DOI:
10.1039/C6RA09763D
(Paper)
RSC Adv., 2016,
6, 55539-55545
Photoluminescence dynamics of copper nanoclusters synthesized by cellulase: role of the random-coil structure†
Received
15th April 2016
, Accepted 25th May 2016
First published on 27th May 2016
Abstract
This study demonstrates the one-pot cellulase-directed synthesis of magic numbered copper nanoclusters (Cu NCs) with blue-, cyan-, and green-luminescence from Cu12, Cu20, and Cu34, respectively. The developed Cu NCs have an excellent luminescence quantum yield, photostability and better colloidal stability. The temperature-dependent response of the blue luminescent nanocluster gets flipped over by changing the aqueous medium to a methanolic one. Moreover, the preferential stabilization of large-sized Cu34 (green-luminescent) NCs–cellulase over a relatively small one, Cu20 NCs–cellulase is found to be dependent on both temperature as well as solvent. The luminescence of Cu20 NCs–cellulase can be greatly augmented by the addition of Zn(II) and Cd(II) ions, whereas the Fe(III) ions quench the luminescence. The random coil structure of the enzyme dictates the size and luminescent properties of the Cu NCs. This unique and beneficial aspect of the Cu NCs warrants them to find possible applications as a sensitive marker for cell imaging, and nanophotonic materials.
Introduction
Noble metal nanoclusters (NCs) composed of several to tens of atoms (core size ≤ 2 nm) have generated burgeoning interest owing to their unique size-dependent luminescent and bioconjugation properties.1–20 They find profound applications in catalysis,2,3 optoelectronics,4 bioimaging5 and sensing applications.6 The low toxicity of these highly luminescent NCs makes them a better candidate to replace toxic quantum dots in biomedical applications.7,8 Metal NCs have molecule-like electronic transitions within their conduction bands due to quantum confinement.9 The spherical Jellium model, a conventional theoretical model, which is generally employed to predict the size of nanoclusters, states that the energy gap between the adjacent levels, i.e. the emission maximum of a NC is the indicator of its size,
, where Eem is the emission maximum of nanocluster of radius (R) with N number of metal atoms, EFermi is the Fermi energy of the bulk metal and rs is the Wigner–Seitz radius of the metals.10,11 The stability of the nanoclusters is dependent on its composition, i.e. number of atoms. Both gas and condensed-phase studies reveal the existence of magic-numbered clusters (N = 2, 8, 18, 20, 34, 40, and etc.) which possess better stability over others due to geometric and electronic shell closings.12,13 The shell closing rules for copper being a group IB element is expected to follow the same as Au, Ag. However, the preferential stabilization of a relatively large-sized magic-numbered cluster from the nearby one has not been explored in detail. In this report, we have observed such stabilization of copper nanoclusters by employing steady state and time-resolved fluorescence spectroscopy and circular dichroism studies.
The nanoclusters tend to agglomerate without proper protection and form nanoparticles, which are nonfluorescent. Inspired by biochemical synthetic routes for Au14 and Ag NCs,15 limited efforts have been made to stabilize Cu NCs using serum albumins,16–18 lysozyme,19 polymer,20 glutathione.21,22 Thus, the investigation of Cu NCs which are susceptible to oxidation needs to be carried out in great detail for primarily two reasons: the development of better synthetic methods and tuning of its luminescent properties for versatile use. Moreover, Cu is an important trace element in metal homeostasis23 and serves as a cofactor in metallochaperones, metalloregulatory and redox-active proteins.24 In this report, we unravel the synthetic methods using cellulase as sole reducing and stabilizing agent for Cu NCs in aqueous solution. Cellulase is a highly efficient enzyme synthesized by microbes for the extracellular degradation of substrates and biomineralization. The presence of metal ion impurities in cellulosic materials plays a key role in retarding the turn over number of enzyme. The catalytic role of enzyme toward degradation of cellulosic materials for bio-fuel production is of great interest due to global energy scarcity. However, a very limited report exists on cellulase mediated synthesis of nanostructures.25 In this context, it is very essential to understand the mechanism by which cellulase interacts with metal ions. Cellulase from Trichoderma reesei ATCC 26921 is composed of 344 amino acid residues of which 8 are cysteine and one is methionine.26,27 We report the synthesis of “magic numbered” Cu NCs by cellulase using two different synthetic routes: simple incubation and sonication followed by incubation of cellulase–metal ion solution, both the methods yield the formation of luminescent Cu NC. We have also observed that the enzyme structure regulates the size and luminescent properties of Cu NCs. The effect of temperature and solvent on luminescent properties of cellulase synthesized NCs have also been investigated.
Experimental
Materials
Cellulase from Trichoderma reesei ATCC 26921 is purchased from Sigma Aldrich and used as received. Copper sulphate and sodium hydroxide are obtained from RANKEM. Methanol (Fischer Scientific, HPLC grade) and metal salts such as zinc nitrate (SDFCL), mercuric chloride (Fisher Scientific), ferric chloride (Fisher Scientific) and cadmium chloride (Fisher Scientific) are obtained from the respective companies. All chemicals used in the experiments of analytical grade and used without any purification. Water from ELGA LAB Water-purifier (18.2 mΩ) is used throughout the experiments.
Cellulase-directed synthesis of copper nanoclusters
In a typical synthesis, 1 mL of 1.5 mg mL−1 cellulase dissolved in water is added to 1 mL, 1.5 mg mL−1 CuSO4 solution and mixed well. Few drops (30 μL) of NaOH (1 M) is added to the enzyme–metal ion solution to make the solution alkaline (pH ∼ 11). The solution is stirred for 2 hours, thereafter incubated at 4 °C for 3 days. The addition of NaOH to enzyme–metal ion solution changes the color of the solution from colorless to greenish and thereafter to reddish brown suggesting the formation of Cu NC. In an another method, the enzyme–metal ion solution is sonicated for 30 min, then stirred for 2 hours followed by incubation at 4 °C for 3 days. The color of solution changes to intense violet indicating the synthesis of Cu NC. It can be added here that the timescale to synthesize nanoclusters solely depends on the nature of the protein. However, the synthesized nanoclusters have been characterized by steady state and time-resolved fluorometric methods. The cellulase synthesized Cu NCs is added to methanol, lower polarity than that of water. The absorption and fluorescence spectra of the solvent mixture (solvent
:
water = 2
:
1) are recorded. The temperature of the aqueous and methanolic (2
:
1) media is also regulated to study its effect on the luminescent properties of different sized Cu NCs.
Instrumentation
The steady state absorption spectra are collected using a LAB India UV-Vis 3200 spectrophotometer. The absorption spectra are recorded from 200 to 1100 nm. Steady state photoluminescence (PL) spectra are recorded on a Varian Cary Eclipse fluorescence spectrophotometer with Peltier temperature controller. Synchronous luminescence studies are performed by scanning the excitation and the emission monochromators of the fluorimeter simultaneously, with a constant offset of 20 nm between the two. The relative quantum yield (ϕs) of the samples is defined as the area under the curve of excitation corrected fluorescence spectrum. The absolute luminescence quantum yield (ϕf) of the sample is determined by the equation, ϕf = (ϕs/ϕr) × φf, where ϕr and φf, are the relative and absolute quantum yield of the reference (coumarin 102).28 Time-resolved photoluminescence measurements are performed using a picosecond pulsed-diode laser-based TCSPC fluorescence spectrometer from LifeSpec II (Edinburg Instruments).29 The laser diodes: 375 nm with pulse width 170 ps and 440 nm of pulse width 180 ps are used as excitation sources. Emissions from the blue luminescent NC at 410 nm (λexc = 375 nm) and from red luminescent NCs at 485 nm and 600 nm (λexc = 440 nm) are collected at right angles to the direction of the excitation beam, at magic angle polarization and the resolution is 12 ps per channel. The data are fitted to multiexponential functions after deconvolution of the instrument response function by an iterative reconvolution technique, using the F900 data analysis software, where reduced χ2 and weighted residuals serve as parameters for goodness of fit. Circular dichroism data are recorded using Jasco J-815 CD spectrophotometer equipped with Peltier temperature controller, in the range of 190–400 nm at the rate of 100 nm s−1 with band width of 1 nm. CD spectra are measured using quartz cuvettes. The spectropolarimeter was purged with N2 prior to the experiment. Each CD plot is an average of three accumulated plots and also baseline corrected. The molar ellipticity is calculated from the observed ellipticity θ as 100θ/cl where c is the concentration of the protein solution in molarity and l is the path length of the cell in centimeters. The nanoclusters formed are imaged by a transmission electron microscope (TEM) (JEOL 2100F) mounted with field emission gun FEG TEM at 200 kV accelerating voltage. The samples for TEM are prepared on amorphous carbon films supported on a copper grid.30
Results & discussion
Cellulase-directed synthesis of copper nanoclusters
Cellulase as the sole reducing and nucleating agent unprecedentedly synthesizes copper nanoclusters in aqueous solution. Simple incubation (method-I) of cellulase–copper ion solution at 4 °C temperature over 72 hours results reddish brown color of solution, suggesting slow kinetics of Cu NC synthesis by cellulase alone (Scheme 1). Whereas, sonication (40 kHz, 15 min) followed by overnight incubation at 4 °C (method-II) of enzyme–metal ion solution changes the color of the solutions from colorless to greenish and finally to color resembling to Prussian blue (Fig. S1, ESI†). The pH of the medium is maintained is at ∼pH 11. In order to understand the cellulase mediated biomineralization process, the use of extraneous reducing agent has been avoided in particular. However, the synthesized nanoclusters are quite stable in aqueous solution for more than a month (Fig. S2 and S3†). The absorption spectra of pure cellulase and copper nanoclusters obtained from two different methods are plotted in Fig. 1A. The absorption spectra of cellulase alone have a peak centered at 278 nm which is characteristics of tryptophan, tyrosine and phenyl alanine residues. The absorption spectra of Cu NCs–cellulase, as expected, are markedly different and broader than that of enzyme alone, confirming the formation of nanoclusters. Cellulase from Trichoderma reesei ATCC 26921 is composed of 344 amino acid residues of which 8 are cysteine and one is methionine.26,27 Upon addition of copper sulphate solution to cellulase solution, the metal ions possibly be coordinated to amino acid residues of enzyme containing such as –NH, –SH, and –OH groups (Scheme 1) and they are subsequently reduced. Inhibition of nanocluster formation in acidic pH indicates the pivotal of role of concentration of hydroxide ion in enzyme-catalyzed nanocluster synthesis. We have reported that cellulase which contains 14 Tyr residues, even at low concentration can effectively synthesizes silver nanoparticles indicating its strong reducing ability.24 The standard reduction potential of Cu2+/Cu is 0.34 V (vs. NHE), whereas, the redox potentials for tyrȮ/tyrOH couple which is regulated by polypeptide chain lies in the range of 0.78–0.93 V (vs. NHE).31,32 The plausible mechanism for nanocluster formation is that the deprotonated tyrosine residues in cellulase at pH 11 undergo oxidation and yielding atomic copper metal. The regeneration of tyrosine residues might occur through 1e− reduction in the medium. The polypeptide chain acts as a scaffold for copper nanoclusters. At the molecular level the nanoparticle synthesis by cellulase is similar to the biomineralization process executed by living organisms in nature.33
 |
| | Scheme 1 Schematic representation of reaction methods employed for the synthesis of Cu NCs in cellulase solution. | |
 |
| | Fig. 1 [A] In aqueous solution, absorption spectra of – (i) cellulase (black line), (ii) Cu NC obtained in method-I (blue line), (iii) Cu NC obtained in method-II (red line). [B] Normalized excitation (dashed line) and emission spectra (solid line) of Cu NCs (method-I) in aqueous solution. [C] Normalized excitation (dotted line) and emission spectra (solid line) of Cu NCs (method-II) in aqueous solution. | |
Photophysical properties of Cu NCs–cellulase
Cu NCs obtained in method-I exhibit maximum excitation and emission peaks at 320 nm and 401 nm respectively (Fig. 1B). Excitation dependent emission spectra confirm the presence of one type of nanocluster and its spectral characteristics in aqueous solution (Fig. S4, ESI†). Surprisingly, the excitation spectra of Cu NCs prepared using sonication method are of vibronic in nature. Two excitation peaks are separated by about 28 nm, with peak maximum at 468 nm. The emission spectra exhibit also two peaks, maxima centered at 484 nm and second one at 511 nm (Fig. 1C). Its excitation dependent emission spectra also prove the vibronic coupling (Fig. S5, ESI†). However, in order to resolve the ground state heterogeneity, the synchronous luminescence technique is employed.34,35 In this technique, excitation and emission wavelengths are varied simultaneously with a constant offset, chosen by trial and error, such that the experiment is performed along the diagonal of the excitation–emission matrix, which passes through the maximum number of peaks and troughs. The resulting spectra have a greater number of narrower peaks compared with that obtained in conventional excitation spectroscopy, allowing discrimination between luminescent molecules that have broad and overlapping spectra. This method has previously been utilized to analyze complex mixtures in as widely separated fields as hydrocarbon analysis in mineral oils,36 and the study of mitochondria.37 In the present experiment, the synchronous luminescence spectra of Cu NCs (method-II) exhibit two peaks in aqueous solution, intense one at 468 nm and relatively weak at 530 nm (Fig. 2A). Excitation spectra (recorded at λems = 600 nm) is relatively broad with a peak at 510 nm and excitation dependent emission spectra exhibits peak maxima at 570 nm (Fig. 2B). The simple incubation method yields blue-luminescent copper nanoclusters whereas sonication method results the formation of cyan-luminescent Cu NCs (major) and green-luminescent Cu NCs (minor). The quantum yields of the synthesized Cu NCs in aqueous solution are estimated to be 0.23 for blue-luminescent Cu NCs, 0.24 and 0.07 for cyan-luminescent Cu NCs and green-luminescent Cu NCs respectively, using coumarin 102 as the reference.38
 |
| | Fig. 2 [A] Synchronous luminescence spectra of – (i) Cu NCs (method-II) in aqueous solution, (ii) Cu NCs (method-II) in methanolic solution, Δλ = 20 nm. [B] Normalized excitation at λems = 600 nm (dashed line) and excitation-dependent emission spectra of the Cu34 NC–cellulase in aqueous solution at λexc: (i) 500 nm, (ii) 510 nm, (iii) 520 nm, and (iv) 530 nm. | |
The atomic composition of synthesized NCs can effectively be deduced from the Jellium model, =(λems/λFermi)3, even the predicted composition is directly matched with mass spectrometry results using MALDI-TOF (biosystems) spectrometer.17,39,40 For blue-luminescent Cu NCs, (Eem = 401 nm and EFermi of Cu, which is 177.12 nm (7 eV)) the number of atoms, N, present in Cu NCs is evaluated to be ∼12 (11.6 ± 0.1), the most stable (magic numbered) atomic composition of the blue-luminescent Cu12 NCs–cellulase. Whereas for cyan-luminescent Cu NCs (Eem = 484 nm), and green-luminescent Cu NCs (Eem = 570 nm) the number of atoms, N, present in Cu NCs is evaluated to be ∼20 (20.4 ± 0.2) and ∼34 (33.7 ± 0.5) respectively. The experimental error in determination of emission maxima (swallow in nature) is also mentioned in the prediction of atomic composition of nanoclusters. Interestingly, both Cu20 NC and Cu34 NC are also the magic numbered copper nanoclusters. To the best of our knowledge this is the first report where an enzyme synthesizes two nearby magic numbered clusters. The transmission electron microscopy (TEM) is also employed to confirm the formation of copper nanoclusters. Typical TEM images of the synthesized red luminescent Cu NCs reveals the formation of the NCs has indeed taken place with a size regime of ∼4 nm. The reason for relatively large sized in TEM might be due to the evaporation of water which lead to aggregation of Cu NCs–cellulase (Fig. S6, ESI†).
Unprecedentedly, the synchronous luminescence spectra of Cu NCs (method-II) in methanolic medium, exhibit the complete reversal of PL intensities, the peak at 530 nm becomes the strong one whereas intensity of 468 nm peak drops dramatically. Such reversal might happen on two accounts: first, the better stabilization of relatively large-sized nanoclusters over smaller one due to solvent-induced structural rearrangement of protein; another, energy transfer from cyan-fluorescent Cu NCs to a green-luminescent one. In order to substantiate the above proposed hypotheses, we have carried out picosecond time-resolved measurements for synthesized Cu NCs. Fig. 3 shows the luminescence decay traces of Cu NCs in aqueous and methanolic media. The decay traces are well fitted with triexponential (Table 1). Since the decays are multiexponential, therefore, then it is pertinent to use an average decay time, which is proportional to the steady state intensity.28 An analysis of the decay traces of Cu20 NCs–cellulase reveals the presence of three components in aqueous solution: a shorter one with a lifetime of 0.38 ns (τ1) and two longer components (τ2 ∼ 2.82 ns and τ3 ∼ 4.67 ns) (Table 1). The lifetimes and its corresponding weightage of Cu20 NCs–cellulase remain unchanged in methanolic media. Interestingly, the decay traces recorded at λems = 600 nm (Cu34 NC–cellulase) in aqueous and methanolic media differ significantly. The amplitude of short component decreases while the same for longer component increases. However, the absence of rise-time bolsters the first hypothesis, the better scaffolding ability of cellulase in methanolic media for green-fluorescent nanoclusters. The time-resolved trace of Cu12 NCs–cellulase in aqueous medium, as expected, exhibits an average luminescent lifetime of ∼3.16 ns [0.1 ns (5%), 2.22 ns (68%); 6.10 ns (27%)] (Fig. S7, ESI†). The origin of multiexponential decay of Cu NC can be rationalized by the electronic transition between the filled d orbital and sp conduction bands.16 The shorter component may be attributed to Cu NCs–solvent interaction, whereas relatively longer components to the Cu–Cu interaction and the transition between Cu–amino acid residues in enzyme.
 |
| | Fig. 3 Photoluminescence decays of Cu NCs (method-II) in – [A] aqueous solution and [B] methanolic solution. (i) λems = 485 nm and (ii) λems = 600 nm. The IRF (λexc = 440 nm) is shown in the dashed line. | |
Table 1 PL decay parameters of Cu NCs–cellulase obtained in method-IIa
| |
λems = 485 nm |
λems = 600 nm |
| a1 |
a2 |
a3 |
τ1 |
τ2 |
τ3 |
〈τ〉 |
χ2 |
a1 |
a2 |
a3 |
τ1 |
τ2 |
τ3 |
〈τ〉 |
χ2 |
| Luminescence lifetimes, τi are in ns and 〈τ〉 = ∑aiτi2/∑aiτi. |
| In aqueous |
0.03 |
0.80 |
0.17 |
0.38 |
2.82 |
4.67 |
3.29 |
1.10 |
0.15 |
0.51 |
0.34 |
0.33 |
2.62 |
4.12 |
3.33 |
0.97 |
| In methanol |
0.02 |
0.79 |
0.19 |
0.33 |
3.08 |
4.60 |
3.47 |
1.13 |
0.05 |
0.89 |
0.06 |
0.30 |
3.75 |
6.82 |
4.07 |
1.25 |
| 20 μM Zn(II) |
0.03 |
0.72 |
0.25 |
0.33 |
2.82 |
4.71 |
3.50 |
1.17 |
0.23 |
0.77 |
|
0.37 |
3.34 |
|
3.24 |
0.96 |
| 20 μM Cd(II) |
0.03 |
0.80 |
0.17 |
0.32 |
2.81 |
4.72 |
3.30 |
1.09 |
0.170 |
0.82 |
|
0.37 |
3.22 |
|
3.15 |
1.01 |
| 20 μM Hg(II) |
0.03 |
0.68 |
0.29 |
0.29 |
2.67 |
4.21 |
3.28 |
1.08 |
0.18 |
0.82 |
|
0.35 |
3.22 |
|
3.15 |
0.99 |
| 20 μM Fe(III) |
0.04 |
0.84 |
0.12 |
0.34 |
2.86 |
5.01 |
3.28 |
1.11 |
0.19 |
0.81 |
|
0.35 |
3.41 |
|
3.33 |
1.14 |
Temperature dependent luminescence properties of Cu NCs
In order to explore the temperature dependence on the luminescent properties of synthesized Cu NCs, we have recorded luminescence spectra of nanoclusters in both aqueous and methanolic media as a function of temperature. With increase in temperature, the PL intensity of the NCs–cellulase in aqueous media is quenched, whereas enhanced luminescence is noticed in methanolic media (Fig. 4A). Expectedly, in aqueous media, Cu20 NCs–cellulase follows the similar luminescence quenching trend like Cu12 NCs–cellulase (Fig. S8A, ESI†). The quenching mechanism can be understood by the thermal unfolding of stabilizing scaffolds of cellulase. Surprisingly, in methanolic media, PL intensity at 465 nm (Cu20 NCs–cellulase) is getting quenched whereas the PL intensity at 540 nm is getting enhanced with increasing temperature (Fig. 4B). To get an insight on the heterogeneity present in methanolic media, synchronous luminescence study has also been carried out, which reveals that with increasing temperature, the peak at 530 nm (Cu34 NCs–cellulase) is initially blue-shifted and then undergoes spectral broadening (Fig. 4B, inset). Interestingly, it also reveals that in aqueous media the PL intensity of (Cu20 NCs–cellulase) is decreasing whereas the PL intensity of Cu34 NCs–cellulase is increasing with rise in temperature (Fig. S9B, ESI†). The observation of luminescence enhancement may be rationalized by the solvent/temperature-triggered preferential scaffolding ability of cellulase for large-sized nanoclusters through redistribution of secondary structure (greater content of random coil) of enzyme by strengthening the hydrogen-bonding interactions in the backbone.41 However, the above results demonstrate that photoluminescence of as-synthesized Cu NCs–cellulase are not only temperature dependent, solvent responsive as well. This unique ability of Cu NCs–cellulase can be highly advantageous for imaging of heterogeneous systems (varied polarity), and for probing the temperature as well. Therefore, as-synthesized luminescent nanoclusters may find applications in the biomedical field and nanophotonics.
 |
| | Fig. 4 [A] Plot of the PL intensity of Cu12 NCs–cellulase in aqueous (filled circle) and in methanolic media (filled square) as a function of temperature. [B] Emission spectra of Cu NCs (method-II) in methanolic medium as a function of temperature (λexc = 440 nm). Inset contains the synchronous luminescence spectra of the same mentioned above (Δλ = 20 nm). | |
Role of the random-coil structure of cellulase
The scaffolding ability of protein is dictated by its secondary structure that plays a central role in biological applications. To get an insight on the post-synthesis structural modification of cellulase, we have performed circular dichroism (CD) studies. The far-UV CD spectra of cellulase alone and Cu NCs are shown in Fig. 5A. The presence of characteristic double minima suggests the existence of α-helical structures in enzyme. However, the CD spectra of nanoclusters stabilized cellulase are remarkably different and indicate total disruption of the helical structure of protein. The random coil structure of cellulase gets significantly populated in order to stabilize the copper nanoclusters. The greater the size of the nanoclusters, higher is the content of random coil. It is quite evident that the random coil structure of cellulase dictates the core-size of the scaffold which promotes preferential stabilization of large-sized nanoclusters over smaller one. Methanol is known to weaken the hydrophobic interactions and thereby resulting in an expansion of the hydrophobic core; at the same time strengthens the interactions between the polar backbone of the protein. Circular dichroism study therefore bolsters again the first hypothesis.
 |
| | Fig. 5 [A] Far-UV CD spectra of – (i) cellulase, (ii) Cu12 NC–cellulase in aqueous solution, (iii) Cu NCs–cellulase (method-II) in aqueous solution and (iv) Cu NC–cellulase (method-II) in methanolic medium. [B] Extent of luminescence enhancement of Cu20 NC–cellulase in presence of metal ions: I0 and I are the luminescence intensities in the absence and presence of metal ions, respectively, in aqueous solution containing metal ions – (i) Zn2+ ions (filled square), (ii) Cd2+ ions (hollow square), (iii) Hg2+ ions (filled circle) and (iv) Fe3+ ions (hollow circle). | |
Metal ions are known to stabilize of secondary structure of protein by interacting with carbonyl and amide groups as well as by cross-linking the sidechains.42–44 “Soft” ions such as Zn(II), Cd(II) and Hg(II), possessing high polarizability and a hard ion, Fe(III) having high positive charge are primarily chosen for this study. “Soft” ions form stable bonds with polarizable ligands such as cysteine thiolates and histidine side whereas hard ions promote electrostatic interactions.45,46 In this study, aqueous solutions of metal salts, Fe(III), Zn(II), Cd(II), and Hg(II) are gradually added to the solution of the Cu20 NCs–cellulase and stepwise change in luminescence intensity (at 485 nm) of the Cu20 NC–cellulase is shown in Fig. 5B and luminescence spectra are presented in Fig. S10.† The augmentation of luminescence intensity to 100% is observed for all the metal ions upto 40 μM and further addition of Fe(III) ions to signals “turn off” behavior. Interestingly, rapid enhancement of luminescence intensities is observed on the addition of “soft ions”, about 300% enhancement for 60 μM of Zn(II) and Cd(II) ions. The most probable explanation for enhancement of PL of Cu20 NC–cellulase in the presence of ions in aqueous medium is that all of them stabilize the scaffolds of the enzyme through bonding interaction. Time-resolved studies at 20 μM concentration of metal ions indicate that the average lifetime nanoclusters remain unchanged (Table 1). The fluorescence enhancement is therefore the ground state dynamics of enzyme nanocluster complex. However, at higher concentration, the soft ions (d10 systems) continue to stabilize the structure, whereas Fe(III) ion (d5 system) may destabilize the scaffold by induction of strong ionic interaction responsible for quenching of luminescence. Moreover, augmenting luminescence intensity is of vital importance to the development of imaging techniques and illumination sources.
Conclusion
In conclusion, magic numbered, highly luminescent Cu NCs, each of which possesses unique luminescent properties has been synthesized by cellulase alone. Upon incubation of enzyme–metal ion solution, the resultant nanoclusters turns out be blue-luminescent one with emission peak at 401 nm, whereas sonication results the formation of two nanoclusters: cyan-luminescent which emits strongly 485 nm and green luminescent having emission maxima at 570 nm in aqueous solution. Sonication of enzyme solution perhaps increases the inner-diameter of the scaffold to accommodate large-sized nanoclusters. The Cu NCs show high photoluminescence quantum yield and better colloidal stability. Solvent-induced conformational change of cellulase play a significant role to reverse the temperature-dependent luminescent properties of Cu12 NCs–cellulase. Moreover, the preferential stabilization of large-sized Cu34 NCs–cellulase over a smaller one Cu20 NCs–cellulase is found to be dependent on both temperatures as well as solvent. The content of the random coil structure of enzyme is the governing parameter to stabilize the different-sized copper nanoclusters. The addition of “soft ions” results remarkable enhancement of the luminescent intensity of nanoclusters whereas the presence of hard ion, Fe3+ quenches the luminescence. The as-synthesized nanoclusters may find profound applications in bioimaging, nanophotonics and interdisciplinary fields. However, it will be of great interest to study the energy transfer between two nanoclusters, which is currently underway in our group.
Acknowledgements
The authors greatly acknowledge RGIPT-Rae Bareli and the Science and Engineering Research Board, Department of Science and Technology, New Delhi for financial support (DST Project No. SB/FT/CS-135/2013). AS and TR thank RGIPT-Rae Bareli for assistantships. Authors sincerely thank Prof. Pratik Sen, IIT Kanpur for time-resolved studies. Authors also thank SAIF-IIT Bombay for TEM data and Tuhin Khan, Anjali Dhir for their help in experiments.
References
- J. Zheng, P. R. Nicovich and R. M. Dickson, Annu. Rev. Phys. Chem., 2007, 58, 409–431 CrossRef CAS PubMed.
- J. Oliver-Meseguer, J. R. Cabrero-Antonino, I. Domínguez, A. Leyva-Pérez and A. Corma, Science, 2012, 338, 1452–1455 CrossRef CAS PubMed.
- H. Zhang, T. Watanabe, M. Okumura, M. Haruta and N. Toshima, Nat. Mater., 2012, 11, 49–52 CrossRef PubMed.
- T.-H. Lee and R. M. Dickson, Opt. Photonics News, 2004, 15, 22–27 CrossRef CAS.
- S. Chattoraj and K. Bhattacharyya, J. Phys. Chem. C, 2014, 118, 22339–22346 CAS.
- J. Zhang, L. Tu, S. Zhao, G. Liu, Y. Wang, Y. Wang and Z. Yue, Biosens. Bioelectron., 2015, 67, 296–302 CrossRef CAS PubMed.
- S. Choi, R. M. Dickson and J. Yu, Chem. Soc. Rev., 2012, 41, 1867–1891 RSC.
- Z. Luo, K. Zheng and J. Xie, Chem. Commun., 2014, 50, 5143–5155 RSC.
- H. Qian, M. Zhu, Z. Wu and R. Jin, Acc. Chem. Res., 2012, 45, 1470–1479 CrossRef CAS PubMed.
- M. Brack, Rev. Mod. Phys., 1993, 65, 677–732 CrossRef CAS.
- W. A. de Heer, Rev. Mod. Phys., 1993, 65, 611–676 CrossRef CAS.
- H. Tsunoyam and T. Tsukuda, J. Am. Chem. Soc., 2009, 131, 18216–18217 CrossRef PubMed.
- S. Kumar, M. D. Bolan and T. P. Bigioni, J. Am. Chem. Soc., 2010, 132, 13141–13143 CrossRef CAS PubMed.
- J. Xie, Y. Zheng and J. Y. Ying, J. Am. Chem. Soc., 2009, 131, 888–889 CrossRef CAS PubMed.
- A. Mathew, P. R. Sajanlal and T. Pradeep, J. Mater. Chem., 2011, 21, 11205–11212 RSC.
- N. Goswami, A. Giri, M. S. Bootharaju, P. L. Xavier, T. Pradeep and S. K. Pal, Anal. Chem., 2011, 83, 9676–9680 CrossRef CAS PubMed.
- C. Wang, C. Wang, L. Xu, H. Cheng, Q. Lin and C. Zhang, Nanoscale, 2014, 6, 1775–1781 RSC.
- S. Ghosh, N. K. Das, U. Anand and S. Mukherjee, J. Phys. Chem. Lett., 2015, 6, 1293–1298 CrossRef CAS PubMed.
- R. Ghosh, A. K. Sahoo, S. S. Ghosh, A. Paul and A. Chattopadhyay, ACS Appl. Mater. Interfaces, 2014, 6, 3822–3828 CAS.
- Y. Ling, J. J. Wu, Z. F. Gao, N. B. Li and H. Q. Luo, J. Phys. Chem. C, 2015, 119, 27173–27177 CAS.
- X. Gao, Y. Lu, M. Liu, S. He and W. Chen, J. Mater. Chem. C, 2015, 3, 4050–4056 RSC.
- C. Wang, L. Ling, Y. Yao and Q. Song, Nano Res., 2015, 8, 1975–1986 CrossRef CAS.
- C. P. Joshi, D. Panda, D. J. Martell, N. M. Andoy, T.-Y. Chen, A. Gaballa, J. D. Helmann and P. Chen, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 15121–15126 CrossRef CAS PubMed.
- H. Tapiero, D. M. Townsend and K. D. Tew, Biomed. Pharmacother., 2003, 57, 386–398 CrossRef CAS.
- T. Rai and D. Panda, Chem. Phys. Lett., 2015, 623, 108–112 CrossRef CAS.
- J. Karlsson, M. Saloheimo, M. Siika-aho, M. Tenkanen, M. Penttilae and F. Tjerneld, Eur. J. Biochem., 2001, 268, 6498–6507 CrossRef CAS PubMed.
- P. K. Foreman, D. Brown, L. Dankmeyer, R. Dean, S. Diener, N. S. Dunn-Coleman, F. Goedegebuur, T. D. Houfek, G. J. England, A. S. Kelley, H. J. Meerman, T. Mitchell, C. Mitchinson, H. A. Olivares, P. J. M. Teunissen, J. Yao and M. Ward, J. Biol. Chem., 2003, 278, 31988–31997 CrossRef PubMed.
- J. R. Lakowicz, Principles of Fluorescence Spectroscopy, Springer, 3rd edn, 2011 Search PubMed.
- S. Rafiq, R. Yadav and P. Sen, J. Phys. Chem. A, 2011, 115, 8335–8343 CrossRef CAS PubMed.
- S. Banerjee and A. Datta, Langmuir, 2010, 26, 1172–1176 CrossRef CAS PubMed.
- M. R. DeFelippis, C. P. Murthy, F. Broitman, D. Weinraub, M. Faraggi and M. H. Klapper, J. Phys. Chem., 1991, 95, 3416–3419 CrossRef CAS.
- J. J. Warren, J. R. Winkler and H. B. Gray, FEBS Lett., 2012, 586, 596–602 CrossRef CAS PubMed.
- S. K. Das, J. Liang, M. Schmidt, F. Laffir and E. Marsili, ACS Nano, 2012, 6, 6165–6173 CrossRef CAS PubMed.
- D. Panda, S. Khatua and A. Datta, J. Phys. Chem. B, 2007, 111, 1648–1656 CrossRef CAS PubMed.
- D. Panda and A. Datta, J. Chem. Sci., 2007, 119, 99–104 CrossRef CAS.
- D. Patra and A. K. Mishra, Analyst, 2000, 125, 1383–1386 RSC.
- M. D. F. Askari and T. Vo-Dinh, Biopolymers, 2004, 73, 510–523 CrossRef CAS PubMed.
- K. Rurack and M. Spieles, Anal. Chem., 2011, 83, 1232–1242 CrossRef CAS PubMed.
- S. Chattoraj, M. Amin, S. Mohapatra, S. Ghosh and K. Bhattacharyya, ChemPhysChem, 2016, 17, 61–68 CrossRef CAS PubMed.
- N. K. Das, S. Ghosh, A. Priya, S. Datta and S. Mukherjee, J. Phys. Chem. C, 2015, 119, 24657–24664 CAS.
- S. Hwang, Q. Shao, H. Williams, C. Hilty and Y. Q. Gao, J. Phys. Chem. B, 2011, 115, 6653–6660 CrossRef CAS PubMed.
- F. Yu, V. M. Cangelosi, M. L. Zastrow, M. Tegoni, J. S. Plegaria, A. G. Tebo, C. S. Mocny, L. Ruckthong, H. Qayyum and V. L. Pecoraro, Chem. Rev., 2014, 114, 3495–3578 CrossRef CAS PubMed.
- M. Luczkowski, R. D. Ricco, M. Stachura, S. Potocki, L. Hemmingsen and D. Valensin, Metallomics, 2015, 7, 478–490 RSC.
- A. J. Doerr and G. L. McLendon, Inorg. Chem., 2004, 43, 7916–7925 CrossRef CAS PubMed.
- M. Bortolato, F. Besson and B. Roux, Proteins, 1999, 37, 310–318 CrossRef CAS.
- P. Hainaut and J. Milner, Cancer Res., 1993, 53, 1739–1742 CAS.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra09763d |
|
| This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.