Next Article in Journal
Localized Overheating Phenomena and Optimization of Spark-Plasma Sintering Tooling Design
Next Article in Special Issue
Synthesis of New RE3+ Doped Li1+xTa1−xTixO3 (RE: Eu, Sm, Er, Tm, and Dy) Phosphors with Various Emission Colors
Previous Article in Journal
Influence of Fracture Width on Sealability in High-Strength and Ultra-Low-Permeability Concrete in Seawater
Previous Article in Special Issue
Fabrication of a Poly(3-octylthiophene-2,5-diyl) Electrochemiluminescence Device Assisted by Perylene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Luminescent Anionic Nano System: d10 Metallocyanide Excimers and Exciplexes in Alkali Halide Hosts

1
Department of Chemistry, University of Maine, Orono, ME 04469, USA
2
Department of Chemistry, University of Rhode Island, Kingston, RI 02881, USA
*
Author to whom correspondence should be addressed.
Materials 2013, 6(7), 2595-2611; https://doi.org/10.3390/ma6072595
Submission received: 2 May 2013 / Revised: 8 June 2013 / Accepted: 18 June 2013 / Published: 25 June 2013
(This article belongs to the Special Issue Luminescent Materials 2013)

Abstract

:
Dicyanoaurate, dicyanoargentate, and dicyanocuprate ions in solution and doped in different alkali halide hosts exhibit interesting photophysical and photochemical behavior, such as multiple emission bands, exciplex tuning, optical memory, and thermochromism. This is attributed to the formation of different sizes of nanoclusters in solution and in doped hosts. A series of spectroscopic methods (luminescence, UV-reflectance, IR, and Raman) as well as theoretical calculations have confirmed the existence of excimers and exciplexes. This leads to the tunability of these nano systems over a wide wavelength interval. The population of these nanoclusters varies with temperature and external laser irradiation, which explains the thermochromism and optical memory. DFT calculations indicate an MLCT transition for each nanocluster and the emission energy decreases with increasing cluster size. This is in agreement with the relatively long life-time for the emission peaks and the multiple emission peaks dependence upon cluster concentration.

1. Introduction

The term “excimer” was first introduced in 1960 by Stevens and Hutton when studying pyrene luminescence [1]. They defined excimer as a dimer associated in the electronic excited state and dissociated in its ground state, which is different from a stable dimer. Actually, the first excimer can be dated back to 1927, when Lord Rayleigh observed the ultraviolet emission spectrum of high-pressure mercury vapor [2]. Nobel gases, such Helium, Neon, Argon, and Krypton also form excimers to give intense emission spectra in the vacuum ultraviolet region [3,4]. Excimers are also common in liquid and solid phases, especially for aromatic compounds. As early as the 1950s, Förster and Kasper observed that the pyrene solution fluorescence depended on the concentration [5,6]. Since then, more and more organic compounds have been found to form excimers: benzene, naphthalene, 1,2-benzanthracene, and perylene, just to name a few [7,8,9,10,11].
For inorganic excimers, a matrix isolation study has been carried out for alkali metals, alkaline earth metals and group VIIB metals with absorption and luminescence spectra [12]. Platinum complexes have been reported to form different excimers in aqueous solution, such as [PtII(4,7-diphenyl-1,10-phenanthroline)(CN)2], [Pt2(P2O5H2)42−Tl], and [Pt2(P2O5H2)42−Au(CN)2] [13,14,15]. In aqueous solution, the formation of excimers is affected by the concentration of monomers; as the concentration increases, the emission peak intensity corresponding to the excimer is enhanced. Closed-shell heavy metals used to be considered as repelling each other due to the fully occupied valence orbitals. However, in the study of inorganic and organometallic compounds, experimental results have presented evidence of interaction between closed-shell metals [16,17,18,19,20,21,22,23,24,25]. The strength of these bonds is weaker than ionic (700–4000 kJ/mol) and covalent bonds (200–1000 kJ/mol) but stronger than a van der Waals bond (<5 kJ/mol), and they are about the energy of hydrogen bonds (10–40 kJ/mol) [26]. The strong aurophilic interaction is ascribed to the relativistic effect of gold. In chemistry, relativistic effect means when the electrons with high speed move close to a heavy nucleus, the mass increases consequently, and so does the effective nuclear charge [27]. Thus the less diffuse orbitals (s and p) contract radially and the orbital energy is stabilized, leading to a stronger shielding of the nuclear attraction. The more diffuse orbitals (d and f) expand radially and results in a destabilization of the orbital energy. Gold has a larger relativistic effect than any other element with Z < 100 [28]. The tendency of Au(I) to form larger clusters via a weak interaction is due to the stabilization of molecular orbitals derived from the filled 5d atomic orbitals by configuration mixing with empty molecular orbitals from 6s or 6p atomic orbitals [29].
Inspired by the Au(I) compounds, our group has reported the first example of ligand-unsupported Ag(I) complex with argentophilic interaction, Tl[Ag(CN)2] [30]. X-ray crystallography, luminescence, and temperature-dependent Raman spectra measurements were carried out. The crystal structure results show that the Ag–Ag bond distance is 3.11 Å and the low energy Ag–Ag vibrational frequency of about 75–125 cm−1 given by Raman indicates the presence of direct argentophilic interaction in Tl[Ag(CN)2]. Since then, our group has reported a series of d10 metal cyano compounds in nano systems [31,32,33,34,35,36,37,38,39,40,41]. We found that due to the formation of different size exciplexes, these nano systems show interesting photophysical and photochemical behaviors, such as emission tunability, thermochromism, and optical memory. Luminescence of these systems not surprisingly depends upon the content of cyanometallic compounds, similar to excimers in aqueous solution. In this paper, we review the research work concerning Ag(CN)2, Au(CN)2, and Cu(CN)2 nano systems, which possess interesting luminescence properties.

2. Ag(CN)2 Excimers and Exciplexes

The photoluminescence of solid state Tl[Ag(CN)2] indicates the existence of argentophilicity in Ag(I) compounds [30]. The Ag–Ag interaction is believed to result in the formation of *[Ag(CN)2] excimers in the bulky state. We carried out a study focusing on the *[Ag(CN)2]n in the nano state by doping Ag(CN)2 in KCl hosts [31,32,33,34,39]. The KAg(CN)2/KCl nano systems were prepared by a simple slow evaporation method and the concentration was determined using atomic absorption method.
Figure 1 shows the emission peaks for the KAg(CN)2/KCl single crystal. Multiple emission peaks at 295, 327, 345, 417, and 548 nm were observed for corresponding excitation wavelengths. These luminescence peaks are labeled as A, B, C, and D respectively, as indicated in Figure 1. For each emission band, the corrected excitation spectra were compared to the absorption of a KAg(CN)2 aqueous solution. The KAg(CN)2 solution absorbs at 196 nm, ascribed to a metal-to-ligand charge transfer process. However, the excitation maxima for A–D bands are at much lower energy and the homometallic interaction is believed to play an important role in the large red-shift of the excitation peaks.
Figure 1. Exciplex tuning by site-selective excitation: emission spectra of a KAg(CN)2/KCl crystal at 77 K with different excitation wavelengths. Intensities are not comparable between different spectra. (Reprinted with permission from [31]. Copyright 1998 American Chemical Society.)
Figure 1. Exciplex tuning by site-selective excitation: emission spectra of a KAg(CN)2/KCl crystal at 77 K with different excitation wavelengths. Intensities are not comparable between different spectra. (Reprinted with permission from [31]. Copyright 1998 American Chemical Society.)
Materials 06 02595 g001
To better understand the argentophilic interaction, electronic structure calculations using both extend Hückel and ab initio methods were conducted for a monomer, dimer and trimer doped in KCl hosts. As shown in Figure 2, a Ag(CN)2 monomer is present in the KCl cubic unit cell with substituting Ag+ and CN for K+ and Cl, thus the total charge of this system is unchanged. For dimers, there are three different configurations, eclipsed, perpendicular, and offset-eclipsed, which all have linear Ag(CN)2 units. Calculations for four trimers with linear and angular arrangements respectively were carried out as well.
Figure 2. Structure of a unit cell of KCl with a defect created by doping two adjacent Ag(CN)2 ions. In the defect, the Ag+ and CN ions are shown to occupy the K+ and Cl sites, respectively. (Reprinted with permission from [31]. Copyright 1998 American Chemical Society.)
Figure 2. Structure of a unit cell of KCl with a defect created by doping two adjacent Ag(CN)2 ions. In the defect, the Ag+ and CN ions are shown to occupy the K+ and Cl sites, respectively. (Reprinted with permission from [31]. Copyright 1998 American Chemical Society.)
Materials 06 02595 g002
Table 1 summarizes the calculated results for the equilibrium distance, binding energy, HOMO-LUMO gap and overlap population for [Ag(CN)2]n nanoclusters in ground and excited states. The bond distance between two Ag atoms is shorter in the excited state than in the ground state. In addition, the binding energy and the overlap population also increase significantly in the excited states in comparison with the ground states. These are consistent with the behavior of excimer and exciplex. The interaction between monomer units is stronger in the excited state, which results in a stable excited dimer, e.g., excimer. The multiple HOMO-LUMO gap energies by calculation are in agreement with the different luminescence bands (A to D) observed experimentally. Interestingly, the HOMO-LUMO gap show a trend that with an increase in the number of interacting Ag(CN)2, the emission energy decreases. This explains the exciplex tuning behavior and the different bands A, B, C, and D are tentatively assigned to excimer, cis-localized exciplexs, trans-localized exciplexes, and delocalized exciplexes respectively. The ab initio calculation results for all [Ag(CN)2]n nanoclusters indicate a antibonding HOMO to bonding LUMO transition, which is also favorable for the formation of excimers and exciplexes.
Table 1. Summary of the Results of Extended Hückel Calculations for the Ground and Excited States of Oligomeric Species of Dicyanoargentate(I). (Adapted with permission from [31]. Copyright 1998 American Chemical Society.)
Table 1. Summary of the Results of Extended Hückel Calculations for the Ground and Excited States of Oligomeric Species of Dicyanoargentate(I). (Adapted with permission from [31]. Copyright 1998 American Chemical Society.)
Species[S][S]2*[S]2[S]2*[S]2[S]3*[S]3[S]3*[S]3[S]3*[S]3
(ec)(ec)(st)(T1,st)(T1,ec)(T1,ec)(T1,st)(T1,st)(T1,ec)(T1,ec)
Ag–Ag eq. diet, A8.00d3.583.002.882.393.493.092.792.463.543.15
Binding energy, eV0.000.131.120.221.320.331.470.612.000.291.14
H–L gap, eV4.954.373.984.284.134.003.553.733.464.173.83
O.P. (Ag–Ag)0.0000.0030.0340.0690.089−0.0080.0270.0390.0790.0030.048
The Ag(CN)2/KCl nano system has an extreme wide luminescence tuning range over 18,000 cm−1, compared to some other systems [42,43,44]. This new optical behavior is referred to as exciplex tuning, in which the exciplex means excited state oligomers, not limited to dimers. The photoluminescence of the crystal can also be tuned by the content of the Ag in the host [32]. The higher energy bands are predominant for lower silver concentration whereas the spectra for higher silver concentration are mostly of lower energy bands. This further confirms the existence of different size nanoclusters in the doped system and the time-resolved luminescence spectra suggest energy transfer processes from dimers to trimers. The increase of trimer luminescence intensity is related to higher statistical population of trimers in Ag concentrated crystals and more acceptors for the energy transfer.
In addition to the novel exciplex tuning behavior, the Ag(CN)2/KCl nano system exhibits interesting thermochromism phenomena [33]. Figure 3 shows the emission spectra of a KCl/KAg(CN)2 crystal at different temperatures. At room temperature, only one band (A) is observed while excited at 235 nm. However, as the temperature is cooled down to liquid nitrogen temperature (~77 K), band A almost disappears and a strong emission band C at lower energy appears. As the temperature is further decreased to 12 K, the emission spectra show predominantly band B at intermediate energy. Since the luminescence pattern changes dramatically with temperature, the possible cause relating to the Ag–Ag bond distance change is excluded. Another possibility that these emission bands are dependent on fluorescence and phosphorescence is also ruled out, because all the emission bands have a lifetime of microsecond magnitude. Therefore, the thermochromism behavior is due to some major structural changes. A kinetic model based on “normal” and back energy transfer is proposed. At liquid helium temperature, no energy transfer is involved and the luminescence is only from direct excitation. With temperature increasing to 80 K, “normal” energy transfer processes from high energy nanoclusters to low energy ones prevail and thus band C intensity increases in this temperature range. However, a further temperature increase leads to back energy transfer with reverse direction between nanoclusters. The back energy transfer is from low energy to high energy, which is not thermodynamically favorable. The higher activation energy than that of normal energy transfer is compensated by higher temperature, e.g., the back energy is significant at higher temperatures. In addition to KCl, different alkali halides, such as NaCl, NaBr, KBr, and NaF were used to study the host effect. Similar experimental results were obtained except for NaF, which is likely due to the small lattice size leaving no space for the Ag(CN) clusters.
Figure 3. Emission spectra of a KCl/KAg(CN)2 crystal as a function of temperature and excitation wavelength. Thin and thick curves represent spectra obtained with excitation wavelengths of 235 and 270 nm, respectively. Intensities are comparable between spectra at the same temperature but not comparable between spectra at different temperatures. (Reprinted with permission from [33]. Copyright 2000 American Chemical Society).
Figure 3. Emission spectra of a KCl/KAg(CN)2 crystal as a function of temperature and excitation wavelength. Thin and thick curves represent spectra obtained with excitation wavelengths of 235 and 270 nm, respectively. Intensities are comparable between spectra at the same temperature but not comparable between spectra at different temperatures. (Reprinted with permission from [33]. Copyright 2000 American Chemical Society).
Materials 06 02595 g003
Figure 4. Emission spectra versus 266 nm laser irradiation time at 77 K for batch 1 of [Ag(CN)2]/KCl crystals. All spectra were scanned with 275 nm excitation. Note the dominance of the short-wavelength bands following irradiation at 77 K and the regeneration of the original spectrum in the recovery step. (Reprinted with permission from [39]. Copyright 2000 American Chemical Society.)
Figure 4. Emission spectra versus 266 nm laser irradiation time at 77 K for batch 1 of [Ag(CN)2]/KCl crystals. All spectra were scanned with 275 nm excitation. Note the dominance of the short-wavelength bands following irradiation at 77 K and the regeneration of the original spectrum in the recovery step. (Reprinted with permission from [39]. Copyright 2000 American Chemical Society.)
Materials 06 02595 g004
Optical memory is another interesting property resulting from the presence of the nanoclusters in the doped KCl host [39]. As shown in Figure 4, the Ag(CN)2/KCl crystal show two emission peaks at 77 K. After irradiation at 266 nm for 90 min, the emission spectra have a significant change with the intensity of high energy peak increasing whereas the low energy peak intensity decreasing. Then, when the crystal is brought up to a warm temperature and cooled down again to 77 K again, the emission pattern recovers back to where it was at the beginning. In the whole process, acquiring emission spectra is “read”, irradiating laser beam on the crystal is “write”, and changing the temperature leads to “erase”. Therefore, this is a reversible write/read/erase optical memory behavior, which is similar to the research done by Zink and coworkers [45,46,47]. During the laser irradiations, the nanoclusters are excited and the distribution of different size excimers and exciplexes is changed, leading to the luminescence intensity change in the emission spectra. The kinetic study was conducted for every five min of laser irradiation and the activation energy was extrapolated via the emission peak intensity versus the logarithm of time. The obtained activation energy around 261 J/mol suggests a fast energy transfer between these *[Ag(CN)2]n excimers and exciplexes.

3. Au(CN)2 Excimer and Exciplexes

Our group has carried out a study on Ag(CN)2 and Au(CN)2 in aqueous as well as other organic solutions and observed non-Beer’s law behavior of the absorption for different concentrations of solution, as shown in Figure 5 [35,36]. The luminescence spectra of K[Au(CN)2] aqueous solution at room temperature and 77 K exhibit different emission peaks and the pattern depends upon the excitation wavelength. Theoretical calculations indicate the energy gap between the HOMO and LUMO decreases as the oligomer size increases. Both experimental and theoretical results suggest the presence of [Au(CN)2]n excimers and exciplexes. Consequently, investigations on KAu(CN)2 doped in an alkali halide host was conducted to further understand the close-shell interaction between Au(I) complexes [38].
Figure 5. Absorption spectra versus concentration of K[Au(CN)2] in aqueous solutions at ambient temperature. (Reprinted with permission from [35]. Copyright 2002 American Chemical Society.)
Figure 5. Absorption spectra versus concentration of K[Au(CN)2] in aqueous solutions at ambient temperature. (Reprinted with permission from [35]. Copyright 2002 American Chemical Society.)
Materials 06 02595 g005
Figure 6 shows the emission spectra of a KAu(CN)2/KCl single crystal at 77 K. Same as the argentocyanide/alkali halide system, the Au(CN)2/KCl also exhibit exciplex tuning behavior. The excitation ranges from 270 nm to 350 nm and the emission peaks are at 335, 390, and 425 nm respectively. However, the pure KAu(CN)2 crystal only shows one emission peak at 390 nm with different excitation wavelengths. Table 2 lists the emission maxima for KAu(CN)2 in an alkali halide host and aqueous solution, as well as the corresponding assignments. The high energy emission peak at 275–285 nm in KAu(CN)2 aqueous solution was not observed for the solid state, implying no Au(CN)2 monomer species is present in the doped KCl crystal. However, for the aqueous solution, due to the limit of concentration, low energy emission peak from the larger size of [Au(CN)2]n does not exist. These results are in agreement with a model where the homometallic interactions between different Au(CN)2 units form excimers and exciplexes. The lifetime measurements show microsecond magnitude phosphorescence, suggesting a metal to ligand charge-transfer process. In addition, Raman spectroscopy results provide further evidence for the formation of different nanoclusters in the doped crystal. A pure KAu(CN)2 crystal exhibits only one peak in the cyanide stretching region (about 2176 cm−1), whereas the doped crystals have additional peaks at 2169 and 2189 cm−1. This is consistent with the photoluminescence results. The homometallic interaction in [Au(CN)2]n nanoclusters affects the back bonding from gold to the cyanide ligand, which results in the shift of cyanide stretching frequency.
Figure 6. Emission spectra of a single crystal of KAu(CN)2/KCl at 77 K with different excitation wavelengths. (Reprinted with permission from [38]. Copyright 2002 American Chemical Society.)
Figure 6. Emission spectra of a single crystal of KAu(CN)2/KCl at 77 K with different excitation wavelengths. (Reprinted with permission from [38]. Copyright 2002 American Chemical Society.)
Materials 06 02595 g006
Table 2. General Qualitative Assignment of the Emission Bands Observed in Solids and Solutions of Au(CN)2-Species. (Adapted with permission from [38]. Copyright 2002 American Chemical Society.)
Table 2. General Qualitative Assignment of the Emission Bands Observed in Solids and Solutions of Au(CN)2-Species. (Adapted with permission from [38]. Copyright 2002 American Chemical Society.)
bandsolids, λmaxem, nmsolutions λmaxem, nmassignment
I275–285*[Au(CN)2]2
II320–355320–350bent *[Au(CN)2]3
III370–395380–390linear *[Au(CN)2]3
IV420–450420–440*[Au(CN)2]4
V455–490*[Au(CN)2]n a
VI600–640*[Au(CN)2]n
a *[Au(CN)2]n represents delocalized exciplexes.
Theoretical calculations were conducted using Extended Hückel methods for [Au(CN)2] monomer, dimer, trimer, and tetramer. The energy results of the ground state and the first excited-state results for these oligomers show that with the increase of the size, the binding energy between [Au(CN)2] units increase and the HOMO-LUMO gap decreases. All the results indicate the bonding between Au-Au is stronger in the excited state and this confirms the existence of [Au(CN)2]n excimers and exciplexes. For the trimer, the geometry also plays an important role and leads to different emission centers. The dynamics of [Au(CN)2]n oligomer geometry changes have been studied by ultrafast spectroscopic experiments [48]. The time-resolved results show direct evidence of a shortening of Au-Au distance and further support the formation of [Au(CN)2]n nanoclusters.

4. Cu(CN)2 Excimer and Exciplexes

As the last member in the coinage group, copper has a smaller relativistic effect than gold and silver and the homometallic interaction between Cu(I) compounds should be less [28]. Moreover, most of the reported luminescent Cu(I) complexes are cationic with chelating imine ligands or bisphosphine ligands or cuprous clusters [49]. However, our group has reported photophysical and photochemical properties of anionic Cu(CN)2 doped in alkali halide hosts which resulted from the close-shell interaction of Cu(I) [41].
Unlike pure KAg(CN)2 and KAu(CN)2 which have linear M(CN)2, the dicyanocuprate salts have a local trigonal Cu(CN)2 structure. X-ray diffraction (XRD) results indicate that both NaCu(CN)2 2H2O and KCu(CN)2 have a polymeric-chain structure. NaCu(CN)2 has zigzag CuCN chains linked to [Na(H2O)2]+ polyhedra and the copper has a trigonal-planar structure, with two bridge cyanides for the polymeric chain and one for the [Na(H2O)2]+ polyhedra. However, in KCu(CN)2, the polymeric chain is along a two-fold screw axis and the three-coordination of copper is not planar. However, when NaCu(CN)2 and KCu(CN)2 is doped in alkali halide hosts, the copper(I) cation becomes six-coordinated within the crystal field. With two cyanides and four surrounding halide atoms; the [Cu(CN)2X4]5− (X for halides) has a local D4h symmetry. Figure 7 gives IR and Raman spectra in the CN stretching region for the pure and NaCl doped dicyanocuprate(I) crystals. The pure NaCu(CN)2·2H2O, KCu(CN)2 have two CN vibrational modes in both IR and Raman, which is due to the local C2v symmetry of trigonal Cu(CN)2. For the doped crystals, the same results were observed for both dicyanocuprate salts, suggesting the linear structure in NaCl hosts. Same results were obtained for NaBr and KBr doped Cu(CN)2 crystals. This structural change is also confirmed by UV-vis diffuse reflectance and aqueous solution absorption spectroscopy experiments.
Figure 7. The υCN region IR and Raman spectra of pure NaCu(CN)2·2H2O, KCu(CN)2 and doped in NaCl. (Reprinted with permission from [41]. Copyright 2012 American Chemical Society.)
Figure 7. The υCN region IR and Raman spectra of pure NaCu(CN)2·2H2O, KCu(CN)2 and doped in NaCl. (Reprinted with permission from [41]. Copyright 2012 American Chemical Society.)
Materials 06 02595 g007
Figure 8. Luminescence spectra at 77 K of (a) NaCu(CN)2 and (b) KCu(CN)2 doped in NaCl. (Reprinted with permission from [41]. Copyright 2012 American Chemical Society.)
Figure 8. Luminescence spectra at 77 K of (a) NaCu(CN)2 and (b) KCu(CN)2 doped in NaCl. (Reprinted with permission from [41]. Copyright 2012 American Chemical Society.)
Materials 06 02595 g008
The Cu(CN)2 doped in alkali halide host systems show a series of interesting photophysical and photochemical properties. Similar to Au(CN)2 and Ag(CN)2, the dicyanocuprate ion in sodium chloride crystal exhibits exciplex tuning, as indicated in Figure 8. Both NaCu(CN)2·2H2O and KCu(CN)2 in NaCl show the same luminescence pattern, which provides further evidence for the linear structure of Cu(CN)2 in doped hosts. Other alkali halides, such as NaBr and KBr, were also used and the result was that all the crystals have exciplex tuning. For each host, both NaCu(CN)2·2H2O and KCu(CN)2 give the same luminescence pattern. The intensity of emission peaks depends upon the concentration of the copper in these crystals. The relative peak intensity of low energy peak to high energy peak increases with higher Cu(CN)2 content. As we have discussed for Ag(CN)2/KCl crystals, the lower energy peaks are assigned to larger size nanoclusters, therefore in the copper case, the increase of Cu(CN)2 concentration consequently leads to the formation of more large size oligomers and the lower energy peaks arise.
Thermochromism and optical memory were also observed for Cu(CN) doped crystals. Between 77 and 200 K, all the peak intensity decreases due to the quenching with increasing temperature. However, above 200 K, the intensity of higher energy peaks abruptly increases and the lower energy peak almost disappears. This is ascribed to the population change during the temperature increasing; the large size nanoclusters break into small size monomers and dimers, which take place between 200 and 250 K. After the reaction is completed, the distribution of nanoclusters no longer changes and the peak intensity is only affected by the temperature. The optical memory experiment was conducted in three steps: (1) the doped crystal is cooled down to liquid nitrogen temperature and during the cooling process, the crystal is irradiated by the 266 nm laser, which is the writing; (2) the emission spectra are recorded at 77 K, the reading; and (3) the crystal is heated up to the room temperature and then cooled down to 77 K again. The luminescence spectra are shown in Figure 9. Excimer and exciplex are more stable and the metal-metal distance is shorter for excited states. The laser irradiation promotes Cu(CN)2 ions into an excited state and thus form larger size nanoclusters. With the temperature decrease, they are frozen in their positions, which leads to the distribution change of different nanoclusters.
Theoretical density functional calculations were carried out for monomers and possible dimers and the possible configurations are shown in Scheme 1. The calculated electron density results give a major metal-to-ligand charge transfer (MLCT) mixed with ligand-to-ligand charge transfer (LLCT) transition and this is in agreement with the long life time (μs) of all emission peaks, suggesting the phosphorescence process. Figure 10 gives the isodensity map of some of the [Cu(CN)2]n in alkali halide hosts. The optimized structure for monomers and dimers were obtained in the ground state and the first excited triplet state. Comparing the bond distances between the ground state and the first excited triplet state, we find that the Cu–C bond distances contract and the C≡N bond distances expand, which is in accordance with the MLCT transition. Time-dependent DFT calculation results show the absorption and emission energies follow the trend that with the size of nanoclusters increase, the corresponding energies red shift. This is also observed from the experimental results. Atomistic calculations show that the potassium salts have greater bridging state selectivity and more favorable oligomerization energy than that of sodium salts. Bromide salts show a slightly greater bridging state selectivity than chloride salts.
Figure 9. Emission spectra of Cu(CN)2 doped in KBr at 77 K without laser irradiation (black solid), with 266 nm laser irradiation (blue solid), and recover, e.g., heat to room temperature then cool down to 77 K without laser irradiation (red dashed). (Reprinted with permission from [41]. Copyright 2012 American Chemical Society.)
Figure 9. Emission spectra of Cu(CN)2 doped in KBr at 77 K without laser irradiation (black solid), with 266 nm laser irradiation (blue solid), and recover, e.g., heat to room temperature then cool down to 77 K without laser irradiation (red dashed). (Reprinted with permission from [41]. Copyright 2012 American Chemical Society.)
Materials 06 02595 g009
Scheme 1. Dimeric configurations of Cu2(CN)4X6 or 7]8− or 9− ions. (Reprinted with permission from [41]. Copyright 2012 American Chemical Society.)
Scheme 1. Dimeric configurations of Cu2(CN)4X6 or 7]8− or 9− ions. (Reprinted with permission from [41]. Copyright 2012 American Chemical Society.)
Materials 06 02595 g011
Figure 10. Isodensity of HOMO (left) and LUMO (right) for Cu(CN)2 monomers (top), μ-Cl2 dimers (middle) and μ-ClCN dimers (bottom). (Reprinted with permission from [41]. Copyright 2012 American Chemical Society.)
Figure 10. Isodensity of HOMO (left) and LUMO (right) for Cu(CN)2 monomers (top), μ-Cl2 dimers (middle) and μ-ClCN dimers (bottom). (Reprinted with permission from [41]. Copyright 2012 American Chemical Society.)
Materials 06 02595 g010

5. Conclusions

The dicyano compounds of coinage metals, Ag, Au, and Cu doped in alkali halide hosts show examples of close-shell homometallic interactions, which results in a series of interesting photophysical and photochemical properties. These nanoclusters are typical excimers and exciplexes, e.g., more stable at the excited states. Especially for copper, the Cu(CN)2 in alkali halide crystals are unique examples for luminescent anionic Cu(I) compounds. The population of different nanoclusters changes with the concentration, temperature, and laser irradiation. Lifetime results of microsecond magnitude suggest phosphorescence and are consistent with the calculation results, which gives a MLCT transition with slightly LLCT process. The exciplex tuning gives a range as wide as 18,000 cm−1 and the strong phosphorescence has possible application for solid state lasers. Thermochromism and optical memory properties can be used to make new photosensors and memory devices.

Acknowledgments

This work was supported by the National Science Foundation (CHE-0443345).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Stevens, B.; Hutton, E. Radiative life-time of the pyrene dimer and the possible role of excited dimers in energy transfer processes. Nature 1960, 186, 1045–1046. [Google Scholar] [PubMed]
  2. Rayleigh, L. Series of emission and absorption bands in the mercury spectrum. Proc. R. Soc. A 1927, 116, 702–719. [Google Scholar] [CrossRef]
  3. Hopefield, J.J. Certain effects accompanying electron diffraction. Phys. Rev. 1930, 35, 1133–1134. [Google Scholar] [CrossRef]
  4. Birks, J.B. Excimers. Rep. Prog. Phys. 1975, 38, 903–974. [Google Scholar] [CrossRef]
  5. Förster, T.H.; Kasper, K. Ein Konzentrationsumschlag der Fluoreszenz. Z. Phys. Chem. (N.F.) 1954, 1, 275–277. [Google Scholar] [CrossRef]
  6. Förster, T.H.; Kasper, K. Ein Konzentrationsumschlag der Fluoreszenz des Pyrens. Z. Elektrochem. 1955, 59, 976–980. [Google Scholar]
  7. Ivanova, T.V.; Mokeeva, G.A.; Sveshnikov, B.Y.A. Dependence of the fluorescence of benzene, toluene, and p-xylene solutions on the concentration of fluorescent substance. Opt. Spectrosc. 1962, 12, 325–328. [Google Scholar]
  8. Döller, E.; Förster, T.H. Der Konzentrationsumschlag der Fluoreszenz des Naphthalins. Z. Phys. Chem. (N.F.) 1962, 31, 274–277. [Google Scholar] [CrossRef]
  9. Birks, J.B.; Christophorou, L.G. Excimer fluorescence of aromatic hydrocarbons in solution. Nature 1962, 194, 442–444. [Google Scholar] [CrossRef]
  10. Birks, J.B.; Christophorou, L.G. Resonance interactions of fluorescent organic molecules in solution. Nature 1962, 196, 33–35. [Google Scholar] [CrossRef]
  11. Birks, J.B.; Christophorou, L.G. “Excimer” fluorescence. IV. Solution spectra of polycyclic hydrocarbons. Proc. R. Soc. A 1964, 277, 571–582. [Google Scholar] [CrossRef]
  12. Miller, J.C.; Andrews, L. Absorption and laser-excited fluorescence spectra of matrix-isolated metal van der waals dimmers. Appl. Spectrosc. Rev. 1980, 16, 1–42. [Google Scholar] [CrossRef]
  13. Kunkely, H.; Vogler, A. Photoluminescence of [PtII(4,7-diphenyl-1,10-phenanthroline)(CN)2] in solution. J. Am. Chem. Soc. 1990, 112, 5625–5627. [Google Scholar] [CrossRef] [Green Version]
  14. Nagle, J.K.; Brennan, B.A. Luminescent exciplex formation involving Tetrakis(μ-diphosphito)diplatinate(II) and -Thallium(I) in aqueous solution. J. Am. Chem. Soc. 1988, 110, 5931–5932. [Google Scholar] [CrossRef]
  15. Pettijohn, C.N.; Jochnowitz, E.B.; Chuong, B.; Nagle, J.K.; Vogler, A. Luminescent Excimers and exciplexes of PtII compounds. Coord. Chem. Rev. 1998, 171, 85–92. [Google Scholar] [CrossRef]
  16. Harvey, P.D.; Gray, H.B. Low-lying singlet and triplet electronic excited states of binuclear (d10-d10) Palladium(0) and Platinum(0) complexes. J. Am. Chem. Soc. 1988, 110, 2145–2147. [Google Scholar] [CrossRef]
  17. Che, C.-M.; Kwong, H.-L.; Yam, V.W.-W.; Cho, K.C. Spectroscopic properties and redox chemistry of the phosphorescent excited state of [Au2(dppm)2]2+. J. Chem. Soc. Chem. Commun. 1989, 885–886. [Google Scholar]
  18. King, C.; Wang, J.-C.; Khan, M.N.I.; Fackler, J.P., Jr. Luminescence and metal-metal interactions in binuclear Gold(I) compounds. Inorg. Chem. 1989, 28, 2145–2149. [Google Scholar] [CrossRef]
  19. Perreault, D.; Drouin, M.; Michel, A.; Miskowski, V.M.; Schaefer, W.P.; Harvey, P.D. Silver and gold dimers. Crystal and molecular structure of Ag2(dmpm)2Br2 and [Au2(dmpm)2](PF6)2 and relation between metal-metal force constants and metal-metal separations. Inorg. Chem. 1992, 31, 695–702. [Google Scholar] [CrossRef]
  20. Fu, W.-F.; Chan, K.-C.; Miskowski, V.M.; Che, C.-M. The intrinsic 3[dσ*pσ] emission of binuclear Gold(I) complexes with two bridging diphosphane ligands lies in the near UV; Emissions in the visible region are due to exciplexes. Angew. Chem. Int. Ed. 1999, 38, 2783–2785. [Google Scholar] [CrossRef]
  21. Leung, K.-H.; Phillips, D.L.; Tse, M.-C.; Che, C.-M.; Miskowski, V.M. Resonance raman investigation of the Au(I)-Au(I) interaction of the 1[dσ*pσ] excited state of Au2(dcpm)2(ClO4)2. J. Am. Chem. Soc. 1999, 121, 4799–4803. [Google Scholar] [CrossRef]
  22. Che, C.-M.; Mao, Z.; Miskowski, V.M.; Tse, M.-C.; Chan, C.-K.; Cheung, K.-K.; Phillips, D.L.; Leung, K.-H. Cuprophilicity: Spectroscopic and structural evidence for Cu–Cu bonding interactions in luminescent dinuclear Copper(I) complexes with bridging diphosphane ligands. Angew. Chem. Int. Ed. 2000, 39, 4084–4088. [Google Scholar] [CrossRef]
  23. Che, C.-M.; Tse, M.-C.; Chan, M.C.W.; Cheung, K.-K.; Phillips, D.L.; Leung, K.-H. Spectroscopic evidence for argentophilicity in structurally characterized luminescent binuclear Silver(I) complexes. J. Am. Chem. Soc. 2000, 122, 2464–2468. [Google Scholar] [CrossRef]
  24. Fu, W.-F.; Chan, K.-C.; Cheung, K.-K.; Che, C.-M. Substrate-binding reactions of the 3[dσ*pσ] Excited state of binuclear Gold(I) complexes with bridging bis(dicyclohexylphosphino)methane ligands: Emission and time-resolved absorption spectroscopic studies. Chem. Eur. J. 2001, 7, 4656–4664. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, H.-X.; Che, C.-M. Aurophilic attraction and luminescence of binuclear Gold(I) complexes with bridging phosphine ligands: Ab initio study. Chem. Eur. J. 2001, 7, 4887–4893. [Google Scholar] [CrossRef] [PubMed]
  26. Pyykkö, P. Strong closed-shell interactions in inorganic chemistry. Chem. Rev. 1997, 97, 597–636. [Google Scholar] [CrossRef] [PubMed]
  27. Rose, J.; Grant, I.P.; Pyper, N.C. The direct and indirect effects in the relativistic modification of atomic valence orbitals. J. Phys. B 1978, 11, 1171–1176. [Google Scholar] [CrossRef]
  28. Pyykkö, P. Relativität, gold, wechselwirkungen zwischen gefüllten schalen und CsAu-NH3. Angew. Chem. 2002, 114, 3723–3728. [Google Scholar] [CrossRef]
  29. Vogler, A.; Kunkely, H. Absorption and emission spectra of tetrameric Gold(I) complexes. Chem. Phys. Lett. 1988, 150, 135–137. [Google Scholar] [CrossRef]
  30. Omary, M.A.; Webb, T.R.; Assefa, Z.; Shankle, G.E.; Patterson, H.H. Crystal structure, electronic structure, and temperature-dependent raman spectra of Tl[Ag(CN)2]: Evidence for ligand-unsupported argentophilic interactions. Inorg. Chem. 1998, 37, 1380–1386. [Google Scholar] [CrossRef] [PubMed]
  31. Omary, M.A.; Patterson, H.H. Luminescent homoatomic exciplexes in Dicyanoargentate(I) ions doped in Alkali halide crystals. 1. “Exciplex tuning” by site-selective excitation. J. Am. Chem. Soc. 1998, 120, 7696–7705. [Google Scholar] [CrossRef]
  32. Omary, M.A.; Hall, D.D.; Shankle, G.E.; Siemiarczuk, A.; Patterson, H.H. Luminescent homoatomic exciplexes in Dicyanoargentate(I) ions doped in Alkali halide crystals. 2. “Exciplex tuning” by Varying the dopant concentration. J. Phys. Chem. B 1999, 103, 3845–3853. [Google Scholar] [CrossRef]
  33. Rawashdeh-Omary, M.A.; Omary, M.A.; Shankle, G.E.; Patterson, H.H. Luminescence thermochromism in Dicyanoargentate(I) ions doped in Alkali halide crystals. J. Phys. Chem. B 2000, 104, 6143–6151. [Google Scholar] [CrossRef]
  34. Patterson, H.H.; Kanan, S.M.; Omary, M.A. Luminescent homoatomic exciplexes in Dicyanoargenate(I) ions doped in Alkali halide crystals. “Exciplex tuning” by site-selective excitation and variation of the dopant concentration. Coord. Chem. Rev. 2000, 208, 227–241. [Google Scholar] [CrossRef]
  35. Rawashdeh-Omary, M.A.; Omary, M.A.; Patterson, H.H. Oligomerization of Au(CN)2 and Ag(CN)2 ions in solution via ground-state aurophilic and argentophilic bonding. J. Am. Chem. Soc. 2000, 122, 10371–10380. [Google Scholar] [CrossRef]
  36. Rawashdeh-Omary, M.A.; Omary, M.A.; Patterson, H.H.; Fackler, J.P., Jr. Excited-state interactions for [Au(CN)2]n and [Ag(CN)2]n Oligomers in solution. Formation of luminescent gold–gold bonded excimers and exciplexes. J. Am. Chem. Soc. 2001, 123, 11237–11247. [Google Scholar] [CrossRef] [PubMed]
  37. Kanan, S.M.; Kanan, M.C.; Patterson, H.H. Photophysical properties of Ag(I)-exchanged zeolite A and the photoassisted degradation of malathion. J. Phys. Chem. B 2001, 105, 7508–7516. [Google Scholar] [CrossRef]
  38. Hettiarachchi, S.R.; Rawashdeh-Omary, M.A.; Kanan, S.M.; Omary, M.A.; Patterson, H.H.; Tripp, C.P. Spectroscopic studies of “exciplex tuning” for Dicyanoaurate(I) ions doped in potassium chloride crystals. J. Phys. Chem. B 2002, 106, 10058–10064. [Google Scholar] [CrossRef]
  39. Hettiarachchi, S.R.; Patterson, H.H.; Omary, M.A. Write/read/erase with laser irradiation of Dicyanoargentate(I) doped and pure crystals. J. Phys. Chem. B 2003, 107, 14249–14254. [Google Scholar] [CrossRef]
  40. Baril-Robert, F.; Li, X.; Welch, D.A.; Schneider, B.Q.; O’Leary, M.; Larochelle, C.L.; Patterson, H.H. Site-selective excitation of “exciplex tuning” for luminescent nanoclusters of Dicyanoargentate(I) ions doped in different Alkali halide crystals. J. Phys. Chem. C 2010, 114, 17401–17408. [Google Scholar] [CrossRef]
  41. Li, X.; Pan, Z.; Welch, D.A.; Baril-Robert, F.; Pike, R.D.; Patterson, H.H. An unusual luminescent anionic Copper(I) system: Dicyanocuprate(I) ion in nano and bulky states. J. Phys. Chem. C 2012, 116, 26656–26667. [Google Scholar] [CrossRef]
  42. Cummings, S.D.; Eisenberg, R. Tuning the excited-state properties of Platinum(II) diimine dithiolate complexes. J. Am. Chem. Soc. 1996, 118, 1949–1960. [Google Scholar] [CrossRef]
  43. Yersin, H.; Gliemann, G. Spectroscopic studies of Mx[Pt(CN)4yH2O. Ann. N.Y. Acad. Sci. 1978, 313, 539–559. [Google Scholar]
  44. Gliemann, G.; Yersin, H. Spectroscopic properties of the quasi one-dimensional Tetracyanoplatinate(II) compounds. Struct. Bond. 1985, 62, 87–153. [Google Scholar]
  45. Hollingsworth, G.; Barrie, J.D.; Dunn, B.; Zink, J.I. Optical memory in Cu+-Doped β″-Alumina. J. Am. Chem. Soc. 1988, 110, 6569–6572. [Google Scholar] [CrossRef]
  46. Barrie, J.D.; Dunn, B.; Hollingsworth, G.; Zink, J.I. Optical spectroscopy of Copper(I) doped β″-Alumina. J. Phys. Chem. 1989, 93, 3958–3963. [Google Scholar] [CrossRef]
  47. Barrie, J.D.; Dunn, B.; Zink, J.I. Optical spectroscopy of Cu+/Ag+ Doped β″-Alumina. J. Am. Chem. Soc. 1990, 112, 5701–5706. [Google Scholar] [CrossRef]
  48. Iwamura, M.; Nozaki, K.; Takeuchi, S.; Tahara, T. Real-time observation of tight Au−Au bond formation and relevant coherent motion upon photoexcitation of [Au(CN)2] oligomers. J. Am. Chem. Soc. 2013, 135, 538–541. [Google Scholar] [CrossRef] [PubMed]
  49. Armaroli, N.; Accorsi, G.; Cardinali, F.; Listorti, A. Photochemistry and photophysics of coordination compounds: Copper. Top. Curr. Chem. 2007, 280, 69–115. [Google Scholar]

Share and Cite

MDPI and ACS Style

Li, X.; Patterson, H.H. A Review of Luminescent Anionic Nano System: d10 Metallocyanide Excimers and Exciplexes in Alkali Halide Hosts. Materials 2013, 6, 2595-2611. https://doi.org/10.3390/ma6072595

AMA Style

Li X, Patterson HH. A Review of Luminescent Anionic Nano System: d10 Metallocyanide Excimers and Exciplexes in Alkali Halide Hosts. Materials. 2013; 6(7):2595-2611. https://doi.org/10.3390/ma6072595

Chicago/Turabian Style

Li, Xiaobo, and Howard H. Patterson. 2013. "A Review of Luminescent Anionic Nano System: d10 Metallocyanide Excimers and Exciplexes in Alkali Halide Hosts" Materials 6, no. 7: 2595-2611. https://doi.org/10.3390/ma6072595

Article Metrics

Back to TopTop