Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Article

pubs.acs.org/ac

“Turn-On” Fluorescent Sensor for Hg2+ Based on Single-Stranded


DNA Functionalized Mn:CdS/ZnS Quantum Dots and Gold
Nanoparticles by Time-Gated Mode
Dawei Huang,† Chenggang Niu,*,† Xiaoyu Wang,†,‡ Xiaoxiao Lv,† and Guangming Zeng*,†

College of Environmental Science and Engineering, Key Laboratory of Environmental Biology and Pollution Control, Ministry of
Education, Hunan University, Changsha 410082, China

College of Chemistry and Chemical Engineering, Xinxiang University, Xinxiang 453003, China

ABSTRACT: An ultrasensitive “turn-on” fluorescent sensor


was presented for determination of Hg2+. This method is
mainly based on Hg2+-induced conformational change of a
thymine-rich single-stranded DNA. The water-soluble long-
lifetime fluorescence quantum dot (Mn:CdS/ZnS) acted as
the fluorophore, which was labeled on a 33-mer thymine-rich
single-stranded DNA (strand A). The gold nanoparticles
(GNPs) functionalized 10-mer single-stranded DNA (strand
B) is selected as the quencher to quench the fluorescence of
Mn:CdS/ZnS. Without Hg2+ in the sample solution, strands A and B could form hybrid structures, resulting in the fluorescence
of Mn:CdS/ZnS being decreased sharply. When Hg2+ is present in the sample solution, Hg2+-mediated base pairs induced the
folding of strand A into a hairpin structure, leading to the release of GNPs-tagged strand B from the hybrid structures. The
fluorescence signal is then increased obviously compared with that without Hg2+. The sensor exhibits two linear response ranges
between fluorescence intensity and Hg2+ concentration. Meanwhile, a detection limit of 0.18 nM is estimated based on 3α/slope.
Selectivity experiments reveal that the fluorescent sensor is specific for Hg2+ even with interference by high concentrations of
other metal ions. This sensor is successfully applied to determination of Hg2+ in tap water and lake water samples. This sensor
offers additional advantage to efficiently reduce background noise using long-lifetime fluorescence quantum dots by a time-gated
mode. With excellent sensitivity and selectivity, this sensor is potentially suitable for monitoring of Hg2+ in environmental
applications.

H eavy metal ion pollution has become an important


worldwide issue for years due to the severe risks in
human health and the environment. As one of the most toxic
It has been previously reported that Hg2+ can selectively link
T−T pairs to form stable T−Hg2+−T base pairs.9,10 The Hg2+-
mediated T−Hg2+−T pair is even more stable than the
heavy metals, mercury with the feature of strong toxicity and Watson−Crick A−T pair.11 Hg2+ can be incorporated into
bioaccumulation can cause serious human health problems DNA duplex without altering the double-helical structure
even at very low concentrations.1,2 The exposure to mercury because the van der Waals radius of mercury (≈1.44 Å) is
can cause a number of toxicological effects such as brain smaller than the base pair spacing of DNA duplex (≈3.4 Å).9 A
damage, kidney failure, and various cognitive and motion large number of fluorescent, colorimetric, and electrochemical
sensors were designed for Hg2+ based on T−Hg2+−T
disorders.3,4 The solvated Hg2+, one of the most stable
coordination chemistry.12−14 As far as fluorescent sensors are
inorganic forms of mercury, is well-known to be highly toxic
concerned, most of them can be suitable for Hg2+ detection in
due to its good water solubility.5 Because of these environ- drinking water because detection limits are lower than the toxic
mental and health problems of Hg2+, it is obviously of great level reported by the U.S. Environmental Protection Agency.
necessity to obtain new and efficient Hg2+ detection methods However, many fluorescent sensors work in a “turn-off”
that are cost-effective, rapid, sensitive, and selective. Traditional mode.15−17 As we know well, the sensors working in “turn-
methods for Hg2+ quantification include atomic absorption/ off” mode may produce false positive results because of a
emission spectroscopy, selective cold vapor atomic fluorescence number of other quenchers.18 As a result, a “turn-on” sensor is
spectrometry, inductively coupled plasma mass spectrometry preferred.12,14,19−21
(ICPMS), and so forth. These methods are very sensitive and Recently, researchers have devoted a lot of efforts to the
selective but require complicated sample preparation and fluorescent approach that involves the use of nanometer
sophisticated instruments which limit their application in
routine Hg2+ monitoring.6−8 Thus, it is still of great challenge Received: October 23, 2012
to develop a method which is sensitive, selective, and Accepted: December 20, 2012
environmentally friendly for aqueous Hg2+ detection.

© XXXX American Chemical Society A dx.doi.org/10.1021/ac303084d | Anal. Chem. XXXX, XXX, XXX−XXX
Analytical Chemistry Article

materials, especially semiconductor quantum dots (QDs) or/ Japan). Atomic fluorescence measurements were performed on
and gold nanoparticles (GNPs), for the detection of Hg2+ based an atomic fluorescence spectrometer (AFS-9700) (Beijing,
on this T−Hg2+−T coordination chemistry.22,23 QDs have China).
many unique photophysical properties such as high fluores- PBS buffer (100 mM) was prepared by mixing an appropriate
cence quantum yields, narrow emission bands, high Stokes content of 200 mM Na2HPO4 and 200 mM NaH2PO4. The
shifts, and stability against photobleaching. Because of these composition of hybridization buffer, incubation buffer, and
excellent properties, QDs were used as fluorescence labels to washing buffer was 10 mM PBS buffer (pH = 7.4), 100 mM
develop fluorescent sensors.24−26 GNPs have been of great NaNO3. In addition, 2 M NaCl was also prepared.
interest because of their high extinction coefficient and a broad Synthesis of Mn-Doped CdS/ZnS Core/Shell QDs. Mn-
absorption spectrum in a visible light. GNPs are unique doped CdS/ZnS core/shell QDs (Mn:CdS/ZnS QDs) were
quenchers for organic dyes or QDs through both energy- prepared according to a three-step synthesis method.33,34 The
transfer and electron-transfer processes.27,28 A lot of sensors resulting Mn:CdS/ZnS QDs were dispersed in hexane. The
have been fabricated based on GNPs as quenchers for DNA, lack of water solubility of the prepared QDs hindered their
small molecules, or protein detection.29−32 Up to now, reaction with the water-soluble alkylthiol-capped oligonucleo-
fluorescent sensors based on QDs and GNPs are still a good tides. Therefore, the authors used 3-mercaptopropionic acid to
choice for the detection of different analytes. prepare the water-soluble QDs according to the literature.38
No matter whether it is “turn-on” or “turn-off” mode, most The resulting water-soluble QDs also have long-lifetime
fluorescent sensors for Hg2+ are based on organic dyes or QDs fluorescence (∼4.8 ms) and exhibit high stability in aqueous
that usually have short-lifetime fluorescence. The background solutions. The water-soluble QDs should be an excellent
signals might interfere with the fluorescence of organic dyes or fluorescence label and could play an important role in a number
QDs, affecting the sensitivity of the fluorescent sensors. of QDs-based biochemical and biomedical applications.
Therefore, it should be desirable to develop a novel method, Synthesis of GNPs. All glassware and mechanical stirrers
which uses a long-lifetime fluorophore to decrease the used for the synthesis were thoroughly cleaned in aqua regia (3
background noises on the basis of time-gated fluorescence parts HCl, 1 part HNO3), rinsed with ultrapure water, and then
assay. The authors found that Mn-doped QDs are good choices oven-dried prior to use. The colloidal solution of GNPs was
which have high quantum yield and long-lifetime fluores- synthesized by means of citrate reduction of
cence.33−37 AuCl3·HCl·4H2O.39 An amount of 5 mL of 38.8 mM sodium
In this work, the authors designed a simple “turn-on” citrate was rapidly added to a boiled 50 mL of 1 mM HAuCl4
fluorescent sensor for Hg2+ in aqueous solution that utilized the solution with vigorous stirring in a 250 mL round-bottom flask
T−Hg2+−T coordination chemistry and its induced displace- equipped with a condenser. The color changed from light
ment of the quencher-labeled oligonucleotides. Water-soluble yellow to wine red. Boiling was continued for 10 min; the
Mn-doped CdS/ZnS core/shell QDs [the QDs have long- heating mantle was then removed, and stirring was continued
lifetime fluorescence (∼4.8 ms) and excellent stability in for an additional 15 min. After the solution was cooled to room
aqueous solution] and GNPs are selected as the fluorophore temperature, the prepared GNPs solution was stored in the 4
and quencher, respectively. This fluorescent sensor has a °C refrigerator before use. The size of the nanoparticles was
detection limit below the U.S. Environmental Protection typically ∼13 nm in average diameter. The concentration of the
Agency limit of acceptable Hg2+ concentration in drinking GNPs was ∼17 nM, which was determined according to the
water. The sensor also exhibits superior selectivity toward Hg2+ Beer’s law by using UV−vis spectroscopy, based on the
even in the presence of other competitive metal ions. extinction coefficient of 2.7 × 108 M−1 cm−1 at λ = 520 nm for
Furthermore, the sensor was employed to detect Hg2+ spiked 13 nm particles.40
in tap water and lake water samples to demonstrate its potential Preparation of DNA-Functionalized GNPs. According to
for practical applications. literature with slight modifications,41 8.0 mL of 17 nM GNPs

■ EXPERIMENTAL SECTION
Chemicals and Apparatus. All oligonucleotides used in
were incubated with 20 μL of 0.1 mM oligonucleotides
overnight. After standing for 16 h at 50 °C, the mixed solution
was changed into 0.1 M NaCl, 10 mM phosphate buffer (pH =
the present study were synthesized and HPLC-purified by 7.4) by addition of the necessary salts and was kept at 50 °C for
Shanghai Sangon Biological Engineering Technology & 40 h. To remove unreacted oligonucleotides, the oligonucleo-
Services Co., Ltd. (Shanghai, China) and dissolved in ultrapure tide-conjugated GNPs were purified three times by centrifuga-
water (18.3 MΩ·cm) produced by a Millpore water purification tion at 13 200 rpm for 30 min. The final product was
system. The sequences were shown as follows: 5′SH−C6− redispersed into 1.2 mL of PBS buffer (10 mM, pH = 7.4) to
TGAAA CTGTA-3′; 5′SH−C6−TACAG TTTCA CCTTT make a stock solution and stored at 4 °C for future usage. The
TCCCC CGTTT TGGTG TTT-3′. AuCl3·HCl·4H2O was number of oligonucleotides probes immobilized on the GNPs
purchased from Shanghai Institute of Fine Chemical Materials was estimated by measuring the absorbance difference at 260
(Shanghai, China). 3-Mercaptopropionic acid (MPA, 99+%) nm before and after modification with oligonucleotides. The
was purchased from Sigma-Aldrich. The chemicals were used as average oligonucleotide loadings were about 10 oligonucleo-
received without further purification. All other chemicals used tides per GNP, and the final concentration of GNPs was 95
were of analytical grade and were used without further nM.
purification. Ultrapure water was used throughout the experi- Preparation of DNA-Functionalized Mn:CdS/ZnS QDs.
ments. DNA-functionalized QDs were prepared according to a
The time-gated fluorescence intensities were measured and previously published protocol with minor modifications.38
recorded with a Perkin-Elmer LS-55 spectrofluorimeter The QDs solution and oligonucleotides solution were mixed
(United Kingdom). UV−vis absorption spectra were recorded together at a ratio of 11 oligonucleotides per QD (100 μL of
by using a Shimadzu UV spectrophotometer (UV-2550, Kyoto, 1.8 μM QDs mixed with 20 μL of 0.1 mM oligonucleotides).
B dx.doi.org/10.1021/ac303084d | Anal. Chem. XXXX, XXX, XXX−XXX
Analytical Chemistry Article

Scheme 1. Schematic Description of the “Turn-On” Fluorescent Sensor for Hg2+ Based on the Hg2+-Mediated Formation in
DNA Duplexesa

a
The 33-mer single-stranded DNA (strand A) with a Mn:CdS/ZnS quantum dot attached at the 5′ end was hybridized with a 10-mer single-stranded
DNA (strand B) with a gold nanoparticle attached at the 5′ end, which resulted in energy transfer from the Mn:CdS/ZnS quantum dot to the gold
nanoparticle, leading to a decrease in the time-gated fluorescence intensity of the Mn:CdS/ZnS quantum dot. In the presence of Hg2+, the folding of
strand A releases strand B and increases the fluorescence of the Mn:CdS/ZnS quantum dot. The drawing of QDs and GNPs modified single-
stranded DNA is only a graphic presentation.

After standing for 12 h, the mixed solution was brought to 0.15 Mg2+, Zn2+, Al3+, Fe3+, Pb2+, Ag+, and Au3+. The lake water
M NaCl and the particles were aged for an additional 12 h. The samples were taken from Taozi lake in Hunan University. The
NaCl concentration was then raised to 0.3 M, and the mixture time-gated fluorescence signal was measured and recorded by a
was allowed to stand for a further 40 h before centrifugalization Perkin-Elmer LS-55 spectrofluorimeter. The parameters of the
using centrifugal filter devices (Amicon Ultra-0.5). Finally, the spectrofluorimeter are set as λex = 400 nm; λem = 609 nm; delay
QDs were redispersed into 4.0 mL of PBS buffer (10 mM, pH time, 0.1 ms; gate time, 1.0 ms; excitation slit, 15 nm; emission
= 7.4) by vortex and stored at 4 °C for future usage. The slit, 20 nm.


number of oligonucleotides probes immobilized on the QDs
was also estimated by measuring the absorbance difference at RESULTS AND DISCUSSION
260 nm before and after modification with oligonucleotides.
Experimental Principle of the Proposed Sensor. The
The average oligonucleotide loadings were about six “turn-on” fluorescent sensor for Hg2+ is outlined in Scheme 1.
oligonucleotides per QD, and the final concentration of QDs The sensor system comprises two single-stranded DNAs
was 45 nM. (strand A and strand B) with an alkanethiol moiety at their
Procedures for Hg2+ Detection. To detect Hg2+ or other 5′-terminus. Strand A is a 33-mer thymine-rich oligonucleotide,
metal ions in buffer or real environmental water samples, 30 μL and strand B is a 10-mer oligonucleotide complementary with
of 95 nM DNA/GNPs, 70 μL of 45 nM DNA/QDs, and 140 strand A. Strand A contains two major parts: the first part (in
μL of 0.01 M PBS buffer were mixed uniformly by vortex and red) is a five-base segment close to the 5′-terminal that could
hybridized for 35 min first. Then, various concentrations of hybridize with strand B close to the 3′-terminal; five self-
Hg2+ (20 μL) were added into the mixture solution and complementary base pairs separated by seven thymine−
incubated for 16 min to form T−Hg2+−T coordination thymine mismatches are introduced to the second part (in
chemistry. Finally, the time-gated fluorescence spectra of blue). There are five-base segments in the second part which
different concentrations of Hg2+ were monitored after the could hybridize with five-base segments of strand B close to the
completion of the reaction. For the sensitivity experiment, the 5′-terminal. Strand A was functionalized with ∼5 nm sized
concentrations of Hg2+ were 0, 0.2, 0.4, 0.6, 0.8, 1.0, 2.0, 4.0, Mn:CdS/ZnS QDs, and strand B was prepared by function-
6.0, 8.0, 10.0, 20.0, 50.0, 100.0, 200.0, 500.0, and 1000.0 nM, alization of ∼13 nm diameter GNPs, according to the 5′-
respectively. For the optimizing experiment, 1.0 μM was terminal −SH reaction. The QDs with ∼5 nm in diameter
selected as the Hg2+ concentration to determine the optimum showed the fluorescence emission at 609 nm under light
hybridization time and incubation time. Various metal ions of excitation at 400 nm. In our experiment, the average distance
10 μM were used in the selectivity experiments. The metal ions between QD and GNP in hybridized structures was estimated
are as follows: Mn2+, Ba2+, Ni2+, Cu2+, Ca2+, Cr2+, Co2+, Cd2+, to be 5.25 nm. In the absence of Hg2+, strands A and B could
C dx.doi.org/10.1021/ac303084d | Anal. Chem. XXXX, XXX, XXX−XXX
Analytical Chemistry Article

hybridize each other because of the 10 complementary base operational temperature, and media pH played crucial roles
pairs. Under these conditions, the QDs and GNPs are close to in the detection sensitivity. The hybridization time between
each other, resulting in fluorescence quenching due to strands A and B was investigated as it may influence the
fluorescence resonance energy transfer. On the contrary, hybridization efficiency. Although longer hybridization time
when Hg2+ was present in the sensor solution, mercury- may yield more stable fluorescence signal, it is unnecessary if
mediated base pairs (T−Hg2+−T) induce the folding of strand the reaction attained the equilibrium. The fluorescence signal
A into a hairpin structure. Consequently, there are only five was recorded along with the hybridization time increasing
base pairs remaining between strands A and B, which is not (Figure 2). It is found that the fluorescence signal decreased
stable enough under the conditions we used in the proposed
method. As a result, strand B will be released from the hybrid
structure, and the time-gated fluorescence of QDs will be
observed upon light excitation. The fluorescence spectra of the
sensor before and after the addition of 1.0 μM Hg2+ is shown in
Figure 1A. Furthermore, the authors used a control experiment
with the results of no fluorescence intensity changed, in which
Hg2+ was added to the solution only containing strand A. The
results indicated that Hg2+ makes negligible contribution to
quench the fluorescence of QDs (Figure 1B).
Optimization of the Experimental Conditions. In the
present strategy, the hybridization and incubation times,

Figure 2. Optimization experiments of hybridization and incubation


time; 1.0 μM was selected as the Hg2+ concentration to determine the
optimum hybridization and incubation time.

with increasing hybridization time and became stable over 32


min, then continued at an almost constant value. To ensure the
completeness of hybridization, the authors chose 35 min as the
optimum hybridization time.
The incubation time after the introduction of Hg2+ into the
solution to prompt the T−Hg2+−T formation was also
investigated, and the results are shown in Figure 2. The
fluorescence intensity increased when Hg2+ was introduced and
then tended to stabilize after more than 16 min. On the basis of
this, the authors chose 16 min as the optimum incubation time.
Furthermore, taking into account operational convenience,
room temperature (25−28 °C) was selected as the operational
temperature for all experiments. In order to facilitate the
hybridization reaction, the media pH for all the experimental
steps was 7.4. Herein, 1.0 μM was used as the Hg 2+
concentration to optimize the experimental conditions.
Sensitivity of the Sensor. On the basis of the above
standard procedures and optimized assay conditions, various
concentrations of Hg2+ were introduced to the buffer to
evaluate the sensitivity of the “turn-on” fluorescent sensor. The
various concentrations of Hg2+ were 0, 0.2, 0.4, 0.6, 0.8, 1.0, 2.0,
4.0, 6.0, 8.0, 10.0, 20.0, 50.0, 100.0, 200.0, 500.0, and 1000.0
nM. As shown in Figure 3 parts A and B, higher concentrations
of Hg2+ resulted in more fluorescence intensity enhancement. It
Figure 1. (A) Time-gated fluorescence emission spectra of the sensor is worth noting that, even at very low concentration of Hg2+,
without and with 1.0 μM Hg2+. The measure conditions are given the time-gated fluorescence intensity exhibited perceptible
below: hybridization and incubation times, operational temperature, change which indicated that the Hg2+ could be detected with
and media pH are 60 min, 30 min, 25−28 °C, and 7.4, respectively.
The parameters of the spectrofluorimeter are set as λex = 400 nm; λem
high sensitivity in this proposed fluorescent sensor.
= 609 nm; delay time, 0.1 ms; gate time, 1.0 ms; excitation slit, 15 nm; The time-gated fluorescence intensity was found to be linear
emission slit, 20 nm. (B) Fluorescence emission spectra of Mn:CdS/ with the concentration of Hg2+ in the range from 0 to 1 × 10−9
ZnS quantum dots modified strand A in the absence and presence of M and from 1 × 10−9 to 1 × 10−8 M (Figure 3C). The
10, 100, and 500 nM Hg2+, indicating that Hg2+ has negligible effect on equations for the resulting calibration plot were y = 5.75x +
the quenching of fluorescence. 0.43 (eq 1) and y = 1.71x + 5.40 (eq 2) (x was the
D dx.doi.org/10.1021/ac303084d | Anal. Chem. XXXX, XXX, XXX−XXX
Analytical Chemistry Article

Hg2+ detection using fluorescent methods, which is lower than


the standards of the U.S. Environmental Protection Agency
(EPA). The EPA regulates the maximum allowable level of
Hg2+ in drinking water to be 10 nM. In the present method,
higher concentrations of Hg2+ resulted in more free GNPs in
the sample solution, and free GNPs have a quenching effect on
the fluorescence of QDs. As shown in Figure 3, parts B and C,
the proposed method exhibits nonlinear character in the range
from 0 to 1000 nM; such behavior may be caused by the free
GNPs.
For examination of the repeatability of this sensor, three
sample solutions with different concentrations of Hg2+ were
prepared, and the relative standard deviations (RSDs) are about
6.2%, 3.8%, and 2.5% for three independent determinations of
0.8, 6.0, and 50 nM Hg2+ under the optimum conditions,
respectively.
Selectivity of the Sensor. To determine the selectivity of
this protocol, two control experiments were conducted. First,
1.0 μM of Hg2+ and 10 μM of other metal ions including Mn2+,
Ba2+, Ni2+, Cu2+, Ca2+, Cr2+, Co2+, Cd2+, Mg2+, Zn2+, Al3+, Fe3+,
Pb2+, Ag+, and Au3+ were added to the sample solution, and
then the time-gated fluorescence intensity was recorded. As
indicated in Figure 4, only Hg2+ exhibits significant response.

Figure 4. Selectivity of the “turn-on” fluorescent Hg2+ sensor. The


concentrations of Hg2+ and other metal ions are 1.0 and 10 μM. Every
data point was the mean of three measurements. The error bars are the
standard deviation.

Second, 1.0 μM of Hg2+ and 10 μM of other metal ions were


mixed together to form a mixture solution as a sample for
Figure 3. (A) Time-gated fluorescence emission spectra of the sample selectivity testing (Figure 4). The fluorescence intensity was
solution after addition of various concentrations (0−1000 nM) of obviously higher than that of other samples without Hg2+.
Hg2+ in buffer. (B) Plot of time-gated fluorescence intensity as a These results clearly indicated that the approach is not only
function of the concentration of Hg2+. (C) Linear region of panel B. insensitive to other metal ions but also selective toward Hg2+ in
The concentrations of Hg2+ were 0, 0.2, 0.4, 0.6, 0.8, 1.0, 2.0, 4.0, 6.0, their presence.
8.0, and 10 nM (C). Every data point was the mean of three Assay of Hg2+ Concentrations in Environmental Water
measurements. The error bars are the standard deviation. Samples. With excellent sensitivity and selectivity in buffer
solution, the proposed method was further tested with real
concentration of Hg2+, y was the time-gated fluorescence environmental water samples to demonstrate its practical
intensity) with correlation coefficient of 0.9939 and 0.9967, application. The environmental water samples used in the
respectively. According to the standard deviation of 0.35 for the study were tap water and lake water samples. These samples
blank signal with 20 parallel measurements and eq 1, a spiked with Hg2+, with concentrations of 0, 1.0, 2.0, and 10.0
detection limit of approximately 0.18 nM was estimated based nM, were tested using the proposed method and AFS (atomic
on a 3α/slope. This is an exceptionally low detection limit for fluorescence spectrometer). The lake water samples were
E dx.doi.org/10.1021/ac303084d | Anal. Chem. XXXX, XXX, XXX−XXX
Analytical Chemistry


Article

filtered by qualitative filter paper and then centrifuged for 15 REFERENCES


min at 12 000 rpm. The concentrations of total mercury in tap (1) Nolan, E. M.; Lippard, S. J. Chem. Rev. 2008, 163, 3443−3480.
water and lake water samples were 0.045 and 0.056 nM which (2) Zahir, F.; Rizwi, S. J.; Haq, S. K.; Khan, R. H. Environ. Toxicol.
were measured by AFS. The results are summarized in Table 1 Pharmacol. 2005, 20, 351−360.
and show good agreement with those achieved by AFS, (3) Harris, H. H.; Pickering, I. J.; George, G. N. Science 2003, 1203.
indicating that the present sensor can also work in environ- (4) Tchounwou, P. B.; Ayensu, W. K.; Ninashvili, N.; Sutton, D.
mental water samples. Environ. Toxicol. 2003, 18, 149−175.
(5) Hylander, L. D.; Goodsite, M. E. Sci. Total Environ. 2006, 368,
352−370.
Table 1. Determination of Hg2+ in Environmental Water
(6) Butler, O. T.; Cook, J. M.; Harrington, C. F.; Hill, S. J.;
Samples Using the Proposed Method and AFS Rieuwerts, J.; Miles, D. L. J. Anal. At. Spectrom. 2006, 21, 217−243.
Hg2+ (nM) (7) Li, Y.; Chen, C.; Li, B.; Sun, J.; Wang, J.; Gao, Y.; Zhao, Y.; Chai,
Z. J. Anal. At. Spectrom. 2006, 21, 94−96.
sample added proposed method meana ± SDb AFS mean ± SD (8) Leermakers, M.; Baeyens, W.; Quevauviller, P.; Horvat, M. TrAC,
tap water 1 0 c 0.045 ± 0.003 Trends Anal. Chem. 2005, 24, 383−393.
tap water 2 1.0 1.07 ± 0.10 0.995 ± 0.075 (9) Miyake, Y.; Togashi, H.; Tashiro, M.; Yamaguchi, H.; Oda, S.;
tap water 3 2.0 2.06 ± 0.14 2.16 ± 0.10 Kudo, M.; Tanaka, Y.; Kondo, Y.; Sawa, R.; Fujimoto, T.; Achinami,
tap water 4 10.0 10.12 ± 0.20 10.10 ± 0.17 T.; Ono, A. J. Am. Chem. Soc. 2006, 128, 2172−2173.
lake water 1 0 c 0.056 ± 0.005 (10) Tanaka, Y.; Oda, S.; Yamaguchi, H.; Kondo, Y.; Kojima, C.;
lake water 2 1.0 1.06 ± 0.14 0.987 ± 0.068 Ono, A. J. Am. Chem. Soc. 2007, 129, 244−245.
(11) Zhu, Z.; Su, Y.; Li, J.; Li, D.; Zhang, J.; Song, S.; Zhao, Y.; Li, G.;
lake water 3 2.0 2.06 ± 0.15 2.10 ± 0.15
Fan, C. Anal. Chem. 2009, 81, 7660−7666.
lake water 4 10.0 10.15 ± 0.29 10.01 ± 0.055
(12) Du, J.; Liu, M. Y.; Lou, X. H.; Zhao, T.; Wang, Z.; Xue, Y.; Zhao,
a
Mean of four determinations. bSD, standard deviation. cNo Hg2+ J. L.; Xu, Y. S. Anal. Chem. 2012, 84, 8060−8066.
concentration could be detected. (13) Tang, X. M.; Liu, H. X.; Zou, B. H.; Tian, D. B.; Huang, H.
Analyst 2012, 137, 309−311.


(14) Wang, H.; Wang, Y. X.; Jin, J. Y.; Yang, R. H. Anal. Chem. 2008,
80, 9021−9028.
CONCLUSIONS (15) Freeman, R.; Finder, T.; Willner, I. Angew. Chem., Int. Ed. 2009,
The authors have developed a “turn-on” fluorescent sensor for 48, 7818−7821.
determination of Hg2+ in aqueous media with excellent (16) Kong, L. T.; Wang, J.; Zheng, G. C.; Liu, J. H. Chem. Commun.
sensitivity and selectivity by using Hg2+-mediated T−Hg2+−T 2011, 47, 10389−10391.
pairs, long-lifetime fluorescence QDs, and GNPs. It combines (17) He, X. X.; Qing, A. H.; Wang, K. M.; Zou, Z.; Shi, H.; Huang, J.
Anal. Methods 2012, 4, 345−347.
the advantages of specific and stable binding interactions
(18) Liu, J. W.; Lu, Y. J. Am. Chem. Soc. 2007, 129, 9838−9839.
between Hg2+ and thymine, the unique photophysical proper- (19) Ren, X. S.; Xu, Q. H. Langmuir 2009, 25, 29−31.
ties and long-lifetime fluorescence of QDs, and the excellent (20) Wang, Z. D.; Lee, J. H.; Lu, Y. Chem. Commun. 2008, 6005−
quenching performance of GNPs. The detection limit (LOD, 6007.
0.18 nM) is much lower than the EPA limit of Hg2+ in (21) Yan, F. Y.; Cao, D. L.; Yang, N.; Yu, Q. H.; Wang, M.; Chen, L.
drinkable water (<10 nM). In addition, the fluorescent sensor Sens. Actuators, B 2012, 162, 313−320.
showed outstanding selectivity for Hg2+ against other metal (22) Ye, B. C.; Yin, B. C. Angew. Chem., Int. Ed. 2008, 47, 8386−
ions. The sensor also can be used for the detection of Hg2+ in 8389.
real environmental water samples. The authors believed that (23) Li, M.; Wang, Q. Y.; Shi, X. D.; Hornak, L. A.; Wu, N. Q. Anal.
the fluorescent sensor using the “turn-on” mode provides a Chem. 2011, 83, 7061−7065.
(24) Mattoussi, H.; Mauro, J. M.; Goldman, E. R.; Anderson, G. P.;
promising method for Hg2+ detection in environmental and
Sundar, V. C.; Mikulec, F. C.; Bawendi, M. G. J. Am. Chem. Soc. 2000,
other applications.


122, 12142−12150.
(25) Freeman, R.; Willner, I. Chem. Soc. Rev. 2012, 41, 4067−4085.
AUTHOR INFORMATION (26) Li, Y. F.; Han, M.; Baia, H. Y.; Wu, Y.; Daia, Z. H.; Bao, J. C.
Corresponding Author Electrochim. Acta 2011, 56, 7058−7063.
*E-mail: cgniu@hnu.edu.cn, cgniu@hotmail.com (C.N.); (27) Kamat, P. V.; Barazzouk, S.; Hotchandani, S. Angew. Chem., Int.
zgming@hnu.edu.cn (G.Z.). Phone: +86-731-88823820. Fax: Ed. 2002, 41, 2764−2767.
+86-731-88822829. (28) Fan, C. H.; Wang, S.; Hong, J. W.; Bazan, G. C.; Plaxco, K. W.;
Heeger, A. J. Proc. Natl. Acad. Sci. U.S.A. 2003, 100, 6297−6301.
Notes (29) Liu, H.; Liang, G.; Abdel-Halim, E. S.; Zhu, J. J. Anal. Methods
The authors declare no competing financial interest.


2011, 3, 1797−1801.
(30) Huang, C. C.; Chiu, S. H.; Huang, Y. F.; Chang, H. T. Anal.
ACKNOWLEDGMENTS Chem. 2007, 79, 4798−4804.
Special thanks are given to Mr. Y. Charles Cao and his group of (31) Mandal, G.; Bardhan, M.; Ganguly, T. J. Phys. Chem. C 2011,
the University of Florida for kind assistance in synthesis of Mn- 115, 20840−20848.
doped CdS/ZnS core/shell QDs. This work was financially (32) Ray, P. C.; Fortner, A.; Darbha, G. K. J. Phys. Chem. B 2006,
supported by the National Natural Science Foundation of 110, 20745−20748.
(33) Yang, Y.; Chen, O.; Angerhofer, A.; Cao, Y. C. J. Am. Chem. Soc.
China (20977026), the National 863 High Technology 2006, 128, 12428−12429.
Research Foundation of China (2006AA06Z407), the Research (34) Yang, Y.; Chen, O.; Angerhofer, A.; Cao, Y. C. J. Am. Chem. Soc.
Fund for the Doctoral Program of Higher Education of China 2008, 130, 15649−15661.
(20090161110009), the project sponsored by SRF for ROCS, (35) Norris, D. J. Nano Lett. 2001, 1, 3−7.
SEM (521294018), and the Fundamental Research Funds for (36) Norris, D. J.; Efros, A. L.; Erwin, S. C. Science 2008, 319, 1776−
the Central Universities of Hunan University (531107040375). 1779.

F dx.doi.org/10.1021/ac303084d | Anal. Chem. XXXX, XXX, XXX−XXX


Analytical Chemistry Article

(37) Nag, A.; Chakraborty, S.; Sarma, D. D. J. Am. Chem. Soc. 2008,
130, 10605−10611.
(38) Mitchell, G. P.; Mirkin, C. A.; Letsinger, R. L. J. Am. Chem. Soc.
1999, 121, 8122−8123.
(39) Katherine, C. G.; Freeman, R. G.; Michael, B. H.; Michael, J. N.
Anal. Chem. 1995, 67, 735−743.
(40) Haiss, W.; Thanh, N. T. K.; Aveyard, J.; Fernig, D. G. Anal.
Chem. 2007, 79, 4215−4221.
(41) Storhoff, J. J.; Elghanian, R.; Mucic, R. C.; Mirkin, C. A.;
Letsinger, R. L. J. Am. Chem. Soc. 1998, 120, 1959−1964.

G dx.doi.org/10.1021/ac303084d | Anal. Chem. XXXX, XXX, XXX−XXX

You might also like