the two hosts
hydrophobic effect) plays an important role, with a cup-shaped conformation, the best host selectiv-
most pronounced for the medium sized CB7 ity is only a factor of 30, for DBO as guest, one order of
and least pronounced for the smallest CX4 (6). magnitude less than that of CBs. Moreover, although CX5
Additionally, the portals of CDs are capable possesses a larger cavity size, it does not provide a higher
Supramolecular Chemistry  9

binding affinity to the larger guest as intuitively expected. recently, we reported water-box simulations for discrete
This is because the CX5 cavity resembles more that of a water molecules inside cavities and concluded that if N
shallow dish, due to the restriction of the lower rim hydro- is the number of water molecules in a cavity and m is the
gen bond network (40, 78), its cavity opening is too wide to average number of hydrogen bonds which these water
effectively accommodate hydrophobic cargoes. The poor molecules form inside the cavity, then the relative driving
binding selectivity of CXs to the neutral azoalkane guests force can be estimated as Z = N (3.62-m), where 3.62 is
further validates the aforementioned points 2 and 4: the the simulated hydrogen bond number in bulk water (6).
preorganisation of the conformationally flexible CXs is not β-CD hosts 4.4 water molecules in its cavity with m = 2.96
as well developed as for the other two macrocycles, result- and, thus, Z = 3.1; CX4 has 0.8 water molecules in its cav-
ing in a reduced size/shape selectivity towards the guests. ity with m = 2.15 and, thus, Z = 1.2; CB7 accommodates
The differential binding affinities of β-CD towards DBH, 7.9 water molecules in its cavity with m = 2.52 and, thus,
DBO and DBN suggest an intermediary guest selectivity. Z = 8.7. These estimates are in good agreement with the
The highest selectivity, both upon varying the guest size experimental results. For example, the cavity of CX4, also
and upon varying the host cavity size, is observed for termed ‘the smallest cup of water’ (79), is occupied by a
the CBs, for example Ka DBN•CB8/Ka DBH•CB8 = 211, Ka DBH•CB7/ single water molecule, such that the release of high-energy
Ka DBH•CB6 = 293 and Ka DBO•CB7/Ka DBO•CB8 = 323. Consequently, water (Z = 1.2) cannot contribute much to the net driving
Downloaded by [University of Sussex Library] at 07:44 05 December 2015

defying the activity-selectivity principle, the CB macrocy- force for neutral guest binding. In contrast, neutral guest
cles display not only the strongest absolute binding to the binding to CB7 shows a strong high-energy water-release
neutral guests but also the highest selectivity towards the effect Z = 8.7, which accounts for the strongest binding
guests. We therefore expanded our investigation for CBs by this host. β-CD, with Z = 3.1, lies in between in terms
to two additional guests: cyclopentanone (as a smallest of water-release contributions. The same arguments have
guest) and (+)-camphor (as a larger guest). As expected, been previously used to assess the trends within the CB
the smaller cyclopentanone forms a tighter complex with series; for CB6, the number N is smaller than for CB7,
CB6 than DBH (more ideal PC, 61% vs. 68 %), and the larger while for CB8, the number m is higher than for CB7, both
(+)-camphor forms a tighter complex with CB8 than DBN; of which result in reduced Z values and hence binding.
both results follow a simple supramolecular structure– This delicate interplay between the individual energetic
activity relationship, that is the binding strength increases frustration and the absolute number of the cavity water
as packing becomes more ideal (see Results, Table 3). molecules provides an explanation for the medium sized
β-CD, CX4 and CB7 are three analogues studied most macrocycle CB7 displaying the highest binding constants,
frequently in tandem, because of their similar preferred not only for the presently selected series of neutral guests,
capacity (see Results); generally, these three also stand but also among synthetic macrocycles in general (20).
out within each class as highest affinity binders (15). β-CD
shows a moderate binding to the azoalkanes, similar to
Conclusion
CX4, but its complexation enthalpy and entropy terms are
distinct. The enthalpy gain of CX4 is more favourable than We have evaluated the binding affinities and the thermo-
that of β-CD, probably resulting from the unique C-H⋯π dynamic parameters for the binding of three azoalkane
interactions offered by the aromatic framework of calix- guests, namely DBH, DBO and DBN, with three different
arenes that the other hosts are lacking. Compared with prototypal macrocyclic host families, CDs, CXs and CBs. All
CDs, CXs have a wider opening, which leads to the rela- offer a hydrophobic inner cavity, but differ in shape, flex-
tively shallower inclusion that is unfavourable for compl- ibility and hydrophobicity, resulting in different selectivi-
exation-induced desolvation (Figures 2 and 3). This is also ties and affinities. Neutral azoalkane guests are perfectly
reflected in MD calculations (see hydrogen bond analysis, suited to comparatively study these three host families,
Table S1), which reveal that CBs and CDs shield the guest because direct host–guest interactions, namely, hydrogen
molecules better from the surrounding water. Moreover, bonding in CDs, charge interactions in CXs and ion–dipole
CXs, inherent to their higher flexibility, always undergo a ones in CBs, can be neglected, in a first approximation.
comparably larger loss of the conformational degrees of The obtained results show that (1) the preorganisation of
freedom upon complexation of guests (17). conformational flexibility governs the binding selectivity
Our study exposes that CB7, despite its comparable significantly. The well-defined CB cavity provides a much
cavity size, forms 2–4 orders of magnitude more stable stringent size/shape-fit than the conformationally flexible
complexes with neutral guests than β-CD and CX4. The CX cavity. (2) CXs have a wider opening and afford rather
main factor that drives the encapsulation of neutral guests a shallow inclusion rather than a deep encapsulation or
into macrocyclic cavities is the release of high-energy even threading; the latter is observed for CDs and CBs.
water, that is the non-classical hydrophobic effect. Very (3) It is the non-classical hydrophobic effect (release of
10 D.-S. Guo et al.

high-energy water) that gives CBs the luxury of showing NMR measurements
superior binding affinities towards neutral guests. NMR experiments were performed at ambient tempera-
ture in D2O (99.8%). The pD values of the solutions were
adjusted by the addition of D2SO4 or NaOD. pH readings
Experimental were taken with a WTW 330i pH meter equipped with a
Materials combined pH glass electrode (SenTix Mic) and converted
to pD (+0.40 units) (83). 1H NMR spectra (400 MHz) were
The bicyclic azoalkane homologues (DBH, DBO and DBN)
recorded using the chemical shift of HOD in D2O preset
were obtained as previously reported (31). α-, β- and
at 4.67 ppm as reference. Concentrations of CB7 (3.0 mM)
γ-cyclodextrin (>95% purity) were purchased from Fluka
and azoalkane (1.5 mM) were chosen to obtain a sufficient
(Seelze, Germany) and used without purification. The calix-
signal strength (Figure S2).
arene homologues (CX4, CX5 and CX6) were synthesised
and purified according to the literature procedures (37, 39). Computational studies
The CB6 sample was purchased from Merck (Darmstadt, The initial molecular geometry of CB7, β-CD and CX4 was
Germany) and used without any further purification. In obtained from X-ray diffraction data, whereas the starting
the calculations of the cucurbit[6]uril concentration, geometries of the guest molecules were built up from
Downloaded by [University of Sussex Library] at 07:44 05 December 2015

20% water content was taken into account, which was standard lengths and bond angles. The AMBER 8 program
determined by 1H NMR. CB7 (30% water content) was was used throughout this work using the param99 (for β-
synthesised according to the literature procedure (80). CD) and the general force field parameters sets (for CB7, CX4
CB8 (20% water content) was purchased from Sigma- and the guest molecules) (84). The electrostatic potentials
Aldrich (Taufkirchen, Germany). All solutions were freshly of the guest molecules were taken from ab initio HF/6–31G*
prepared on a daily basis in ultrapure water, and the pH/ calculations performed within the Gaussian 03 W package
pD was adjusted using HCl and NaOH. Care was taken to (85); atomic charges reproducing these electrostatic poten-
minimise the amount of NaOH used as excessive Na+ may tials were obtained by means of the RESP methodology. All
also complex with CB7 and affect its binding towards the systems were solvated in a truncated octahedral periodic
investigated guest (30, 81). box of TIP3P water molecules with a closeness parameter of
15 Å. Periodic boundary conditions were adopted and the
Instruments and measurements particle-mesh Ewald method was used for the treatment of
long-range electrostatic interaction. The non-bonded cut-
Isothermal titration calorimetry off was set to 10.0 Å. Energy minimisation was performed
The microcalorimetric measurements were performed on for each solvated complex using the conjugate gradient
an isothermal titration calorimeter (VP-ITC), purchased algorithm, heated up to 298 K for 60 ps, followed by 200 ps
from MicroCal Inc., USA. The ITC instrument was peri- for achieving equilibration at 298 K and 1 atm. Production
odically calibrated using the internal electric heater (82) runs were carried out for 5 ns at 298 K and a barostat at
Titration experiments were carried out at 25 °C in aqueous 1.0 atm (under an NPT ensemble). A 2-fs time step with a
solution, pH 6; each experiment consisted of 25–30 con- saving of the structure every 2 ps was used, and the non-
secutive injections (5–10 μL) of guest solution (0.2–75 mM) bonded pair list was updated every 25 steps.
injected into the microcalorimetric reaction cell, filled with
macromolecule solution (0.03–5.00 mM) (Figure S1). All
solutions were degassed prior to the titration. Heats of Supplemental material
dilution were subtracted from each data set. The data Supplemental data for this article can be accessed here:
were analysed according to a 1:1 binding model in Origin http://dx.doi.org/10.1080/10610278.2015.1105374
7.0 software; this affords simultaneously the binding sto-
ichiometry (N), the complex stability constant (Ka), the
standard molar reaction enthalpy (∆H°) and the standard
Note
deviation from the titration curve. The knowledge of com- 1. 
We have previously measured the binding constants
plex stability constant (Ka) and molar reaction enthalpy of Na+ with CX4 and CX5 by fluorescence competitive
titrations, giving the values of 85 and 173 M−1 (please
(∆H°) enabled the calculation of standard free energy (∆G°)
see Supporting Information of Ref. 73).
and entropy changes (∆S°) according to
ΔG◦ = −RT lnKa = ΔH ◦ − T ΔS◦
Acknowledgements
where R is the gas constant and T is the absolute The Chinese team gratefully acknowledges the support of the
temperature. NNSFC (21172119) and 21322207). The German team thanks
Supramolecular Chemistry  11

the Deutsche Forschungsgemeinschaft (DFG grant NA-686/5 (21) Stancu, A.D.; Buschmann, H.-J.; Mutihac, L. J. Incl. Phenom.
and SPP 1807), the COST Action CM1005 ‘Supramolecular Macrocycl. Chem. 2013, 75, 1–10.
Chemistry in Water’, and the Fonds der Chemischen Industrie (22) Pessêgo, M.; Basilio, N.; Moreira, J.A.; García-Río, L.
for support. Both teams thank the Robert Bosch Foundation for ChemPhysChem 2011, 12, 1342–1350.
bilateral collaborative project support (Wissenschaftsbrücke (23) Liu, Y.; Li, C.-J.; Guo, D.-S.; Pan, Z.-H.; Li, Z. Supramol. Chem.
China). D.-S. Guo also appreciates the DAAD for supporting his 2007, 19, 517–523.
research visit at Jacobs University Bremen. (24) Bakirci, H.; Koner, A.L.; Schwarzlose, T.; Nau, W.M. Chem.
Eur. J. 2006, 12, 4799–4807.
(25) Nau, W.M.; Zhang, X.Y. J. Am. Chem. Soc. 1999, 121, 8022–
Disclosure statement 8032.
(26) Shinkai, S.; Araki, K.; Kubota, M.; Arimura, T.; Matsuda, T. J.
No potential conflict of interest was reported by the authors.
Org. Chem. 1991, 56, 295–300.
(27) Bonal, C.; Israëli, Y.; Morel, J.P.; Morel-Desrosiers, N. J.
Funding Chem. Soc., Perkin Trans. 2001, 2, 1075–1078.
(28) Israeli, Y.; Detellier, C. Phys. Chem. Chem. Phys. 2004, 6,
This work was supported by the NNSFC [21172119], [21322207]; 1253–1257.
Deutsche Forschungsgemeinschaft DFG [grant NA-686/5], [SPP (29) Guo, D.-S.; Zhang, H.-Q.; Ding, F.; Liu, Y. Org. Biomol. Chem.
1807]; Fonds der Chemischen Industrie; Robert Bosch Founda- 2012, 10, 1527–1536.
Downloaded by [University of Sussex Library] at 07:44 05 December 2015

tion; DAAD. (30) Márquez, C.; Hudgins, R.R.; Nau, W.M. J. Am. Chem. Soc.
2004, 126, 5806–5816.
(31) Bakirci, H.; Zhang, X.Y.; Nau, W.M. J. Org. Chem. 2005, 70,
References 39–46.
(32) Schmidtchen, F.P. Chem. Eur. J. 2002, 8, 3522–3529.
(1) Rekharsky, M.V.; Inoue, Y. Chem. Rev. 1998, 98, 1875–1918.
(33) Rekharsky, M.V.; Inoue, Y. J. Am. Chem. Soc. 2002, 124,
(2) Lagona, J.; Mukhopadhyay, P.; Chakrabarti, S.; Isaacs, L.
12361–12371.
Angew. Chem. Int. Ed. 2005, 44, 4844–4870.
(34) Mayer, B.; Zhang, X.Y.; Nau, W.M.; Marconi, G. J. Am. Chem.
(3) Guo, D.-S.; Wang, K.; Liu, Y. J. Incl. Phenom. Macrocycl.
Soc. 2001, 123, 5240–5248.
Chem. 2008, 62, 1–21.
(35) Gutsche, C.D.; Bauer, L.J. J. Am. Chem. Soc. 1985, 107,
(4) Norouzy, A.; Azizi, Z.; Nau, W.M. Angew. Chem. Int. Ed.
6052–6059.
2015, 54, 792–795.
(36) Rogers, J.S.; Gutsche, C.D. J. Org. Chem. 1992, 57, 3152–
(5) Schneider, H.J. Angew. Chem. Int. Ed. 2009, 48, 3924–3977.
3159.
(6) Biedermann, F.; Nau, W.M.; Schneider, H.J. Angew. Chem.
(37) Steed, J.W.; Johnson, C.P.; Barnes, C.L.; Juneja, R.K.;
Int. Ed. 2014, 53, 11158–11171.
Atwood, J.L.; Reilly, S.; Hollis, R.L.; Smith, P.H.; Clark, D.L. J.
(7) Liu, L.; Guo, Q.-X. J. Incl. Phenom. Macrocycl. Chem. 2002,
Am. Chem. Soc. 1995, 117, 11426–11433.
42, 1–14.
(38) Atwood, J.L.; Clark, D.L.; Juneja, R.K.; Orr, G.W.; Robinson,
(8) Praetorius, A.; Bailey, D.M.; Schwarzlose, T.; Nau, W.M. Org.
K.D.; Vincent, R.L. J. Am. Chem. Soc. 1992, 114, 7558–7559.
Lett. 2008, 10, 4089–4092.
(39) Arena, G.; Contino, A.; Lombardo, G.G.; Sciotto, D.
(9) Florea, M.; Nau, W.M. Angew. Chem. Int. Ed. 2011, 50,
Thermochim. Acta 1995, 264, 1–11.
9338–9342.
(40) Guo, D.-S.; Wang, L.-H.; Liu, Y. J. Org. Chem. 2007, 72, 7775–
(10) Nau, W.M.; Florea, M.; Assaf, K.I. Isr. J. Chem. 2011, 51,
7778.
559–577.
(41) Liu, Y.; Li, Q.; Guo, D.-S.; Chen, K. Cryst. Growth Des. 2007,
(11) Rekharsky, M.V.; Schwarz, F.P.; Tewari, Y.B.; Goldberg, R.N. J.
7, 1672–1675.
Phys. Chem. 1994, 98, 10282–10288.
(42) Bakirci, H.; Koner, A.L.; Nau, W.M. Chem. Commun. 2005,
(12) Liu, Y.; Yang, E.C.; Yang, Y.W.; Zhang, H.Y.; Fan, Z.; Ding, F.;
5411–5413.
Cao, R. J. Org. Chem. 2004, 69, 173–180.
(43) Francisco, V.; Piñeiro, A.; Nau, W.M.; García-Río, L. Chem.
(13) Liu, Y.; Ma, Y.-H.; Chen, Y.; Guo, D.-S.; Li, Q. J. Org. Chem.
Eur. J. 2013, 19, 17809–17820.
2006, 71, 6468–6473.
(44) Basilio, N.; García-Río, L.; Martín-Pastor, M. J. Phys. Chem. B
(14) Kon, N.; Iki, N.; Miyano, S. Org. Biomol. Chem. 2003, 1,
2010, 114, 7201–7206.
751–755.
(45) Francisco, V.; Basilio, N.; García-Río, L. J. Phys. Chem. B
(15) Bakirci, H.; Koner, A.L.; Nau, W.M. J. Org. Chem. 2005, 70,
2012, 116, 5308–5315.
9960–9966.
(46) Fucke, K.; Anderson, K.M.; Filby, M.H.; Henry, M.; Wright,
(16) Cui, J.; Uzunova, V.D.; Guo, D.-S.; Wang, K.; Nau, W.M.; Liu, Y.
J.; Mason, S.A.; Gutmann, M.J.; Barbour, L.J.; Oliver, C.L.;
Eur. J. Org. Chem. 2010, 1704–1710.
Coleman, A.W.; Atwood, J.L.; Howard, J.A.K.; Steed, J.W.
(17) Liu, Y.; Guo, D.-S.; Zhang, H.-Y.; Ma, Y.-H.; Yang, E.-C. J. Phys.
Chem. Eur. J. 2011, 17, 10259–10271.
Chem. B 2006, 110, 3428–3434.
(47) Buschmann, H.J.; Jansen, K.; Schollmeyer, E. Thermochim.
(18) Rekharsky, M.V.; Mori, T.; Yang, C.; Ko, Y.H.; Selvapalam, N.;
Acta 1998, 317, 95–98.
Kim, H.; Sobransingh, D.; Kaifer, A.E.; Liu, S.M.; Isaacs, L.;
(48) Biedermann, F.; Vendruscolo, M.; Scherman, O.A.; De
Chen, W.; Moghaddam, S.; Gilson, M.K.; Kim, K.M.; Inoue, Y.
Simone, A.; Nau, W.M. J. Am. Chem. Soc. 2013, 135, 14879–
Proc. Nat. Acad. Sci. USA 2007, 104, 20737–20742.
14888.
(19) Assaf, K.I.; Nau, W.M. Supramol. Chem. 2014, 26, 657–669.
(49) Zhang, X.Y.; Gramlich, G.; Wang, X.J.; Nau, W.M. J. Am.
(20) Biedermann, F.; Uzunova, V.D.; Scherman, O.A.; Nau, W.M.;
Chem. Soc. 2002, 124, 254–263.
De Simone, A. J. Am. Chem. Soc. 2012, 134, 15318–15323.
12 D.-S. Guo et al.

(50) Mock, W.L.; Shih, N.-Y. J. Org. Chem. 1986, 51, 4440–4446. (75) Biedermann, F.; Elmalem, E.; Ghosh, I.; Nau, W.M.;
(51) Zhang, X.Y.; Nau, W.M. Angew. Chem. Int. Ed. 2000, 39, Scherman, O.A. Angew. Chem. Int. Ed. 2012, 51, 7739–
544–547. 7743.
(52) Mecozzi, S.; Rebek, J., Jr. Chem. Eur. J. 1998, 4, 1016–1022. (76) Liu, K.; Liu, Y.; Yao, Y.; Yuan, H.; Wang, S.; Wang, Z.; Zhang, X.
(53) Lee, T.-C.; Kalenius, E.; Lazar, A.I.; Assaf, K.I.; Kuhnert, N.; Angew. Chem. Int. Ed. 2013, 52, 8285–8289.
Grün, C.H.; Jänis, J.; Scherman, O.A.; Nau, W.M. Nat. Chem. (77) Jiang, L.X.; Deng, M.L.; Wang, Y.L.; Liang, D.H.; Yan, Y.;
2013, 5, 376–382. Huang, J.B. J. Phys. Chem. B 2009, 113, 7498–7504.
(54) Dsouza, R.N.; Pischel, U.; Nau, W.M. Chem. Rev. 2011, 111, (78) Coruzzi, M.; Andreetti, G.D.; Bocchi, V.; Pochini, A.; Ungaro,
7941–7980. R. J. Chem. Soc., Perkin Trans. 1982, 2, 1133–1138.
(55) Márquez, C.; Nau, W.M. Angew. Chem. Int. Ed. 2001, 40, (79) Hontama, N.; Inokuchi, Y.; Ebata, T.; Dedonder-Lardeux,
4387–4390. C.; Jouvet, C.; Xantheas, S.S. J. Phys. Chem. A 2010, 114,
(56) Ghosh, I.; Nau, W.M. Adv. Drug Delivery Rev. 2012, 64, 764– 2967–2972.
783. (80) Márquez, C.; Huang, F.; Nau, W.M. IEEE Trans. Nanobiosci.
(57) Connors, K.A. Chem. Rev. 1997, 97, 1325–1358. 2004, 3, 39–45.
(58) Böhmer, V. Angew. Chem. Int. Ed. 1995, 34, 713–745. (81) Hennig, A.; Bakirci, H.; Nau, W.M. Nat. Methods 2007, 4,
(59) Dalgarno, S.J.; Hardie, M.J.; Makha, M.; Raston, C.L. Chem. 629–632.
Eur. J. 2003, 9, 2834–2839. (82) Wiseman, T.; Williston, S.; Brandts, J.F.; Lin, L.N. Anal.
(60) Atwood, J.L.; Barbour, L.J.; Hardie, M.J.; Raston, C.L. Coord. Biochem. 1989, 179, 131–137.
Downloaded by [University of Sussex Library] at 07:44 05 December 2015

Chem. Rev. 2001, 222, 3–32. (83) Glasoe, P.K.; Long, F.A. J. Phys. Chem. 1960, 64, 188–190.
(61) Liu, Y.; Li, Q.; Guo, D.-S.; Chen, K. CrystEngComm 2008, 10, (84) Case, D.A.; Darden, T.A.; Cheatham, T.E.; Simmerling,
675–680. C.L.; Wang, J.; Duke, R.E.; Luo, R.L.; Merz, K.M.; Wang, B.;
(62) Lee, J.W.; Samal, S.; Selvapalam, N.; Kim, H.J.; Kim, K. Acc. Pearlman, D.A.; Crowley, M.; Brozell, S.; Tsui, V.; Gohlke, H.;
Chem. Res. 2003, 36, 621–630. Mongan, J.; Hornak, V.; Cui, G.; Beroza, P.; Schafmerister, C.;
(63) Ghale, G.; Nau, W.M. Acc. Chem. Res. 2014, 47, 2150–2159. Caldwell, J.W.; Ross, W.S.; Kollman, P.A. AMBER8; University
(64) Guo, D.-S.; Liu, Y. Acc. Chem. Res. 2014, 47, 1925–1934. of California: San Francisco, CA, 2004.
(65) Perret, F.; Lazar, A.N.; Coleman, A.W. Chem. Commun. (85) Frisch, M.; Trucks, G.; Schlegel, H.; Scuseria, G.; Robb, M.;
2006, 2425–2438. Cheeseman, J.; Montagomery, J.; Verven, J.R.; Kudin, K.;
(66) Stella, V.J.; Rajewski, R.A. Pharm. Res. 1997, 14, 556–567. Burant, J.; Millam, J.M.; Iyenger, S.; Tomasi, J.; Barone, V.;
(67) Uekama, K.; Hirayama, F.; Irie, T. Chem. Rev. 1998, 98, Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson,
2045–2076. G.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda,
(68) Assaf, K.I.; Nau, W.M. Chem. Soc. Rev. 2015, 44, 394–418. R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao,
(69) Kim, H.J.; Jeon, W.S.; Ko, Y.H.; Kim, K. Proc. Nat. Acad. Sci. O.; Nakai, H.; Klene, M.; Li, X.; Konx, J.; Hratchian, H.; Cross,
USA 2002, 99, 5007–5011. J.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Strat-
(70) Ong, W.; Gómez-Kaifer, M.; Kaifer, A.E. Org. Lett. 2002, 4, mann, R.; Yazyev, O.; Austin, A.; Cammi. R.; Pomelli, C.;
1791–1794. Ochterski, J.; Ayala, P.; Moromuka, K.; Voth, G.; Salvador,
(71) Masson, E.; Shaker, Y.M.; Masson, J.-P.; Kordesch, M.E.; P.; Dannenberg, J.; Zakrzewski, V.; Dapprich, S.; Daniels, A.;
Yuwono, C. Org. Lett. 2011, 13, 3872–3875. Strain, M.; Farkas, O.; Malick, D.; Rabuck, A.; Raghavacha-
(72) Zhao, H.-X.; Guo, D.-S.; Liu, Y. J. Phys. Chem. B 2013, 117, ri, K.; Foresman, J.; Ortiz, J.; Cui, Q.; Baboul, A.; Clifford, S.;
1978–1987. Cioslwski, J.; Setvanov, B.; Liu, G.; Liashenko, A.; Piskorz, P.;
(73) Guo, D.-S.; Uzunova, V.D.; Su, X.; Liu, Y.; Nau, W.M. Chem. Komaromi, I.; Marton, R.; Fox, D.; Keith, T.; Al-Laham, M.;
Sci. 2011, 2, 1722–1734. Peng, C.; Nanayakkara, A.; Challacombe, M.; Gill, P.; John-
(74) Basilio, N.; Francisco, V.; García-Río, L. Int. J. Mol. Sci. 2013, son, B.; Chen, W.; Wong, M.; Gonzalez, C.; Pople, J. Gaussi-
14, 3140–3157. an 03, Revision D.01; Gaussian, Inc.: Wallingford, CT, 2004.

You might also like