Sunteți pe pagina 1din 5

Computational and Theoretical Chemistry 987 (2012) 5761

Contents lists available at SciVerse ScienceDirect

Computational and Theoretical Chemistry


journal homepage: www.elsevier.com/locate/comptc

Adsorption complexes of copper and copper oxide in the deep eutectic solvent
2:1 ureacholine chloride
Jessica M. Rimsza b, L. Ren Corrales a,b,
a
b

Department of Chemistry and Biochemistry, The University of Arizona, Tucson, AZ 85721, United States
Department Materials Science and Engineering, The University of Arizona, Tucson, AZ 85721, United States

a r t i c l e

i n f o

Article history:
Received 26 May 2011
Accepted 1 November 2011
Available online 12 November 2011
Keywords:
Deep eutectic solvent
Density functional theory
Adsorption
Proton transfer

a b s t r a c t
Adsorption of metal oxide particulates by soft surface solvent action using the deep eutectic solvent (DES)
mixture of choline chloride and urea is examined at the molecular level. Quantum chemical calculations
are employed to determine the binding energy of neutral and anionic urea to copper oxide and elemental
copper, the main complexes believed to participate in surface cleaning of this DES. The possibility of the
existence of urea anions in this deep eutectic solvent is studied in terms of the possibility of proton transfer taking place. Results show that the anion formation is stabilized in a dielectric and that the relative
binding of neutral and anionic urea to copper oxide is similar whereas elemental Cu signicantly favors
binding with the urea anion. This work also shows that the hydrogen bond interactions of the urea with
the chloride ion maintains an open cluster structure as suggested by experiment and that proton transfer
can occur with increasing temperature or with metallic complex formation leading to more aggressive
solvent action.
2011 Elsevier B.V. All rights reserved.

1. Introduction
Deep eutectic solvents (DESs) are a unique class of multicomponent solvent systems with varying properties [1] that make them
useful in electrochemical processing of metal composites [2,3],
synthesis of framework metal oxide structures [4,5], biological
catalysis [6,7], formation of ionogels [8], lubrication [9], synthesis
of solar cells [10], and are under consideration for the soft surface
removal of contaminants generated during photoresist etching in
the back end of the line processing of semiconductors [11,12].
DES systems are comprised of mixed molecular components that
have low volatility and are liquids at room temperature. The latter
denitions satisfy two of three fundamental denitions of room
temperature ionic liquids (RTILs) [13]. A 2:1 eutectic mixture of
urea with choline chloride has the ionic components of the choline
cation and chloride anion while having hydrogen bonding of the
urea molecules. It shows ionic conductivity [3] satisfying a third
property of an ionic liquid, but only because of the choline cation
and chloride anion. DES also show a strong preference for metal
oxide adsorption [14,15]. The extent to which proton transfer
might enter into the possible chemistry of this DES in general is
also of interest because of their proximity in compositions to protic
ionic liquids [16] and because of other decompositions that are
possible [17]. In this particular system the only proton transfer that
Corresponding author at: Department of Chemistry and Biochemistry, The
University of Arizona, Tucson, AZ 85721, United States.
E-mail address: lrcorral@email.arizona.edu (L.R. Corrales).
2210-271X/$ - see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.comptc.2011.11.003

might occur is from urea transforming to a urea anion, as is found


by FAB-MS [18]. Such a possible state is interesting to explore because as shown below, the presence of the urea anion can transform the DES from a soft surface solvent to an aggressive one.
This work examines simple binding complexes that can form of
the copper oxide and elemental copper with urea, its anion, and
with the chloride ion.
One application under consideration for this DES is as a surface
contaminant cleaning agent to selectively remove, CuO for example, from a silicon dioxide surface. It is believed that the removal
of CuO from a silicon dioxide surface involves the formation of a
complex with urea anions. Subsequently, it is of interest to characterize both the degree of urea anion that can form in this DES, and
the binding strengths of urea and its corresponding anion with
CuO. The aggressive nature of this DES is characterized in terms
of proton dissociation as it can lead to the undesirable etching of
pure copper from a surface. This work, employs rst-principles
methods that provide signicant insight into observed behavior
of theses solvent systems as well as provide impetus for more theoretical and experimental characterization.
Objectives of this work are to characterize the molecular states
of the 2:1 ureacholine chloride DES to determine the degree of
urea anion formation and the binding strengths of simple molecular complexes that can form with CuO and Cu. This is done via a
computational approach that employs density functional theory
methods (DFT). The stability of molecular ions is considered by
approximating the bulk uid via embedding clusters within a

58

J.M. Rimsza, L.R. Corrales / Computational and Theoretical Chemistry 987 (2012) 5761

dielectric continuum [19]. The next section describes the approach,


followed by results then by a discussion and conclusions.
2. Methods
The CP2K code [20] that uses the Gaussian Plane Wave method
[21] was employed to carry out the cluster calculations using a cubic cell of length 14 . The molecular optimized Gaussian TZV2P
and DZVP basis set comprised of a triple(double)-zeta valence basis
set that includes two(one) set(s) of polarizable functions (d-type
and p-type) [22] along with the BLYP or PBE exchangecorrelation
function [23] with an energy cutoff of 280 Ry were used. Basis sets
for Cu were at the DZVP level in all calculations. In the cluster calculations of the copper complexes, the DZVP basis set was used
throughout to maintain consistency with the available copper basis
set for the CP2K code. Reported binding energies are the difference
between the energy of the complex minus the energy of the individual molecules that make up the complex. The energy for proton
transfer is taken to be the difference between the normal and the
proton transferred molecular congurations.

Fig. 1. Components of DES. (a) The neutral cluster (left) where both urea molecules
are neutral with a choline cation and a chloride ion. (b) The ion cluster (right) where
the left urea molecule is now an anion leading to the formation of an HCl. Note that
the urea anion packs closer to the choline cation. Relative energies of these states as
a function of dielectric is shown in Fig. 2.

3. Results
Energies of the optimized molecular structures are shown in
Table 1 for the two basis sets and two exchangecorrelation functions used in this work. Optimized structural results include that
urea takes on a C2 symmetry [19,24], and choline chloride takes
on a Cs symmetry [25] where geometries are similar to literature
results, see Table 3.
To examine the relative stability of the neutral molecules versus
the counterpart ionic molecules, a dielectric is introduced into the
cell via creating a spherical cavity in which the molecular cluster
sits and is surrounded by a dielectric eld, effectively producing
a continuum [26]. Fig. 1a and b shows the relative optimized structures of the neutral and ion pairs for a cluster representing the 2:1
ratio of urea to choline chloride. Forming the urea anion also requires forming HCl or a protonated urea (which is not stable in this
cluster) while maintaining the choline cation. The formation of the
HCl requires rotating the proton away from the urea molecules.
Most likely in a liquid state the proton would nd its way back
to the urea anion to reform the neutral molecule or proceed in proton transfer via the urea hydrogen bond network.
In Fig. 2a are plotted the results of optimizing the urea neutral
and anion clusters within a dielectric of K = 0, 15, 30, and 60 that
also compare three sets of calculations that include using the
TZV2P basis set within both the B3LYP and PBE exchangecorrelation functions as well as the DZVP basis set within the PBE exchangecorrelation function. The reference energy used here is
the sum of the molecular fragments at K = 1 that make up the neutral cluster at the same level of basis set and exchange function.
The energy required to form the urea anion is given by the differ-

Table 1
Absolute energies of the atomic and molecular components.
Atom/
Molecule
molecule charge

Cu
Cl
CuO
HCl
Urea
Urea
Choline

0
1
0
0
0
1
+1

Energy
(Hartrees)
DZVPBLYP

Energy
(Hartrees)
TZV2P
BLYP

47.702656
15.013270
15.041127
63.643547
15.520170
15.525772
43.862401
43.872887
43.315956
43.333740
62.280176
62.285862

Energy
(Hartrees)
DZVPPBE

Energy
(Hartrees)
TZV2PPBE

47.88624

15.061043
15.087469
63.838422
15.568315
15.574242
43.957053
43.966119
43.410759
43.425217
62.4718093
62.476726

Fig. 2. Relative energy of possible urea states. (a) Comparison of the urea neutral
cluster (lower curves) energy against the urea anion cluster (top curves) energy as a
function of the dielectric constant. Circles (red) are for TZV2P-B3LYP calculations,
boxes (green) are for TZV2P-PBE calculations, and triangles (blue) are for DZVP-PBE
calculations. (b) The sum of the molecular species that make up the urea neutral
(circles) and urea anion (triangles) clusters. The black curves are using a cavity
radius of 3.5 , and the gray curves have a cavity radius of 7.0 , indicating similar
trends in that both states are stabilized about the same amount with a smaller
cavity leading to greater stabilization. (For interpretation of the references to color
in this gure legend, the reader is referred to the web version of this article.)

ence between the ionic and neutral states [27] of the molecular
fragments that give 143.1 kJ/mol for the BLYP and 142.3 kJ/mol
for PBE. In Fig. 2b are shown the relative differences of the sum

59

J.M. Rimsza, L.R. Corrales / Computational and Theoretical Chemistry 987 (2012) 5761

of molecular fragments for the urea neutral cluster (circles) and the
urea anion clusters (triangles) with the neutral at K = 1 as the reference point. The black lines used a cavity radius of 3.5 while the
gray curves used a radius of 7.0 . The clusters of Fig. 2a were done
with the larger cavity. The results of Fig. 2b shows that a smaller
cavity leads to signicantly more stabilization, however, the trend
of the two clusters remain the same. In the formation of the 2:1
urea to choline chloride clusters there is an additional gain in energy in forming the clusters due to forming hydrogen bonds and
other electrostatic interactions. That additional energy is the difference between the two cluster energies minus the anion formation
energy. Overall, the trend is such that even by adding the dielectric
the neutral cluster is far more stable and, hence, is expected to be
the most populated state at room temperature as considered by
the presence of only a dielectric eld. Differences in the BLYP
and PBE with the TZV2P basis set are that the latter give lower energy but the relative differences remain similar. Difference in basis
sets also lead to similar relative differences.
In Table 2, the binding of copper and copper oxide with neutral
and anionic urea along with the chloride ion are listed for a K = 1
determined with the TZVP basis set and two different exchange
correlation functions. Binding of copper oxide with the neutral
and anion states of urea, Fig. 3a and b respectively, shows a gain
in energy of about 120 kJ/mol for the anion complex. There are several other meta-stable congurations that include binding with the
nitrogen atoms that are of higher energy and nearly degenerate
with each other that show the same relative trends between the
neutral and anion states. Calculations differences between BLYP
and PBE exchangecorrelation are minimal when comparing the
relative differences and otherwise the absolute energy is lower
for the latter exchange functional. Binding of elemental Cu (e.g. a
copper atom) with urea and its anion show a threefold increase
of the binding energy for the anion complex. It is of interest that
this binding energy is not only stronger than the neutral complex,
it is as strong as the neutral complex of CuO, while being weaker
than the anion complex with CuO. The fact that the urea neutral
and anion molecules binds to copper via the oxygen of the carbonyl, Fig. 4a and b respectively, is consistent with crystallographic
data of how urea likes to bind with copper [28]. It is also seen that
Cu binds with the nitrogen in both cases, Fig. 4c and d. Binding of
the chloride anion with CuO and Cu have the opposite energy trend
than the urea molecules in that binding to the elemental Cu is by
far stronger than binding to the oxide. There is a similarity in the
binding of elemental Cu with the chloride and urea anions whereas
with the CuO the urea anion would be preferred. In contrast to the
chloride anion the urea molecule binds more strongly to the CuO
than to elemental Cu.

Table 3
Binding energies of the 2:1 ureacholine chloride DES System with Cu and CuO with
neutral and anionic urea.
Complex

Binding energy (kJ/mole)


BLYP

Urea(0) + CuO
Urea(0) + Cu
Urea( 1) + CuO
Urea( 1) + Cu
Cl( 1) + CuO
Cl( 1) + Cu

159.5
49.01 ( 48.98)
277.8
153.9 ( 108.4)
67.96
168

Binding energy (kJ/mole)


PBE
160.8
59.30 ( 58.8)
285.7
165.1 ( 115.1)
69.89
180.7

Fig. 3. Copper Oxide urea complexes. (a) Urea(0) + CuO and (b) Urea( 1) + CuO
showing similar structures with the copper atom clearly bonding with the oxygen
atom of the carbonyl. In (b) a small bend to the OCu.

Fig. 4. Copper urea complexes. (a) Urea(0) + Cu with the Cu bound to the oxygen,
(b) Urea(0) + Cu with the Cu bound to the amines, (c) Urea( 1) + Cu with the Cu
bound to the oxygen, and (c) Urea( 1) + Cu with the Cu bound to the nitrogen.
Structures with the copper atom bonding with the oxygen atom of the carbonyl
show a much lower binding energy for the anion.

4. Discussion
The calculation of the binding strength and geometries provides
an insight to the expected structures in a DES liquid comprised of
urea and choline chloride. In that the majority of the solvent is
likely to contain neutral urea molecules, there would be a preferential binding of CuO with urea which is by far more favorable than
binding with the Cl anion. The ability of the solution to make the
anion is small, but may be enhanced by the formation of the anion
metal complexes which forms a strong binding complex. Such
complexes may be further thermodynamically stabilized in the
presence of the hydrogen bonding network provided by urea in

this system and requires further consideration because they would


release protons into the solvent system.
In the presence of a dielectric eld provided by a solvent comprised of dipolar molecules proton transfer may be further stabilized. However, the ability of this ureacholine chloride DES to
form anion states of urea is found to be limited even in the presence of a dielectric eld. This work has focused on considering only
an effective dielectric eld and has not explored the role of hydrogen bond network in stabilizing the release of a proton by a urea
molecule. Ionic liquids are expected to form more complex solvent
structures about these solutes such as those found in this work and

Table 2
Carbonyl bond length changes.
Molecule/complex

Urea(0)

Urea( 1)

Urea(0) + CuO

Urea( 1) + CuO

Urea(0) + Cu

Urea( 1) + Cu

CO bond distance (nm)

0.123

0.127

0.127

0.131

0.126

0.131

60

J.M. Rimsza, L.R. Corrales / Computational and Theoretical Chemistry 987 (2012) 5761

in crystallographic analogues recently provided in the literature.


This work provides guidance of what can be expected in molecular
dynamics simulations using potentials based on the interactions of
these levels of theory and basis sets.
Using the most basic means of embedding a cluster into a
dielectric continuum is not as sophisticated as COSMO or PCM
methods, but as shown elsewhere [29], the relative trends are similar and are not expected to change especially given the large difference between the neutral and molecular ionic states. Perhaps
the largest approximation being made in this work is that the
DES is in fact a highly hydrogen bonded network system that is ignored in using a dielectric continuum model. The use of DFT methods is motivated by the need to examine the extended liquid by ab
initio molecular dynamics and by classical molecular dynamics
methods to further investigate hydrogen bond network behavior
of this system.
The 2:1 ureacholine chloride DES has conductivity comparable
to that of a dilute potassium chloride solution that is more indicative of the chloride anion being the principle migration species as it
is more facile than the choline cation or urea anion [30]. There is
evidence that the experimental conductivity has different behavior
at low temperature versus higher temperature as shown in Fig. 2 of
Ref. [18] and Fig. 4 of Ref. [31], almost as if there is a high viscosity
versus a low viscosity behavior that has not been addressed. In the
low temperature range the viscosity decreases rapidly with
increasing temperature to an asymptotic limit that encompasses
the range where a break in slope of the conductivity curve is observable. In addition to that unexplained behavior it remains experimentally unclear if this DES contains free protons or protonated
molecular complexes along with urea anions in appreciable quantities that may appear as a function of increasing temperature, and
may also contribute to a change in conductivity. This work shows
that a proton would associate with the chloride ion that in a highly
hydrogen bonded networked system with other chloride ions present it is possible that proton hopping could ensue. Thus, further
characterization is needed by both the theoretical and experimental means to establish the full nature of this complex solvent
system.
A comment on applying a dielectric eld is warranted in that for
the large clusters a cavity radius of 7 is required. This has the effect of reducing the strength of the dielectric about each molecule
in the large cluster. Thus, the energy of the individual molecules in
two different size cavities, 7 and 3.5 is insightful. It is observed
that a smaller cavity leads to greater stabilization of the ionic species. This is true for the choline cation, chloride anion and the urea
anion. In contrast, there is little change for the urea and HCl molecules. Thus, the net effect is that the choline cation is present in
both clusters, so effectively canceling out its contributions to the
overall dielectric effect. While the urea anion show similar response to the dielectric as the choline cation, the chloride ion is almost doubled in stabilization in the dielectric as the urea anion.
Hence, the overall effect is that the presence of a dielectric tends
to stabilize the system without the urea anion. However, this does
not consider the formation of complexes in the hydrogen bonded
network of the actual liquid.

5. Conclusions
First-principles methods based on density function theory (DFT)
were employed to determine the extent of molecular ionic character of urea in a DES and how urea in its neutral and anion states
bind with CuO and elemental Cu. Overall, the theoretical approach
as used here predicts that this DES does not innately contain many
urea anions and that the neutral urea molecule complexes CuO
more favorably than Cu and its anion counterpart binds more

aggressively to both metal states. The latter is stable enough that


it may induce the formation of an anion and hence release a proton. It is found that the urea anion has a threefold increase in binding with elemental Cu as compared to neutral urea indicating that
etching could occur in the presence of the anion. The chloride anion is shown to favor binding with elemental Cu over CuO. In this
particular DES, an increase in temperature would shift the equilibrium between the molecular neutral and ion states so as to increase the molecular ion concentration, hence, increasing the
etching ability of this solvent system.
Acknowledgments
The authors would like to acknowledge extensive discussions
with S. Raghavan and D.P.R. Thanu who provided experimental
data and much insight into these complex solvent systems.
References
[1] A.P. Abbot, D. Boothby, G. Capper, D.L. Davies, R.K. Rasheed, Deep eutectic
solvents formed between choline chloride and carboxylic acids: versatile
alternatives to ionic liquids, J. Am. Chem. Soc. 126 (2004) 91429147.
[2] A.P. Abbot, F. Khalid, M. Gero, K.J. Ryder, S. Karl, Electrodeposition of copper
composites from deep eutectic solvents based on choline chloride, Phys. Chem.
Chem. Phys. 11 (2009) 4269.
[3] C.A. Nkuku, R.J. LeSuer, Electrochemistry in deep eutectic solvents, J. Phys.
Chem. 111 (2007) 1327113277.
[4] S. Natarajan, S. Mandal, Open-framework structures of transition-metal
compounds, Angew. Chem. Int. Ed. 47 (2008) 47984828.
[5] F. Himeur, I. Stein, D.S. Wragg, A.M.Z. Slawin, P. Lightfoot, The ionothermal
synthesis of metal organic frameworks, Ln(C9O6H3)((CH3NH)2CO)2, using deep
eutectic solvents, Solid State Sci. 12 (2010) 418421.
[6] H. Olivier-Bourbigou, L. Magna, D. Morvan, Ionic liquids and catalysis: recent
progress from knowledge to applications, Appl. Catal. A-Ge. 373 (2010) 156.
[7] B. Singh, H. Lobo, G. Shankarling, Selective N-alkylation of aromatic primary
amines catalyzed by bio-catalyst or deep eutectic solvent, Catal. Lett. 141
(2011) 178182.
[8] K. Lunstroot, K. Driesen, P. Nockemann, K. Van Hecke, L. Van Meerevelt, C.
Goerller-Walrand, K. Binnemans, S. Bellayer, L. Viau, J. Le Bideau, A. Vioux,
Lanthanide-doped luminescent ionogels, Dalton Trans. (2009) 298306.
[9] S.D. A Lawes, S.V. Hainsowrth, P. Blake, K.S. Ryder, A.P. Abbot, Lubrication of
steel/steel contacts by choline chloride ionic liquids, Tribol. Lett. 37 (2010) 2.
[10] M. Steichen, M. Thomassey, S. Siebentritt, P. Dale Jr., Controlled
electrodeposition of CuGa from a deep eutectic solvent for low cost
fabrication of CuGaSe2 thin lm solar cells, Phys. Chem. Chem. Phys. 13
(2011) 42924302.
[11] T. Conard, S. List, M. Claes, B. Beckhoff, Advanced metrologies for cleans
characterization: ARXPS, GIXF and NEXAFS, Solid State Phenom. 134 (2007)
281.
[12] N. Venkataraman, A.K. Muthukumaran, S. Raghavan, Evaluation of copper
oxide in copper selectively of chemical systems for BEOL cleaning through
electrochemical investigations, Mater. Res. Soc. Symp. Proc. 990 (2007) 197
202.
[13] D.R. MacFarlane, K.R. Seddon, Ionic liquids progress on the fundamental
issues, Aus. J. Chem. 60 (2007) 35.
[14] P. Nockemann, B. Thijs, T.N. Parac-Vogt, K. Van Hecke, L. Van Meervelt, B.
Tinant, I. Hartenbach, T. Schleid, V.T. Ngan, M.T. Nguyen, K. Binnemans,
Carboxyl-functionalized task-specic ionic liquids for solubilizing metal
oxides, Inorg. Chem. 47 (2008) 99879999.
[15] A.P. Abbot, G. Frisch, K.S. Ryder, Metal complexation in ionic liquids, Annu.
Rep. Prog. Chem. Sect. A: Inorg. Chem. 104 (2008) 2145.
[16] S.T. Handy, Room temperature ionic liquids: different classes and physical
properties, Curr. Org. Chem. 9 (2005) 959988.
[17] K. Haerens, E. Matthijs, K. Binnemans, B. Van de Bruggen, Electrochemical
decomposition of choline chloride based ionic liquid analogues, Green Chem.
11 (2009) 13571365.
[18] A.P. Abbot, G. Capper, D.L. Davies, R.K. Rasheed, V. Tambyrajah, Novel solvent
properties of choline chloride/urea mixtures, Chem. Commun. (2003) 7071.
[19] S.G. Raptis, J. Anastassopoulou, T. Theophanides, Vibrational and theoretical
studies of urea and magnesium-urea complexes, Theor. Chem. Acc. 105 (2000)
156164.
[20] Use of CP2k CP2K Project Homepage. BerliOS, 14 February 2011. Web. 21
April 2011. <http://cp2k.berlios.de/>.
[21] Lippert Lippert 1997.
[22] J. VandeVondele, J. Hutter, Gaussian basis sets for accurate calculations on
molecular systems in gas and condensed phases, J. Chem. Phys. 127 (2007)
114105.
[23] M-C Kim, E. Sim, K. Burke, Communication: avoiding unbound anions in
density functional calculations, J. Chem. Phys. 134 (2011) 171103.

J.M. Rimsza, L.R. Corrales / Computational and Theoretical Chemistry 987 (2012) 5761
[24] A. Gobbi, G. Frenking, Y-conjugated compounds: the equilibrium geometries
and electronic structures of guanidine, guanidinium cation, urea, and 1,l
diaminoethylene, J. Am. Chem. Soc. 115 (1993) 23622372.
[25] A.S. Davies, W.O. George, S.T. Howard, Ab initio and DFT computer studies of
complexes
of
quaternary
nitrogen
cations:
trimethylammonium,
tetramethylammonium, trimethylethylammonium, choline and acetylcholine
with hydroxide, uoride and chloride anions, Phys. Chem. Chem. Phys. 5
(2003) 45334540.
[26] R.A. Marcus, On the thoery of oxidation-reduction reactions involving electron
transfer I, J. Chem. Phys. 24 (1956) 966978.
[27] H. Markusson, J.P. Belleres, P. Johansson, C.A. Angell, P. Jacobsson, Prediction of
macroscopic properties of protic ionic liquids by ab initio calculations, J. Phys.
Chem. A 111 (2007) 87178723.

61

[28] T. Prior, R. Kift, Synthesis and crystal structures of two metal urea nitrates, J.
Chem. Crystallogr. 39 (2009) 558563.
[29] J. Gibbs, L.R. Corrales, Degree of dissociation and adsorption complexes of the
protic ionic liquid monoethanolamine and formic acid, J. Phys. Chem. B
(submitted for publication).
[30] D. Yue, Y. Jing, J. Sun, X. Wang, Y. Jia, Structure and ion transport behavior
analysis of ionic liquid analogues based on magnesium chloride, J. Mol. Liq.
158 (2011) 124130.
[31] A.P. Abbot, G. Clapper, S. Gray, Design of improved deep eutectic solvents using
hole theory, Chem Phys Chem. 7 (2006) 803806.

S-ar putea să vă placă și