Sunteți pe pagina 1din 14

Nanoscale

PAPER

Gold nanoclusters with bright near-infrared


Cite this: Nanoscale, 2018, 10, 3792
photoluminescence†
Goutam Pramanik,a Jana Humpolickova,a Jan Valenta,b Paromita Kundu,c Sara Bals,c
Petr Bour, a Martin Dracinsky a and Petr Cigler *a

The increase in nonradiative pathways with decreasing emission energy reduces the luminescence
quantum yield (QY) of near-infrared photoluminescent (NIR PL) metal nanoclusters. Efficient surface
ligand chemistry can significantly improve the luminescence QY of NIR PL metal nanoclusters. In contrast
to the widely reported but modestly effective thiolate ligand-to-metal core charge transfer, we show that
metal-to-ligand charge transfer (MLCT) can be used to greatly enhance the luminescence QY of NIR PL
gold nanoclusters (AuNCs). We synthesized water-soluble and colloidally stable NIR PL AuNCs with
unprecedentedly high QY (∼25%) upon introduction of triphenylphosphonium moieties into the surface
Received 15th August 2017, capping layer. By using a combination of spectroscopic and theoretical methods, we provide evidence for
Accepted 8th January 2018
gold core-to-ligand charge transfer occurring in AuNCs. We envision that this work can stimulate the
DOI: 10.1039/c7nr06050e development of these unusually bright AuNCs for promising optoelectronic, bioimaging, and other
rsc.li/nanoscale applications.

Introduction cent (NIR PL) AuNCs functionalized with a variety of thiol-con-


taining ligands, including tiopronin,24 thioctic (lipoic)
Photoluminescent nanosystems have garnered great interest acid,25–27 polymers,28 11-mercaptoundecanoic acid,29 zwitter-
due to their excellent optical properties, low cost, photostabil- ionic structures,30 glutathione,31–33 mercaptosuccinic acid,16
ity, convenience, and highly sensitive detection.1 Recently, and proteins.12,34–36
ultrasmall (<2 nm) photoluminescent gold nanoclusters The increase in nonradiative pathways with decreasing
(AuNCs) have been identified as promising candidates for cell energy of emitted light (i.e., the “energy gap law”)37 makes it
labeling, biosensing,2 photo-therapy applications,3 cancer extremely challenging to prepare AuNCs that emit in the NIR
radiotherapy,4,5 and antimicrobial agents.6 Photoluminescent region with high quantum yields (QYs). The photo-
AuNCs are biocompatible and readily bioconjugable, and they luminescence QY of AuNCs can be enhanced via charge trans-
show good photostability and low toxicity.7 Photoluminescent fer from the ligands to the metal core (i.e., LMCT) through the
AuNCs can emit from the blue to the near-IR (NIR) spectral Au–S bonds and is parallel with the ligand’s capability of
region depending on the number of atoms within the cluster.8 donating electron density to the metal core through the S–Au
In recent years, various approaches have been developed to bond (i.e., charge transfer capability of the ligand).38 The
prepare highly photoluminescent AuNCs with emission in the luminescence of AuNCs can also be “turned on” upon intro-
UV-vis region, for example: (1) engineering the particle surface duction of sulfur-containing ligands39 or enhanced by either
by using different ligands, such as dendrimers,9 DNA,10,11 pep- sulfur oxidation at the Au–ligand interface40 or electronic
tides, and proteins;12–16 (2) controlling the metal core size;17 polarization of the bonds between the Au core and thiolate
(3) aggregation-induced emission;18–21 and (4) rigidification of ligands.41 Nevertheless, the currently known NIR-emitting
the Au(I)-thiolate shell.22,23 Recent advances in the synthesis of AuNCs achieve relatively low QYs.
AuNCs have enabled the development of NIR-photolumines- One promising approach to a great increase in QY is based
on an efficient but experimentally laborious doping of the Au
core with a precise number of Ag atoms.3 In contrast, ligand-
a
Institute of Organic Chemistry and Biochemistry of the CAS, Flemingovo nam. 2, based PL enhancement approaches are more general and syn-
166 10 Prague 6, Czech Republic. E-mail: cigler@uochb.cas.cz thetically straightforward and can directly produce water-
b
Department of Chemical Physics and Optics, Faculty of Mathematics and Physics, soluble AuNCs. In this report, we present results from our
Charles University, Ke Karlovu 3, 121 16 Prague 2, Czech Republic
c study of ligand-mediated improvement of the NIR lumine-
EMAT, University of Antwerp, Groenenborgerlaan 171, B-2020 Antwerp, Belgium
† Electronic supplementary information (ESI) available. See DOI: 10.1039/ scence efficiency of AuNCs. We found that the weak emission
c7nr06050e of AuNCs stabilized with thioctic acid (TA) and polyethylene

3792 | Nanoscale, 2018, 10, 3792–3798 This journal is © The Royal Society of Chemistry 2018
Nanoscale Paper

glycol (PEG) can be enhanced dramatically by introducing tri- be further functionalized. However, TA-protected AuNCs are
phenylphosphonium cations in the capping layer. Motivated unstable in acidic solutions due to the protonation of the car-
by this finding, we investigated in detail the photophysical pro- boxylate group. Hence after the preparation of TA-protected
perties of these NIR PL AuNCs. We present experimental evi- AuNCs using a one-step reduction procedure in water,25 we
dence suggesting the enhancement of the luminescence QY subsequently functionalized them with thiol-terminated PEG
via charge transfer interactions from the gold core to the tri- (MW 2000), yielding AuNC 1. This led to the improvement of
phenylphosphonium cations, i.e. by metal-to-ligand charge aqueous solubility of AuNCs in acidic pH and a further
transfer (MLCT). increase of their colloidal stability. Then we covalently modi-
fied 1 with 3-(aminopropyl)triphenylphosphonium (TPP) salts
via amidic coupling to surface carboxyl groups, thereby obtain-
Results and discussion ing AuNC 2 (Fig. 1a). We characterized the AuNC structure by
high-resolution transmission electron microscopy (HRTEM),
To synthesize NIR PL AuNCs, we have utilized a bottom-up inductively coupled plasma optical emission spectrometry
approach where the gold precursor is first treated with suitable (ICP-OES), 31P and 1H NMR, and zeta potential measurements.
ligands, followed by the reduction of the gold. Among the HRTEM revealed the formation of monodisperse AuNCs
various ligands known for the stabilization of AuNCs, multi- (Fig. 2). AuNCs 1 and 2 both showed a narrow metal core size
thiol-based ligands (with at least two coordinating groups) distribution with an average diameter of 1.15 ± 0.2 nm, as
provide enhanced colloidal stability to the AuNCs because of judged from the image analysis of more than 100 individual
the increased number of binding sites between each ligand particles. The 31P NMR spectrum of 2 (Fig. S3a†) showed a
and the AuNC surface.30 For primary modification of NIR PL single peak at δ = 23.7 ppm, which is similar to that of the
AuNCs, we utilized the bidentate thiol anchor TA providing a phosphonium cation in 3-(aminopropyl)triphenylphospho-
remarkable colloidal stability to the AuNCs over a broad range nium bromide (Fig. S3b†). The 31P NMR spectrum of 2 also
of stringent conditions compared to simple monothiol indicated that the triphenylphosphonium cation was neither
ligands.42 TA also provides a one-phase growth route to nano- degraded nor oxidized during the synthesis. The integration of
particles with discrete size control in water and offers a car- signal intensities in 1H NMR of 2 (Fig. S6 in ESI†) provided us
boxylic acid group on the surface of nanoparticles, which can the molar ratios PEG : TA ≈ 5 : 2 and TA : TPP ≈ 2 : 1. If we

Fig. 1 (a) Schematic representation of the structure and preparation of NIR PL AuNCs 1 and 2. X− = Br−, I−, Cl−, or BF4−. (b) Normalized absorption
spectra of 200 μg mL−1 aqueous solutions of 1 and 2 (X− = BF4−). Inset: Photograph of the solutions under white light. (c) Normalized emission
spectra of 200 μg mL−1 aqueous solutions of 1 and 2 (X− = BF4−) collected upon excitation at 365 nm. Inset: Photograph of the solutions upon UV
illumination (365 nm). (d) Time-resolved photoluminescence decay curves for 1 and 2. Inset: Plot of the emission rates (k) of 1 and 2 (X− = BF4−) as a
function of reciprocal temperature (1000/T). Fits are based on eqn (2) (note the rate scale is logarithmic with base e).

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 3792–3798 | 3793
Paper Nanoscale

(Fig. 1c inset). We observed that the PL intensity depends on


the Au : P ratio (estimated using ICP OES), with a maximum
enhancement at Au : P ≈ 8 : 1 (Fig. S2 in the ESI†). Considering
the approximate size of the nanocluster Au25 (based on the
size obtained from HRTEM and on the PL emission maxima),
this value suggests ∼3 TPP moieties per one AuNC. This result
is fairly consistent with the 2 TPPs per AuNC estimated from
1
H NMR (see above).
The time-resolved PL decay profiles of the two sets of
AuNCs are shown in Fig. 1d. The PL lifetimes were extracted
using a three exponential fit from the PL decay curves using
the equation

IðtÞ ¼ A1 e t=τ1 þ A2 e t=τ2 þ A3 e t=τ3 þ A4 ; ð1Þ

where t is the time and Ai is the amplitude associated with


each decay time, τi. A4 represents processes that occur on a
longer time scale than addressed by the experiment. Table 1
lists the individual components extracted from the fit for each
AuNC. We attribute the major component τ1 (∼10 ns) to emis-
sion from singlet excited states of Au nanoclusters.8,10,44 The
Fig. 2 Bright field TEM images of 1 (a) and 2 (b) and the corresponding long life-time components (>100 ns) are characteristics of
HRTEM images of a single AuNC 1 (c) and 2 (d) showing the typical size Au(I)–thiol complexes, suggesting that such complexes contrib-
around approximately 1 nm. The insets of HRTEM images show the FFTs
from the respective particle in the [011] zone where the spots corres-
ute to the luminescence of AuNCs due to the ligand–metal
pond to {111} and {200} lattice planes of Au. charge transfer and Au(I)–Au(I) interactions.45,46 Upon conju-
gation with TPP, the amplitude of τ1 decreases, while those of
τ2 and τ3 (∼0.45 μs and 2.09 μs, respectively) increase greatly.
The pronounced long lifetimes for NIR PL can be attributed to
approximate the size of the nanocluster roughly as Au25 (based the luminescence arising from gold shell states in which
on the size obtained from HRTEM and on the PL emission significant ligand contribution is present.36,38,47
maxima), we obtain 18 attachment positions for thiols. Using To further investigate the origin of luminescence enhance-
the above stated ratios, we obtain approximately 10 PEG, 4 TA ment in 2, we carried out temperature-dependent lumine-
(dithiol ligand), and 2 TPP moieties attached to each AuNC 2. scence measurements. The inset of Fig. 1d presents the emis-
The zeta potentials of 1 and 2 in phosphate buffer (10 mM, sion rate k (inverse luminescence decay time, k = τPL−1) as a
pH 7.4) were −10.0 ± 1.8 mV and +1.6 ± 0.3 mV, respectively. function of 1000/T. We fit the observed temperature depen-
The higher zeta potential of 2 compared to 1 is consistent with dence with the equation
a charge switch from a part of negatively charged carboxylic
groups to positively charged TPP. kðTÞ ¼ k0 þ k′e ΔE=ðkB TÞ ; ð2Þ
As expected, unlike the UV-vis absorption spectra of larger
where k0, k′, ΔE, and kB are the temperature independent emis-
Au nanoparticles, the spectra of the AuNCs did not display a
sion rate, frequency factor, energy barrier of non-radiative
strong surface plasmon resonance around 520 nm (Fig. 1b).
decay, and Boltzmann constant, respectively. The best para-
The AuNC solutions are therefore only light yellow in color
meters obtained by fitting the data for 1 and 2 are presented
(inset Fig. 1b). However, the enhanced absorption of 2 com-
in Table 2. Our experiments assessing the PL lifetime and
pared to 1 in the UV region revealed a strong interaction
temperature dependence of decay rates (measured from
between the gold core and TPP (Fig. 1b). This enhancement of
5–60 °C) provided parameters characteristic for MLCT, which
absorption can be attributed to the charge transfer from the
has been well-documented in ruthenium complexes.48,49
metallic core to the ligands, as described by Sementa et al. for
other highly delocalized sterically hindered ligands present on
AuNCs.43 Table 1 The amplitude weighted lifetimes of the PL decay curves
Both 1 and 2 showed broad emission with the maxima (Fig. 1d) for 1 and 2 fitted with a three-exponential function. The corres-
around 750 nm (Fig. 1c). However, upon introduction of TPP, ponding relative amplitude (A) for each τ is indicated in parentheses
the PL intensity dramatically increased (Fig. 1c). The observed
τ1 (ns) τ2 (ns) τ3 (μs) τ4 (μs)
change in the PL intensity corresponds to the increase in PL
Set of τ used ∼10 ∼457 ∼2.09 >10
QYs from 10% to 25% estimated for 1 and 2 (X− = BF4−), for the fit
respectively (Fig. S1 and Table S1 in the ESI†). The PL increase A1 A2 A3 A4
is visible to the unaided eye, as documented by photographs AuNC 1 0.79 0.12 0.03 0.06
of the AuNC solutions upon UV illumination (λex = 365 nm) AuNC 2 0.57 0.23 0.17 0.03

3794 | Nanoscale, 2018, 10, 3792–3798 This journal is © The Royal Society of Chemistry 2018
Nanoscale Paper

Table 2 The best parameters obtained by fitting the data from tem- Because charge transfer involving the triphenylphospho-
perature-dependent luminescence measurements for 1 and 2 using nium cation is also influenced by the counteranion,51 we
eqn (2)
investigated the effect of phosphonium salt counteranions on
the emission properties of 2. We prepared AuNC 2 variants
k0 (MHz) k′ (MHz) ΔE (meV)
bearing a set of four different counteranions (BF4−, Cl−, Br−,
AuNC 1 0.07 22 92 and I−) by reactions of the corresponding triphenylphospho-
AuNC 2 0.25 61 144
nium salts. The anions differ in oxidizability,51 forming a
series (BF4− > Cl− > Br− > I−) ranging from the nonoxidizable
BF4− anion to the most oxidizable I−. As Fig. 4b shows, the PL
Based on our HRTEM measurements (Fig. 2), we ruled out the
intensity depends on the counteranion. Decreases in the emis-
possibility of aggregation-induced PL enhancement20 of 2,
sion intensity of 2 follow the increasing oxidizability of the
because AuNC aggregates were absent in the samples.
counteranions. This trend can be interpreted according to the
To further investigate the Au core to phosphonium cation
proposed mechanism of PL enhancement (Fig. 4a). The charge
charge transfer, we employed X-ray photoelectron spectroscopy
transfer from the counteranion to the phosphonium cation
(XPS). The binding energy (BE) shift of Au 4f band is influ-
decreases in the order I− > Br− > Cl− > BF4− and competes with
enced by the oxidation state of gold. The BE of Au 4f7/2 tran-
the gold core-to-phosphonium ion charge transfer (i.e.,
sition of both 1 and 2 (Fig. 3) falls between the energies of Au
MLCT). The more readily oxidizable the counteranion, the
(0) (84 eV) and Au(I) (86 eV) of gold thiolate, suggesting the
more charge transfer takes place from the counteranion to
coexistence of Au(0) and Au(I) in both AuNCs.44,50 However,
phosphonium cation and the less gold core-to-phosphonium
upon conjugation of TPP, the BE of both Au 4f7/2 and 4f5/2
cation charge transfer occurs (and vice versa). These obser-
shifts towards higher values. Because both the AuNCs have a
vations further support the involvement of the MLCT mecha-
similar size and were stored under similar conditions without
nism in the enhancement of the PL emission of 2.
exposure to oxidizing agents, we attribute the BE shift of Au 4f
Because the intrinsic hydrophobicity of the TPP moiety con-
band to charge transfer from the Au core to phosphonium
taining three phenyl groups can be ( partially) responsible for
cations. The charge transfer makes Au atoms in the cluster
the enhancement of QY due to the increased local hydrophobi-
more positively charged, which in turn increases their core
city in 2, we performed two different experiments investigating
level binding energy. The XPS result of the Au 4f binding
energy shift provides direct experimental evidence for charge
transfer from the Au core to the phosphonium cation (Fig. 3).

Fig. 3 Recorded XPS spectra of the Au 4f peaks (black line) of (a) 1 and Fig. 4 (a) Proposed luminescence enhancement mechanism in 2.
(b) 2. The fitted curve (blue) shows that the spectra comprise two doub- (b) The relative photoluminescence spectra of 2 with different countera-
lets of Au(0) (dark red) and Au(I) (green). nions collected upon excitation at 365 nm.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 3792–3798 | 3795
Paper Nanoscale

this potential environmental effect. First, we introduced a


model hydrophobic moiety 2-aminoanthracene to 1, providing
3 (Scheme S1 in the ESI†). The characteristic intrinsic fluo-
rescence of 2-aminoanthracene in the blue region of the spectra
indicates the presence of the moiety on 3. However, in contrast to
1 and 2, the AuNC 3 did not show any PL peak in the wavelength
range of 600–850 nm (same concentration of all the AuNCs;
Fig. S4 in the ESI†), which would be expected upon an increase of
local hydrophobicity. Second, we recorded the PL spectra of 2 in
binary mixtures of water and dimethyl sulfoxide (DMSO) with
increasing volume fractions of DMSO. DMSO is a known disrup-
tor of hydrogen-bond network in water52 and its presence can
strongly modulate the local polarity around fluorophores and
influence their QYs.53 Nevertheless, we did not observe any
change of PL spectra of 2 with an increasing concentration of
DMSO (either in intensity or as a solvatochromic shift; Fig. S5 in
the ESI†). Our results suggest that the polarity effects of the TPP
do not contribute to the enhancement of the PL of AuNCs. Fig. 5 Examples of orbitals in the Au13-TPP model calculated at the
To analyze further the chemical structure and electronic BP86/6-31G**/MWB60 level. The red and green colors correspond to
behaviour of the AuNCs we used 1H NMR spectroscopy. The the positive and negative lobes.
1
H NMR spectrum of 1 (Fig. S6†) shows that the signals corres-
ponding to the TA are missing. However, after conjugation of
TPP with the carboxylic acid group on the surface of 1, the TA Finally, to demonstrate a broader applicability of our PL
signals reappear. The disappearance of TA signals in 1 may be enhancement method, we prepared another NIR PL AuNC 4
explained by the interaction of TA with a paramagnetic gold stabilized with mercaptosuccinic acid and PEG. Then we con-
core, leading to extreme signal broadening due to the para- jugated it with TPP salt via amidic coupling to obtain 5
magnetic relaxation enhancement. It is likely that the carboxy- (Scheme S2 in ESI†). We observed a similar enhancement of
late group also interacts with the nanoparticle surface, absorption and PL upon conjugation of TPP to the case of 2
because even the signals of hydrogen atoms close to the car- (Fig. S8†). This experimental result shows that the method
boxylate group are missing in the spectrum (Fig. S6†). The developed in this study can be extended to other thiolate-pro-
reappearance of the signals after TPP addition indicates that tected NIR emitting AuNCs for PL enhancement.
the structure of the interacting radical (gold core) is different. By using a simple synthetic procedure, we were able to
Signals of hydrogen atoms distant from the sulphur atoms are prepare high QY colloidally stable AuNCs emitting in the NIR
broad but in the same positions as for free TA and signals of region. In comparison with other reported methods3 of substi-
hydrogen atoms close to sulphur (and AuNC) are significantly tution our two step synthetic method in water (without the
broadened and shifted from their original positions. These need for an organic solvent) at room temperature uses very
spectral patterns indicate that TA interacts with a paramag- simple chemistry, which could be appealing for broader
netic centre (gold core), but the nature of the radical (gold research community interested in long term in vitro and in vivo
core) has changed upon TPP introduction leading to a sup- imaging of cells and tissues with a high signal-to-background
pression of paramagnetic relaxation and relative line narrow- ratio by avoiding auto fluorescence, reduced light scattering,
ing, most likely because of the MLCT effect. and high tissue penetration. Most importantly, our method
We have also simulated the possibility of electronic charge produces water soluble AuNCs, which does not require ligand
transfer from AuNC to TPP under electronic excitations by the exchange for biological applications in an aqueous environ-
time dependent density functional theory (TDDFT). From the ment. By using state-of-the-art characterization techniques such
emission spectra (Fig. 1c), we suppose that our AuNCs are as HRTEM, XPS, UV-vis absorption, temperature dependent fluo-
closely related to the Au25 cluster. The Au25 cluster is based on rescence lifetime, NMR spectroscopy, counteranion dependent
a centered icosahedral Au13 core, which is capped by an emission enhancement, in combination with DFT modelling,
exterior shell composed of twelve Au atoms.54 For simplicity, we were able to provide further insight into the mechanism of
we approximated our particles with the Au13 core bearing one PL enhancement. Our results indicate a gold-to-phosphonium
TPP molecule. The Au13 AuNC interacted strongly with TPP ion charge transfer, i.e., metal-to-ligand (MLCT) responsible for
and a substantial amount of electronic charge was transferred enhancement of PL. In contrast to commonly observed PL
from the Au13 nanocluster to the TPP. We found the frontier enhancement by ligand-to-metal charge transfer (LMCT) via the
orbitals (HOMO/LUMO) centered at the phosphorus atom and S–Au bond from the electron-rich groups (e.g., carboxylic, and
gold cluster, respectively (Fig. 5). This implies an extensive amino groups) present in the ligand (e.g., glutathione),38,55 our
charge transfer during the excitation, large transition dipole findings of MLCT mediated that PL enhancement offers an
moments and hence strong fluorescence intensities. alternative route to achieve high QY NIR PL AuNCs.

3796 | Nanoscale, 2018, 10, 3792–3798 This journal is © The Royal Society of Chemistry 2018
Nanoscale Paper

Conclusions Analytical Instruments, Kleve, Germany). NMR spectra were


recorded on a Bruker Avance III 500 spectrometer (499.88 MHz
We discovered unprecedentedly high enhancement of lumine- for 1H and 125.71 MHz for 13C) equipped with a 5 mm PFG
scence in AuNCs explicable by efficient charge transfer from cryoprobe. All 1H and 13C spectra were acquired for samples in
the gold nanocluster to triphenylphosphonium cations present CD3CN and referenced to the solvent signal (1.94 ppm in 1H).
in the ligand-capping layer. Our findings provide a design Signals of all hydrogen atoms of TA were assigned by using a
principle useful for the synthesis of high-QY NIR PL AuNCs for combination of 1D and 2D (H,H-COSY, H,C-HSQC and H,
bioimaging applications. We believe that our work will stimu- C-HMBC) techniques. HRTEM was performed with an FEI
late further experimental and theoretical research on photolu- Osiris instrument operated at 200 kV. The samples were pre-
minescent gold nanoclusters for broad applications, including pared by drop-casting on an ultrathin film C grid. Zeta poten-
light-harvesting systems that involve charge transfer. Moreover, tial experiments were carried out on a Zetasizer Nano ZSP
our results show that in addition to the size and composition (Malvern Instruments, UK) with a 633 nm laser at 25 °C. For
of the Au core and ligand attachment chemistry, the electronic zeta potential measurements, the universal dip cell in disposa-
properties of surface ligands play an extremely important role ble cuvettes was used. 10 μL of AuNC solution (200 μg mL−1)
in PL behavior of AuNCs. We envision that our simple syn- was mixed with 1 mL phosphate buffer solution (10 mM, pH
thetic protocol can be extended to fabricate highly photolumi- 7.4) and the solution was used for the measurement. For zeta
nescent NIR emitting AuNCs protected with custom-designed potential measurements, the average value of at least three
ligands with different functionalities. data points is reported.

Experimental section Conflicts of interest


Preparation of AuNC 1
There are no conflicts to declare.
In a typical reaction, 1.7 μL of 470 mg mL−1 HAuCl4·3H2O was
added to 3.9 mL deionized water containing TA (1.3 mg,
6.3 μmol) and 10 μL of 2 M NaOH solution. The mixture was
stirred at room temperature for 15 min, then 80 μL NaBH4
Acknowledgements
(1.9 mg mL−1 freshly prepared stock solution) was added, fol- The authors acknowledge support from the GACR project Nr.
lowed by further stirring of the reaction mixture overnight. The 18-12533S. J. V. acknowledges funding from the Ministry of
solution was purified by applying three cycles of centrifu- Education, Youth and Sports of the Czech Republic via the
gation/filtration using a membrane filtration device (Millipore) V4+Japan project No. 8F15001 (cofinanced by the International
with a molecular weight cut-off 3 kDa. Next, thiol-terminated Visegrad Fund). P. B. acknowledges GACR project No.
polyethylene glycol (MW 2000; 2.6 mg; 1.3 μmol) was added to 16-05935S and Ministry of Education, Youth and Sports of the
the solution, and the mixture was stirred overnight. The dis- Czech Republic project No. LTC17012.
persion was purified by applying three cycles of centrifugation/
filtration using a membrane filtration device (Millipore) with a
molecular weight cut-off of 3 kDa.
References
Preparation of AuNC 2
1 C. Dong, Z. Liu, J. Liu, C. Wu, F. Neumann, H. Wang,
AuNC 1 solution (4 mL, 0.2 mg mL−1) and 3-(aminopropyl)tri- M. Schäfer-Korting, B. Kleuser, J. Chang, W. Li, N. Ma and
phenylphosphonium bromide (or the respective TPP salt with R. Haag, Adv. Healthcare Mater., 2016, 5, 2214–2226.
BF4−, Cl−, or I−) (2 mg, ∼5 μmol) were mixed together, and the 2 L.-Y. Chen, C.-W. Wang, Z. Yuan and H.-T. Chang, Anal.
pH was adjusted to the range of 4.5–6.0 with 1 M HCl. The Chem., 2015, 87, 216–229.
reaction was started by adding excess N-(3-dimethyl- 3 S. Wang, X. Meng, A. Das, T. Li, Y. Song, T. Cao, X. Zhu,
aminopropyl)-N′-ethylcarbodiimide hydrochloride (EDC·HCl) M. Zhu and R. Jin, Angew. Chem., Int. Ed., 2014, 53, 2376–
(10 mg, 52 μmol) and was stirred overnight. The dispersion 2380.
was purified by applying three cycles of centrifugation/fil- 4 N. Goswami, Z. Luo, X. Yuan, D. T. Leong and J. Xie, Mater.
tration using a membrane filtration device (Millipore) with a Horiz., 2017, 4, 817–831.
molecular weight cut-off of 3 kDa to obtain AuNC 2. 5 X.-D. Zhang, Z. Luo, J. Chen, X. Shen, S. Song, Y. Sun,
S. Fan, F. Fan, D. T. Leong and J. Xie, Adv. Mater., 2014, 26,
Characterization 4565–4568.
UV-vis absorption and photoluminescence spectra of AuNCs 6 K. Zheng, M. I. Setyawati, D. T. Leong and J. Xie, ACS Nano,
were recorded using a TECAN M1000 microplate reader. 2017, 11, 6904–6910.
Phosphorus and gold contents were estimated using an optical 7 Y. Ju-Nam, Y.-S. Chen, J. J. Ojeda, D. W. Allen, N. A. Cross,
emission spectrometer (OES) with radial observation of induc- P. H. E. Gardiner and N. Bricklebank, RSC Adv., 2012, 2,
tively coupled plasma (ICP) (SPECTRO Arcos – SPECTRO 10345–10351.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 3792–3798 | 3797
Paper Nanoscale

8 J. Zheng, C. Zhang and R. M. Dickson, Phys. Rev. Lett., 31 K. G. Stamplecoskie and P. V. Kamat, J. Am. Chem. Soc.,
2004, 93, 077402. 2014, 136, 11093–11099.
9 J. Zheng, J. T. Petty and R. M. Dickson, J. Am. Chem. Soc., 32 Z. Wu, C. Gayathri, R. R. Gil and R. Jin, J. Am. Chem. Soc.,
2003, 125, 7780–7781. 2009, 131, 6535–6542.
10 R. Zhou, M. Shi, X. Chen, M. Wang and H. Chen, Chem. – 33 Y. Negishi, Y. Takasugi, S. Sato, H. Yao, K. Kimura and
Eur. J., 2009, 15, 4944–4951. T. Tsukuda, J. Am. Chem. Soc., 2004, 126, 6518–6519.
11 W.-Y. Chen, G.-Y. Lan and H.-T. Chang, Anal. Chem., 2011, 34 X. Wen, P. Yu, Y.-R. Toh, A.-C. Hsu, Y.-C. Lee and J. Tang,
83, 9450–9455. J. Phys. Chem. C, 2012, 116, 19032–19038.
12 H. Kawasaki, K. Hamaguchi, I. Osaka and R. Arakawa, Adv. 35 X. Wen, P. Yu, Y.-R. Toh and J. Tang, J. Phys. Chem. C, 2012,
Funct. Mater., 2011, 21, 3508–3515. 116, 11830–11836.
13 J. Xie, Y. Zheng and J. Y. Ying, J. Am. Chem. Soc., 2009, 131, 36 P.-C. Chen, C.-K. Chiang and H.-T. Chang, J. Nanopart. Res.,
888–889. 2012, 15, 1336.
14 T.-H. Chen and W.-L. Tseng, Small, 2012, 8, 1912–1919. 37 J. Michl and V. Bonacic-Koutecky, Electronic Aspects of
15 P. L. Xavier, K. Chaudhari, A. Baksi and T. Pradeep, Nano Organic Photochemistry, John Wiley & Sons, Inc., New York,
Rev., 2012, 3, 14767. 1990.
16 M. A. H. Muhammed, S. Ramesh, S. S. Sinha, S. K. Pal and 38 Z. Wu and R. Jin, Nano Lett., 2010, 10, 2568–2573.
T. Pradeep, Nano Res., 2010, 1, 333–340. 39 S. E. Crawford, C. M. Andolina, A. M. Smith, L. E. Marbella,
17 J. Zheng, P. R. Nicovich and R. M. Dickson, Annu. Rev. Phys. K. A. Johnston, P. J. Straney, M. J. Hartmann and
Chem., 2007, 58, 409–431. J. E. Millstone, J. Am. Chem. Soc., 2015, 137, 14423–14429.
18 Z. Luo, X. Yuan, Y. Yu, Q. Zhang, D. T. Leong, J. Y. Lee and 40 J. Jiang, C. V. Conroy, M. M. Kvetny, G. J. Lake,
J. Xie, J. Am. Chem. Soc., 2012, 134, 16662–16670. J. W. Padelford, T. Ahuja and G. Wang, J. Phys. Chem. C,
19 N. Goswami, F. Lin, Y. Liu, D. T. Leong and J. Xie, Chem. 2014, 118, 20680–20687.
Mater., 2016, 28, 4009–4016. 41 G. Wang, R. Guo, G. Kalyuzhny, J.-P. Choi and
20 N. Goswami, Q. Yao, Z. Luo, J. Li, T. Chen and J. Xie, R. W. Murray, J. Phys. Chem. B, 2006, 110, 20282–20289.
J. Phys. Chem. Lett., 2016, 7, 962–975. 42 E. Oh, K. Susumu, R. Goswami and H. Mattoussi,
21 A. Yahia-Ammar, D. Sierra, F. Mérola, N. Hildebrandt and Langmuir, 2010, 26, 7604–7613.
X. Le Guével, ACS Nano, 2016, 10, 2591–2599. 43 L. Sementa, G. Barcaro, A. Dass, M. Stener and
22 K. Pyo, V. D. Thanthirige, K. Kwak, P. Pandurangan, A. Fortunelli, Chem. Commun., 2015, 51, 7935–7938.
G. Ramakrishna and D. Lee, J. Am. Chem. Soc., 2015, 137, 44 C. Zhou, C. Sun, M. Yu, Y. Qin, J. Wang, M. Kim and
8244–8250. J. Zheng, J. Phys. Chem. C, 2010, 114, 7727–7732.
23 H.-H. Deng, X.-Q. Shi, F.-F. Wang, H.-P. Peng, A.-L. Liu, 45 A. Vogler and H. Kunkely, Coord. Chem. Rev., 2001,
X.-H. Xia and W. Chen, Chem. Mater., 2017, 29, 1362– 219–221, 489–507.
1369. 46 V. W.-W. Yam and E. C.-C. Cheng, Chem. Soc. Rev., 2008,
24 G. Wang, T. Huang, R. W. Murray, L. Menard and 37, 1806–1813.
R. G. Nuzzo, J. Am. Chem. Soc., 2005, 127, 812–813. 47 T. Huang and R. W. Murray, J. Phys. Chem. B, 2001, 105,
25 L. Shang, N. Azadfar, F. Stockmar, W. Send, V. Trouillet, 12498–12502.
M. Bruns, D. Gerthsen and G. U. Nienhaus, Small, 2011, 7, 48 E. Sakuda, Y. Ando, A. Ito and N. Kitamura, Inorg. Chem.,
2614–2620. 2011, 50, 1603–1613.
26 C.-A. J. Lin, T.-Y. Yang, C.-H. Lee, S. H. Huang, 49 E. A. Medlycott and G. S. Hanan, Chem. Soc. Rev., 2005, 34,
R. A. Sperling, M. Zanella, J. K. Li, J.-L. Shen, H.-H. Wang, 133–142.
H.-I. Yeh, W. J. Parak and W. H. Chang, ACS Nano, 2009, 3, 50 Y. Negishi, K. Nobusada and T. Tsukuda, J. Am. Chem. Soc.,
395–401. 2005, 127, 5261–5270.
27 D. Mishra, F. Aldeek, E. Lochner, G. Palui, B. Zeng, 51 C. Imrie, T. A. Modro and C. C. P. Wagener, J. Chem. Soc.,
S. Mackowski and H. Mattoussi, Langmuir, 2016, 32, 6445– Perkin Trans. 2, 1994, 1379–1382.
6458. 52 K.-I. Oh, K. Rajesh, J. F. Stanton and C. R. Baiz, Angew.
28 X. Huang, Y. Luo, Z. Li, B. Li, H. Zhang, L. Li, I. Majeed, Chem., Int. Ed., 2017, 56, 11375–11379.
P. Zou and B. Tan, J. Phys. Chem. C, 2011, 115, 16753– 53 J. R. Lakowicz, Principles of Fluorescence Spectroscopy,
16763. Springer, Boston, MA, 3rd edn, 2006.
29 H.-C. Chang, Y.-F. Chang, N.-C. Fan and J. A. Ho, ACS Appl. 54 M. Zhu, C. M. Aikens, F. J. Hollander, G. C. Schatz and
Mater. Interfaces, 2014, 6, 18824–18831. R. Jin, J. Am. Chem. Soc., 2008, 130, 5883–5885.
30 F. Aldeek, M. A. H. Muhammed, G. Palui, N. Zhan and 55 K. G. Stamplecoskie, Y.-S. Chen and P. V. Kamat, J. Phys.
H. Mattoussi, ACS Nano, 2013, 7, 2509–2521. Chem. C, 2014, 118, 1370–1376.

3798 | Nanoscale, 2018, 10, 3792–3798 This journal is © The Royal Society of Chemistry 2018
Electronic Supplementary Material (ESI) for Nanoscale.
This journal is © The Royal Society of Chemistry 2018

Supplementary Information
Gold Nanoclusters with Bright Near-Infrared Photoluminescence

Goutam Pramanik, Jana Humpolickova, Jan Valenta, Paromita Kundu, Sara Bals, Petr Bour,
Martin Dracinsky, and Petr Cigler*

Materials
All reagents were purchased from commercial sources and used as received. Aqueous
solutions in all experiments were prepared using Milli-Q water.
3-(aminopropyl)triphenylphosphonium (TPP) bromide salt was prepared as described in the
literature.1 The corresponding chloride, iodide, and tetrafluoroborate salts of TPP were
obtained by exchange of the Br– salt on an anion exchange column loaded with Amberlite®
IRA-400 in the corresponding anion cycle, following published procedures.2

Determination of Quantum Yield


Quantum yield (QY) were measured by a relative comparison method using the following
equation
2
𝐹𝑠𝑎𝑚𝑝𝑙𝑒 𝐴𝑟𝑒𝑓 𝜂𝑠𝑎𝑚𝑝𝑙𝑒
𝑄𝑌𝑠𝑎𝑚𝑝𝑙𝑒 = 𝑄𝑌𝑟𝑒𝑓 ∗ ∗ ∗
𝐹𝑟𝑒𝑓 𝐴𝑠𝑎𝑚𝑝𝑙𝑒 2
𝜂𝑟𝑒𝑓 (Equation
S1)

where F, A, and η are the measured fluorescence intensity (integrated areas of


photoluminescence spectra of the AuNCs and Rhodamine 6G in the region between 500 –
850 nm), the absorbance at the excitation wavelength (480 nm), and the refractive index of the
solvent, respectively. The subscripts “sample” and “ref” stand for AuNC samples and
reference (Rhodamine 6G) respectively. The emission spectra of the 1 and 2 were measured in
water and Rhodamine 6G was measured in ethanol upon excitation at 480 nm at the same
conditions. Following this, the areas under the curves were determined using
Microsoft Excel. To avoid inner filter effects, the absorption of the AuNCs and reference were
kept below 0.05 at the excitation wavelength (480 nm). The refractive index for water is 1.33
(ηsample), and for ethanol is 1.36 (ηref). All the measurements were performed at 25 °C. QYref is
the quantum yield of the Rhodamine 6G (0.95) in ethanol at 25 oC.

Table S1. Calculation of quantum yields of the AuNCs.

Samples Asample Aref Fsample Fref ηsample ηref QYref QYsample


(at 480 nm) (at 480 nm)
1 0.0395 0.0322 1836 12915 1.33 1.36 0.95 0.1053
2 0.0396 0.0322 4505 12915 1.33 1.36 0.95 0.2576

1
Fig. S1 Normalized emission spectra of Fig. S2 Relative emission spectra of
aqueous solutions of 1, 2 (X¯ = BF4¯) and aqueous solutions of 2 at different
Rhodamine 6G collected upon excitation at Au:P ratio upon excitation at 365
480 nm. nm.

X-ray photoelectron spectroscopy (XPS) measurements


XPS measurements were performed using Omicron Nanotechnology equipment. The primary
X-ray beam was monochrome radiation from an Al lamp with an energy of 1486.7 eV. The
constant analyzer energy (CAE) mode was used, and intensity calibration was based on
copper and copper spectra-derived calibration constants. The ion gun used argon ions with an
energy of ~5 keV; ion etching was carried out in the preparation chamber. A circular area with
a diameter of approximately 0.8 mm was used for analysis. Measured spectra were evaluated
with CasaXPS software. The binding energy of a known element (usually C 1s with binding
energy 284.8 eV) was used for calibration of the binding energy axis.

Photoluminescence lifetime measurements


The photoluminescence lifetime measurements were carried out using a time correlation
single photon counting (TCSPC) system integrated into Fluoromax-4 (Horiba). We used a
pulsed excitation signal at 370 nm with a repetition rate of 1 MHz, provided by a NanoLED-
370 light source (100 ps, fwhm), and the emission was collected using a TBX photomultiplier
detector. We note the temperature-dependent study concentrates only on the slow component
of luminescence decay. We fitted the photoluminescence decay by three exponentials using
equation (1), but it can be fitted also by a distribution of lifetimes, for example by a stretched
exponential (Kohlrausch) function

𝐼(𝑡) = 𝐴𝑒
( ),

𝑡 𝛽
𝜏
(Equation S2)
where parameter β is around 0.6. This value indicates that we have a process with certain
distribution of lifetimes, e. g. due to variation of ligands and their conformations. For
evaluation of the temperature evolution of emission rate we use just the most representative
lifetime, i. e. an average lifetime (obtained by the integration method, i. e. independent of the
fitting model). The slow microsecond components of decay according to equation (1) look
like minor compared to the nanosecond one, but this is due to the short pulse excitation which
is not exciting slow processes efficiently.3 We have checked that under long microsecond
pulses the slow decay components largely dominates (nanosecond component is negligible or
null; data not shown).

2
Fig. S3 31P NMR (161.98 MHz, D2O) of (a) 2 and (b) 3-(aminopropyl) triphenylphosphonium
bromide. The peaks were assigned relative to H3PO4 (0.00 ppm).

Scheme S1. Schematic representation of the structure and preparation of AuNC 3.

Scheme S2. Schematic representation of the structure and preparation of mercaptosuccinic


acid stabilized AuNCs 4 and 5.

Preparation of AuNC 3

AuNC 1 solution (4 mL, 0.2 mg/mL) and excess of 2-aminoanthracene (6 mg, 31 μmol)
solution in 1 mL DMSO were mixed together. Then the pH of the solution was adjusted to the
range of 4.5–6.0 with 1 M HCl. The reaction was started by adding excess N-(3-
3
dimethylaminopropyl)-N′-ethylcarbodiimide hydrochloride (EDC.HCl) (10 mg, 52 μmol) and
stirred overnight. Then the reaction mixture was lyophilized. Then 4 mL water was added to
the reaction mixture and sonicated for 2 minutes. Then the insoluble part was removed by
filtration. The filtrate was purified by applying three cycles of centrifugation/filtration using a
membrane filtration device (Millipore) with a molecular weight cut‐off of 3 kDa to obtain
AuNC 3 and the absorption and emission spectra were acquired immediately.

Preparation of AuNCs 4 and 5


AuNC 4 was prepared using the same procedure as 1, only mercaptosuccinic acid (2 mg,
13.3 μmol) was used instead of TA. AuNC 5 was prepared from 4 by the same procedure as 2
from 1.

Fig. S4 Relative photoluminescence Fig. S5 Relative photoluminescence


spectra of aqueous solutions of AuNC 1, 2 spectra of aqueous and DMSO mixed
and 3 upon excitation at 365 nm. solution of AuNC 2 upon excitation at
365 nm.

4
Fig. S6 1H NMR (500 MHz, CD3CN) spectra of TA, thiol-terminated PEG, AuNC 1 and 2. In
the top panel the whole spectra are shown, the bottom panel shows detailed signal assignment.
For estimation of ligand (PEG:TA) ratios on 2, the following peaks were integrated: terminal
CH3-O group on PEG (3H, 3.20 ppm) and sum of selected methylene groups from TA (6H,
1.24 – 1.80 ppm). For estimation of TPP:TA ratio, the aromatic region of TPP (15H, 7.43 –
7.95 ppm) was used and the same methylene groups for TA as above.

5
Computations
To understand the AuNCs fluorescence properties, we constructed a simpler molecular model,
where centered icosahedral Au13 core is connected to one TPP molecule (Fig. S7). The initial
Au13 geometry was taken from the X-ray structure of a larger 25-atom cluster,4 initial TPP and
Au13TPP geometries were constructed with a fully extended TPP backbone. Au13 was
negatively and TPP, Au13 TPP were positively charged. Then geometries of the systems were
optimized by energy minimization at the BP86/56-31G** level; the MWB606 pseudopotential
and basis set was used to Au atoms and the Grimme’s GD2 dispersion correction7 applied.
The Gaussian program8 was used for all the quantum-chemical computations. For the
optimized structures energies of the excited states were calculated using the time dependent
density functional theory.9 Qualitatively similar results were obtained using the B3LYP10
functional and CPCM11 model for a solvent environment.

Fig. S7 Optimized geometries of the Au13-, TPP+ and Au13TPP+. The following color code is
used: Au – yellow (larger spheres), S – yellow, C – grey, H – white, O – red, N – blue, P –
orange.

Fig. S8 (a) Normalized absorption spectra of aqueous solutions of AuNC 4 and 5. (b)
Normalized emission spectra of aqueous solutions of 4 and 5 collected upon excitation at
365 nm.

References

1 C.-J. Zhang, J. Wang, J. Zhang, Y. M. Lee, G. Feng, T. K. Lim, H.-M. Shen, Q. Lin and B.
Liu, Angew. Chem., 2016, 128, 13974–13978.
2 V. J. A. Jameson, H. M. Cochemé, A. Logan, L. R. Hanton, R. A. J. Smith and M. P.
Murphy, Tetrahedron, 2015, 71, 8444–8453.
6
3 M. Greben and J. Valenta, Rev. Sci. Instrum., 2016, 87, 126101.
4 M. Zhu, C. M. Aikens, F. J. Hollander, G. C. Schatz and R. Jin, J. Am. Chem. Soc., 2008,
130, 5883–5885.
5 J. P. Perdew, Phys. Rev. B, 1986, 33, 8822–8824.
6 D. Figgen, G. Rauhut, M. Dolg and H. Stoll, Chem. Phys., 2005, 311, 227–244.
7 S. Grimme, J. Comput. Chem., 2006, 27, 1787–1799.
8 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman,
G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich,
J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F.
Izmaylov, J. L. Sonnenberg, Williams, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng,
A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng,
W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima,
Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery Jr., J. E. Peralta,
F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A.
Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S.
Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W.
Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian
16, Wallingford, CT, 2016.
9 F. Furche and R. Ahlrichs, J. Chem. Phys., 2002, 117, 7433–7447.
10A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652.
11A. Klamt and G. Schüürmann, J. Chem. Soc. Perkin Trans. 2, 1993, 0, 799–805.

S-ar putea să vă placă și