• Photonics Research
  • Vol. 9, Issue 2, 213 (2021)
Xiongbin Wang1、2、†, Qiushi Wang3、†, Yulong Chen1、4、†, Jiagen Li5, Ruikun Pan3, Xing Cheng4, Kar Wei Ng1, Xi Zhu5, Tingchao He6、7、*, Jiaji Cheng3、8、*, Zikang Tang1、9、*, and Rui Chen2、10、*
Author Affiliations
  • 1Institute of Applied Physics and Materials Engineering, University of Macau, Avenida da Universidade, Taipa, Macau 999078, China
  • 2Department of Electrical and Electronic Engineering, Southern University of Science and Technology, Shenzhen 518055, China
  • 3School of Materials Science and Engineering, Hubei University, Wuhan 430062, China
  • 4Department of Materials Science and Engineering, Southern University of Science and Technology, Shenzhen 518055, China
  • 5Shenzhen Institute of Artificial Intelligence and Robotics for Society (AIRS), Shenzhen 518172, China
  • 6College of Physics and Optoelectronic Engineering, Shenzhen University, Shenzhen 518060, China
  • 7e-mail: tche@szu.edu.cn
  • 8e-mail: jiajicheng@hubu.edu.cn
  • 9e-mail: zktang@um.edu.mo
  • 10e-mail: chenr@sustech.edu.cn
  • show less
    DOI: 10.1364/PRJ.413592 Cite this Article Set citation alerts
    Xiongbin Wang, Qiushi Wang, Yulong Chen, Jiagen Li, Ruikun Pan, Xing Cheng, Kar Wei Ng, Xi Zhu, Tingchao He, Jiaji Cheng, Zikang Tang, Rui Chen. Metal-to-ligand charge transfer chirality-based sensing of mercury ions[J]. Photonics Research, 2021, 9(2): 213 Copy Citation Text show less

    Abstract

    Chiral ligand conjugated transition metal oxide nanoparticles (NPs) are a promising platform for chiral recognition, biochemical sensing, and chiroptics. Herein, we present chirality-based strategy for effective sensing of mercury ions via ligand-induced chirality derived from metal-to-ligand charge transfer (MLCT) effects. The ligand competition effect between molybdenum and heavy metal ions such as mercury is designated to be essential for MLCT chirality. With this know-how, mercury ions, which have a larger stability constant (Kf) than molybdenum, can be selectively identified and quantified with a limit of detection (LOD) of 0.08 and 0.12 nmol/L for D-cysteine and L-cysteine (Cys) capped MoO2 NPs. Such chiral chemical sensing nanosystems would be an ideal prototype for biochemical sensing with a significant impact on the field of biosensing, biological systems, and water research-based nanotoxicology.

    1. INTRODUCTION

    Ligand-induced chirality in semiconductor nanoparticles (NPs) has revolutionized the motif of inorganic material-based nanotechnology, because it timely opens the floodgate of these materials for promising applications in chiral recognitions and synthesis [15], bioimaging and display devices [615], and metamaterials in advanced optical devices [1619] simply via interactions between the chiral ligands and achiral core [1929]. Recently, ligand-induced chiral transition metal oxide ceramics have attracted tremendous interest not only due to their existing applications in optoelectronics [12,18,30] and biomedicine [31,32], but also their ability to provide incisive information on the metal-ligand interactions in terms of chiroptical activities. With both chemical linkage and coordination effects, electronic transitions between metal and ligands induce chiral responses in the visible region with a g-factor up to the order of 103, which is 1 to 2 orders higher than that of traditional chiral ligand capped cadmium chalcogenide, where g-factor is defined as Δε/ε=ΔA/A, and ΔA is the absorbance difference between left- and right-handed circularly polarized light. Kotov and his coworkers [18], for instance, demonstrated that paramagnetic Co3O4 NPs with crystal lattice distortions caused by the chiral ligands exhibited both chiroptical activity in the visible range and intensive magnetic field-induced light modulation in the ultraviolet (UV) range, offering a versatile tool box for new technologies and knowledge at the nexus of chirality and magnetism. Moreover, their follow-up research work [33] showed the observation of induced chirality in WO3x·H2O NPs via using proline (Pro) and aspartic acid (Asp) as surface ligands. The formed C-O-W linkages and weak coordination between amino groups and the core are found to be essential for the chiral metal-to-ligand charge transfer (MLCT) in the visible range and chiroplasmonic band in the near-infrared (NIR) region, respectively. To go further, the mentioned strong MLCT band is as well witnessed in our previous work [34] in a sub-stoichiometric chiral Cys-capped MoO3x system with a large g-factor close to 7×103, in which the MLCT band transition is contributed by the transition from the metal-δ orbitals coupling with ligand-based π and π* orbitals. The g-factor originated from the MLCT band is about 100 times higher compared with the g-factor that comes from chiroplasmonic band transition in the NIR region [34]. Such strong chiroptical activity naturally motivates us to raise questions as this: is it possible to affect the MLCT chirality simply via the ligands, and can the materials be applied in the field of nanotechnology with specific applications such as chiral sensing?

    To address the question, it is demonstrated herein chiral Cys reduced and capped MoO2 NPs (L-/D-Cys-MoO2) can serve as platform for ultralow concentration detection of mercury ions (Hg2+) in water based on circular dichroism (CD) spectropolarimetry. The observed low limit of detection (LOD) and high selectivity on Hg2+ due to the large stability constant (log Kf) of Mo(cysteine)2 highlight the potential of chiral MoO2 NPs in the realm of chiral sensing and recognition. On the other hand, owing to the ligand competition effect, the surface ligand density of the chiral MoO2 will be greatly impacted during Hg2+ introduction. The variation of ligand density linked to MLCT chirality with respect to CD response would unravel the chirogenesis and fundamental relations of ligand density versus MLCT chirality, providing intuitive experimental basis for studying the chiral electronic transitions in transition metal oxide ceramics.

    2. EXPERIMENT SECTION

    A. Materials

    Molybdenum disulfide (MoS2, 99.5%) powder and hydrogen peroxide (H2O2, GR, mass fraction 30% in H2O) solution were purchased from Aladdin. L-Cys (98%), D-Cys (99%), and all soluble heavy metallic salts were obtained from Sigma-Aldrich. All chemicals were not purified for using as received. All heavy metal ion solution was prepared by adding the corresponding salt in water. The water that was used in all experiments had a resistivity higher than 18  MΩ·cm1.

    B. Synthesis of Aqueous Molybdenum Trioxide Nanoparticles

    The preparation of molybdenum trioxide MoO3 NPs was reported previously [1,34]. First of all, 80 mg pristine MoS2 powder was dissolved in 46.25 mL water in a beaker. Then, the dispersion was treated with sonication at room temperature. 3.75 mL H2O2 solution was added into the dispersion subsequently. The mixture was kept constantly stirred until the color turned from black to yellow. It would last about 20 h. Then, the dispersion was heated at about 70°C to remove excess H2O2. After about 1 h, the color of the dispersion would turn to transparent, which means the successful synthesis of MoO3 NPs.

    C. Preparation of Chiral MoO2 Nanodots

    The chiral MoO2 nanodots were prepared according to the procedure given in the literature [1,31]. The as-prepared MoO3 nanocrystals solution could be reduced and capped with the presence of chiral Cys molecules. In a typical process, 65 mg D-Cys was added into 1.5 mL MoO3 solution that was obtained previously. Subsequently, the mixture was treated with sonication for 5 min. After that, the mixture was placed in the dark for one day to make sure that it could react sufficiently. The same treatment was carried out for obtaining D-Cys molecule capped MoO2 nanodots.

    D. Purification of Chiral Cys Capped MoO2 Nanodots

    After obtaining chiral molecule capped MoO2 nanodots, the solution should be purified for further use. As in a typical process, the resulting solution was treated with centrifugation at 14,000 r/min for 20 min. Then, the supernatant was taken out from the centrifuge tube. The precipitate was dissolved in water. The mixture was centrifuged again at 6000 r/min for 10 min. After that, the supernatant was re-dissolved into another centrifuge tube. Finally, the mixture was again centrifuged at 5000 r/min for 5 min. The precipitate was dissolved in water for further use.

    E. Characterizations

    The UV-visible absorption measurement was carried out using a TU-1901 double-beam UV-Vis spectrophotometer (Beijing Purkine General Instrument Co., Ltd., China). For X-ray photoelectron spectroscopy (XPS) measurements, the samples were prepared by dropping the as-prepared solution on a silicon substrate, and the experiment was performed on a Thermo Scientific K-Alpha system. CD experiments were realized on a JASCO J-1500 CD spectrometer. The scan rate was 20 nm/min, and the data pitch was 0.1 nm. For all CD experiments, Milli-Q water with a quartz cuvette (0.1 cm optical path length from Hellma) was used. The transmission electron microscopy (TEM) pictures were captured by a Tecnai F30 microscope with operating voltage at 300 kV. A thermogravimetric analysis (TGA) experiment was conducted via Perkin Elmer STA 6000. The sample for measurement was dried under inert gas atmosphere. The measurement region is from 30 K to 800 K with a heating speed of 5°C per minute and 20°C per minute for the cooling process.

    F. Stimulation Method

    The nanocluster establishment and quantum chemical calculations were conducted with density functional theory (DFT). Ground state geometries were initiated and optimized at minimal energy state. The UV and CD simulated spectra were calculated at time-dependent density functional theory (TD-DFT) based on Gaussian 09 software. All the TD-DFT calculations were utilizing a B3LYP and LanL2DZ basis that was set for all the elements in simulation [3537]. The small MoO2 NPs capped with L- and D-Cys nanoclusters were constructed as a tetragonal crystal structure to form Mo4O8 neutral charge geometry to simulate the interactions between the molybdenum oxide cluster and chiral light.

    G. Calculation of Ligand Density

    The ligand density of NPs could be determined by the mass loss curve owing to the desorption and decomposition of the chemisorbed molecules [33,38]. First of all, the NPs for calculating were assumed to be spherical, and the diameter could be regarded as the average value of each NP obtained from the TEM image. The mass of an individual MoO2 NP could be expressed as mMoO2=πρD36,where D is the average diameter of NPs and ρ is the density of the NP. The total amount of ligands (Wcys) and MoO2 (WMoO2) in the system was obtained as Wcys=Wφ,WMoO2=WWcys,where W is the total mass of sample for TGA measurement and φ is the mass fraction of Cys molecules on the NP surface. Then, the total number (N) for each sample was determined as N=WMoO2mMoO2.Finally, the ligand density could be determined as Ligand density=WcysMcysN·Na·1πD2=φ1φ·ρD6Mcys·Na,where Na is the Avogadro constant and Mcys is the molecular weight of Cys. Here, D was 22.2 nm according to the TEM size distribution. ρ was assumed to be equal to the bulk MoO2 density (6.47  g/cm3). Mcys was equal to 121.15  g/cm3 (CAS52-90-4).

    3. RESULTS AND DISCUSSION

    Illustration of the synthesis process of chiral Cys-MoO2 NPs and their application for Hg2+ sensing.

    Figure 1.Illustration of the synthesis process of chiral Cys-MoO2 NPs and their application for Hg2+ sensing.

    (a) TEM image of D-Cys-MoO2 NPs. The corresponding scale bar is 100 nm. (b) Histogram distribution of the diameter of NPs. Measurements of (c) circular dichroism spectrum and (d) absorption spectrum of MoO3 (black line), L-Cys-MoO2 (red line), and D-Cys-MoO2 NPs (blue line). XPS spectra of (e) MoO3 NPs and (f) D-Cys-MoO2 NPs with deconvoluted molybdenum 3d peaks. The blue and orange peak areas are corresponding to the different valence states of Mo(VI) and Mo(IV), respectively.

    Figure 2.(a) TEM image of D-Cys-MoO2 NPs. The corresponding scale bar is 100 nm. (b) Histogram distribution of the diameter of NPs. Measurements of (c) circular dichroism spectrum and (d) absorption spectrum of MoO3 (black line), L-Cys-MoO2 (red line), and D-Cys-MoO2 NPs (blue line). XPS spectra of (e) MoO3 NPs and (f) D-Cys-MoO2 NPs with deconvoluted molybdenum 3d peaks. The blue and orange peak areas are corresponding to the different valence states of Mo(VI) and Mo(IV), respectively.

    The chiroptical property of the MoO3 NPs and Cys capped MoO2 NPs is characterized by a CD spectrometer along with absorption measurement [Figs. 2(c) and 2(d)]. Without chiral ligands, the MoO3 NPs are inactive in CD, while in the case of Cys reduced Cys-MoO2 NPs, strong CD responses are recorded with opposite line shape depending on the enantiomer of chiral molecules used for the synthesis. Due to the MLCT effect, the observed CD signal is located in the visible range (350–650 nm) accompanied with merging UV-Vis absorption peaks within the same region (around 380 and 560 nm). The g-factors, which are calculated based on measured CD and absorption signal, are about 4.1×103 at 384 nm and 4×103 at 568 nm, respectively. The XPS survey scan is performed for both MoO3 and DCys-MoO2 samples. Figures 2(e) and 2(f) also plot the Mo 3d orbital spectra before and after reductive treatment under Cys molecules. Figure 2(e) shows two apparent peaks located at 233.15 and 236.35 eV, which correspond to the 3d5/2 and 3d3/2 orbitals for the Mo(VI), respectively. The result indicates that typical α-MoO3 has been obtained. After the Cys reduction, the XPS peaks of Mo show obvious downshift, which move to 229 and 232.8 eV, respectively, as illustrated in Fig. 2(f) [34,39]. These binding energies are indexed as 3d5/2 and 3d3/2 orbitals for the Mo(IV). It confirms that excessive Cys could completely reduce the Mo in MoO3 to the IV state without any intermediate states.

    Chiroptical sensing of Hg2+ using Cys-MoO2 NPs. (a) CD and (b) absorption measurements for Hg2+ mixing with aqueous D-Cys-MoO2 and L-Cys-MoO2 NPs solution. The concentration of Hg2+ in the mixture varied from 0.1 nM to 30 nM. (c) Calculated g-factor curves of specimens in (a). (d) Differences of g-factor [values at 384 nm shown in (c)] versus Hg2+ concentration and corresponding fitting curve. The inset image is the calibration plot.

    Figure 3.Chiroptical sensing of Hg2+ using Cys-MoO2 NPs. (a) CD and (b) absorption measurements for Hg2+ mixing with aqueous D-Cys-MoO2 and L-Cys-MoO2 NPs solution. The concentration of Hg2+ in the mixture varied from 0.1 nM to 30 nM. (c) Calculated g-factor curves of specimens in (a). (d) Differences of g-factor [values at 384 nm shown in (c)] versus Hg2+ concentration and corresponding fitting curve. The inset image is the calibration plot.

    Apparently, with the increase of Hg2+ concentration, the recorded CD and UV signals, no matter for L-Cys-MoO2 or D-Cys-MoO2 systems, decrease gradually [Figs. 3(a) and 3(b)], which indicates the chemical reactions are initiated between Hg2+ and Cys molecules that are located on the MoO2 surface.

    Considering that the absorption of the chiral system may change probably due to the volume increment or aggregations caused by introduction of Hg2+, anisotropic g-factor is used here for evaluating the sensing performance to rule out the interference of concentration variation. At low Hg2+ concentration, the CD signal of the mixture is indistinguishable for both D-Cys-MoO2 and L-Cys-MoO2 systems [as shown in Fig. 3(a)]. With the assistance of g-factor, the change of chiroptical property is better elucidated even when an ultrasmall amount (0.1 and 1 nM, 1 nM = 1 nmol/L) of Hg2+ is added to the chiral system as shown in Fig. 3(c) (black and red lines, for example). Moreover, Fig. 3(c) shows the relation of Hg2+ versus g-factor for the optical activity, suggesting a clear decreasing of g-factor values with the increase of Hg2+. Figure 3(d) states the relationship between the increment of g-factor Δg=g0g at around 384 nm and Hg2+ concentration, where g and g0 are the anisotropy factor without and with the presence of Hg2+ in solutions. The scatter plot and fitting curves of Δg show a growth tendency with the increase of Hg2+ concentration, which indicates that Cys molecules loaded at the surface of MoO2 NPs had reacted with Hg2+ [Fig. 3(d)]. A linear increment of anisotropy factor at 384 nm versus the Hg2+ concentration ranging from 0.1 to 4 nM is established with correlation coefficients of 0.992 (for D-Cys-MoO2 system) and 0.990 (for L-Cys-MoO2 system). The LOD is determined to be 0.12 nM and 0.08 nM for L-Cys-MoO2 and D-CysMoO2, respectively. The LOD for Hg2+ detection is evaluated by 3σ/S, where σ is the standard deviation of g-factor measured in the absence of Hg2+, and S is the slope of the calibrated plot. For comparison, the analytical characteristics of some Hg2+ sensors are summarized in Table 1. Obviously, the LOD of our sensing system is comparable to or better than the previously reported Hg2+ sensors based on surface-enhanced Raman scattering (SERS), absorption, PL, electrochemical and CD spectrum.

    Comparison of the Proposed Probe with Previously Reported Hg2+ Sensors Based on Different Methodsa

    MethodsSystemDetection Range (nM)LOD (nM)Ref.
    SERSAu NPs/rGO/SiO2/Si0.1–60000.1[42]
    Au TNAs/graphene/Au NPs1–45,0008.3[43]
    4-MPY-Ag NPs1–1000.34[44]
    AbsorptionDNA-Au NPs0–5000500[45]
    N-T-Au NPs50–2500.8[46]
    Hcy-SH0–100072[47]
    PLAPBA-MoS25–41,0001.8[48]
    N-doped-CNDs0–300,00080[49]
    DNA-SWNTs50–800014.5[50]
    ElectrochemistryHNTs-Fe3O4-MnO22.5–7501[51]
    MSO-Au NPs0–1000.5[52]
    DNA-Fc1–20000.5[53]
    CDDNA-Au NRs0.25–500.15[54]
    Ag-L-Cys NPs0–10009[55]
    D-Cys-MoO2NPs0.1–300.08This work
    L-Cys-MoO2NPs0.1–300.12This work

    rGO, reduced graphene oxide; TNAs, triangular nanoarrays; 4-MPY, 4-mercaptopyridine; N-T, thymine derivative; APBA, 3-aminobenzeneboronic acid; CNDs, carbon nanodots; SWNTs, single walled carbon nanotubes; HNTs, halloysite nanotubes; MSO, mercury-specific oligonucleotide; Fc, ferrocene; NRs, nanorods.

    (a) TGA curves of L-Cys-MoO2 NPs with different amounts of mercury after dialysis. (b) Ligand density varies with mercury ion concentrations. (c) CD and (d) absorption spectra of pure L-Cys-MoO2 NPs as well as L-Cys-MoO2 mixing with Hg2+ (10 nM in the mixture) under different reaction times.

    Figure 4.(a) TGA curves of L-Cys-MoO2 NPs with different amounts of mercury after dialysis. (b) Ligand density varies with mercury ion concentrations. (c) CD and (d) absorption spectra of pure L-Cys-MoO2 NPs as well as L-Cys-MoO2 mixing with Hg2+ (10 nM in the mixture) under different reaction times.

    TD-DFT simulation for different amounts of D-Cys capped Mo4O8 nanoclusters. Calculated frontier molecular orbital of (a), (c) HOMO and (b), (d) LUMO for one Cys molecule capped and six Cys molecule capped Mo4O8 nanoclusters. Calculated (e) CD spectra and (f) absorption spectra.

    Figure 5.TD-DFT simulation for different amounts of D-Cys capped Mo4O8 nanoclusters. Calculated frontier molecular orbital of (a), (c) HOMO and (b), (d) LUMO for one Cys molecule capped and six Cys molecule capped Mo4O8 nanoclusters. Calculated (e) CD spectra and (f) absorption spectra.

    Selectivity of D-Cys-MoO2-based Hg2+ sensor. (a) CD spectra of D-Cys-MoO2 solution mixed with different heavy metal ions: Zn2+, Cd2+, Pb2+, Ag+, Cu2+, and Hg2+. (b) CD signal at 384 nm of mixtures measured in (a). The concentrations of all metal ions are settled at 10 nM.

    Figure 6.Selectivity of D-Cys-MoO2-based Hg2+ sensor. (a) CD spectra of D-Cys-MoO2 solution mixed with different heavy metal ions: Zn2+, Cd2+, Pb2+, Ag+, Cu2+, and Hg2+. (b) CD signal at 384 nm of mixtures measured in (a). The concentrations of all metal ions are settled at 10 nM.

    4. CONCLUSION

    In summary, Cys-induced optically active MoO2 NPs are synthesized and applied for Hg2+ detection with high precision. Due to the big stability constant of Hg(cysteine)2, such a chiral sensing system is exceptionally sensitive to Hg2+ while inactive to other traditional heavy metal ions that coexist in waste water. Further CD observations, TGA analysis, and TD-DFT-based simulations confirm the ligand competition phenomenon between mercury and molybdenum, which unveils the underlying chirogenesis of MLCT chirality. Such transition metal oxide-based chiral NPs would be potential candidates for effective sensing of Hg2+ providing new horizons to the areas of biochemical sensing, chiroptics, and environmental remediation.

    References

    [1] J. Hao, Y. Li, X. Xu, F. Zhao, R. Pan, J. Li, H. Liu, K. Wang, J. Li, X. Zhu, M.-H. Delville, M. Zhang, T. He, J. Cheng. Metal-to-ligand charge transfer chirality sensing of D-glucose assisted with GOX-based enzymatic reaction. Adv. Mater. Technol., 5, 2000138(2020).

    [2] I. V. Martynenko, V. A. Kuznetsova, I. K. Litvinov, A. O. Orlova, V. G. Maslov, A. V. Fedorov, A. Dubavik, F. Purcell-Milton, Y. K. Gun’ko, A. V. Baranov. Enantioselective cellular uptake of chiral semiconductor nanocrystals. Nanotechnology, 27, 075102(2016).

    [3] D. Meng, W. Ma, X. Wu, C. Xu, H. Kuang. DNA-driven two-layer core–satellite gold nanostructures for ultrasensitive microRNA detection in living cells. Small, 16, 2000003(2020).

    [4] A. Kühnle, T. R. Linderoth, B. Hammer, F. Besenbacher. Chiral recognition in dimerization of adsorbed cysteine observed by scanning tunnelling microscopy. Nature, 415, 891-893(2002).

    [5] S. Li, J. Liu, N. S. Ramesar, H. Heinz, L. Xu, C. Xu, N. A. Kotov. Single- and multi-component chiral supraparticles as modular enantioselective catalysts. Nat. Commun., 10, 4826(2019).

    [6] J. E. Govan, E. Jan, A. Querejeta, N. A. Kotov, Y. K. Gun’Ko. Chiral luminescent CdS nano-tetrapods. Chem. Commun., 46, 6072-6074(2010).

    [7] M. P. Moloney, Y. K. Gun’Ko, J. M. Kelly. Chiral highly luminescent CdS quantum dots. Chem. Commun., 38, 3900-3902(2007).

    [8] O. Cleary, F. Purcell-Milton, A. Vandekerckhove, Y. K. Gun’Ko. Chiral and luminescent TiO2 nanoparticles. Adv. Opt. Mater., 5, 1601000(2017).

    [9] J. Ahn, E. Lee, J. Tan, W. Yang, B. Kim, J. Moon. A new class of chiral semiconductors: chiral-organic-molecule-incorporating organic–inorganic hybrid perovskites. Mater. Horiz., 4, 851-856(2017).

    [10] J. Hao, Y. Li, J. Miao, R. Liu, J. Li, H. Liu, Q. Wang, H. Liu, M.-H. Delville, T. He, K. Wang, X. Zhu, J. Cheng. Ligand-induced chirality in asymmetric CdSe/CdS nanostructures: a close look at chiral tadpoles. ACS Nano, 14, 10346-10358(2020).

    [11] J. Cheng, J. Hao, H. Liu, J. Li, J. Li, X. Zhu, X. Lin, K. Wang, T. He. Optically active CdSe-dot/CdS-rod nanocrystals with induced chirality and circularly polarized luminescence. ACS Nano, 12, 5341-5350(2018).

    [12] B. Zhao, H. Yu, K. Pan, Z. A. Tan, J. Deng. Multifarious chiral nanoarchitectures serving as handed-selective fluorescence filters for generating full-color circularly polarized luminescence. ACS Nano, 14, 3208-3218(2020).

    [13] L. Wang, Y. Xue, M. Cui, Y. Huang, H. Xu, C. Qin, J. Yang, H. Dai, M. Yuan. A chiral reduced-dimension perovskite for an efficient flexible circularly polarized light photodetector. Angew. Chem. (Int. Ed.), 59, 6442-6450(2020).

    [14] C. Hao, X. Wu, M. Sun, H. Zhang, A. Yuan, L. Xu, C. Xu, H. Kuang. Chiral core–shell upconversion nanoparticle@MOF nanoassemblies for quantification and bioimaging of reactive oxygen species in vivo. J. Am. Chem. Soc., 141, 19373-19378(2019).

    [15] J. Yeom, P. P. G. Guimaraes, H. M. Ahn, B.-K. Jung, Q. Hu, K. McHugh, M. J. Mitchell, C.-O. Yun, R. Langer, A. Jaklenec. Chiral supraparticles for controllable nanomedicine. Adv. Mater., 32, 1903878(2020).

    [16] J. Govan, Y. K. Gun’ko. Recent progress in chiral inorganic nanostructures. Nanoscience, 3, 1-30(2016).

    [17] F. P. Milton, J. Govan, M. V. Mukhina, Y. K. Gun’Ko. The chiral nano-world: chiroptically active quantum nanostructures. Nanoscale Horiz., 1, 14-26(2015).

    [18] J. Yeom, U. S. Santos, M. Chekini, M. Cha, A. F. de Moura, N. A. Kotov. Chiromagnetic nanoparticles and gels. Science, 359, 309-314(2018).

    [19] W. Ma, L. Xu, A. F. de Moura, X. Wu, H. Kuang, C. Xu, N. A. Kotov. Chiral inorganic nanostructures. Chem. Rev., 117, 8041-8093(2017).

    [20] W. A. Paiva-Marques, F. Reyes Gómez, O. N. Oliveira, J. R. Mejía-Salazar. Chiral plasmonics and their potential for point-of-care biosensing applications. Sensors, 20, 944(2020).

    [21] J. Cheng, E. H. Hill, Y. Zheng, T. He, Y. Liu. Optically active plasmonic resonance in self-assembled nanostructures. Mater. Chem. Front., 2, 662-678(2018).

    [22] Y. Dong, Y. Zhang, X. Li, Y. Feng, H. Zhang, J. Xu. Chiral perovskite: chiral perovskites: promising materials toward next-generation optoelectronics. Small, 15, 1970209(2019).

    [23] X. Gao, B. Han, X. Yang, Z. Tang. Perspective of chiral colloidal semiconductor nanocrystals: opportunity and challenge. J. Am. Chem. Soc., 141, 13700-13707(2019).

    [24] Y. Y. Lee, R. M. Kim, S. W. Im, M. Balamurugan, K. T. Nam. Plasmonic metamaterials for chiral sensing applications. Nanoscale, 12, 58-66(2020).

    [25] A. Visheratina, N. A. Kotov. Inorganic nanostructures with strong chiroptical activity. CCS Chem., 2, 583-604(2020).

    [26] Z. Wang, F. Cheng, T. Winsor, Y. Liu. Optical chiral metamaterials: a review of the fundamentals, fabrication methods and applications. Nanotechnology, 27, 412001(2016).

    [27] R. Zhang, Y. Zhou, X. Yan, K. Fan. Advances in chiral nanozymes: a review. Microchim. Acta, 186, 782(2019).

    [28] X. Zhao, S.-Q. Zang, X. Chen. Stereospecific interactions between chiral inorganic nanomaterials and biological systems. Chem. Soc. Rev., 49, 2481-2503(2020).

    [29] Y. Li, X. Wang, J. Miao, J. Li, X. Zhu, R. Chen, Z. Tang, R. Pan, T. He, J. Cheng. Chiral transition metal oxides: synthesis, chiral origins, and perspectives. Adv. Mater., 32, 1905585(2020).

    [30] J. Lv, D. Ding, X. Yang, K. Hou, X. Miao, D. Wang, B. Kou, L. Huang, Z. Tang. Biomimetic chiral photonic crystals. Angew. Chem. (Int. Ed.), 58, 7783-7787(2019).

    [31] Y. Li, Z. Miao, Z. Shang, Y. Cai, J. Cheng, X. Xu. A visible- and NIR-light responsive photothermal therapy agent by chirality-dependent MoO3−x nanoparticles. Adv. Funct. Mater., 30, 1906311(2020).

    [32] S. Li, M. Sun, C. Hao, A. Qu, X. Wu, L. Xu, C. Xu, H. Kuang. Chiral CuxCoyS nanoparticles under magnetic field and NIR light to eliminate senescent cells. Angew. Chem. (Int. Ed.), 59, 13915-13922(2020).

    [33] S. Jiang, M. Chekini, Z.-B. Qu, Y. Wang, A. Yeltik, Y. Liu, A. Kotlyar, T. Zhang, B. Li, H. V. Demir, N. A. Kotov. Chiral ceramic nanoparticles and peptide catalysis. J. Am. Chem. Soc., 139, 13701-13712(2017).

    [34] Y. Li, J. Cheng, J. Li, X. Zhu, T. He, R. Chen, Z. Tang. Tunable chiroptical properties from the plasmonic band to metal–ligand charge transfer band of cysteine-capped molybdenum oxide nanoparticles. Angew. Chem. (Int. Ed.), 57, 10236-10240(2018).

    [35] C. Lee, W. Yang, R. G. Parr. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B, 37, 785-789(1988).

    [36] P. J. Hay, W. R. Wadt. Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms Sc to Hg. J. Chem. Phys., 82, 270-283(1985).

    [37] W. R. Wadt, P. J. Hay. Ab initio effective core potentials for molecular calculations. Potentials for main group elements Na to Bi. J. Chem. Phys., 82, 284-298(1985).

    [38] B. M. Amoli, S. Gumfekar, A. Hu, Y. N. Zhou, B. Zhao. Thiocarboxylate functionalization of silver nanoparticles: effect of chain length on the electrical conductivity of nanoparticles and their polymer composites. J. Mater. Chem., 22, 20048-20056(2012).

    [39] Y. Sun, X. Hu, J. C. Yu, Q. Li, W. Luo, L. Yuan, W. Zhang, Y. Huang. Morphosynthesis of a hierarchical MoO2 nanoarchitecture as a binder-free anode for lithium-ion batteries. Energy Environ. Sci., 4, 2870-2877(2011).

    [40] G. Berthon. Critical evaluation of the stability constants of metal complexes of amino acids with polar side chains (technical report). Pure Appl. Chem., 67, 1117-1240(1995).

    [41] D. Liu, W. Qu, W. Chen, W. Zhang, Z. Wang, X. Jiang. Highly sensitive, colorimetric detection of mercury(II) in aqueous media by quaternary ammonium group-capped gold nanoparticles at room temperature. Anal. Chem., 82, 9606-9610(2010).

    [42] X. Ding, L. Kong, J. Wang, F. Fang, D. Li, J. Liu. Highly sensitive SERS detection of Hg2+ ions in aqueous media using gold nanoparticles/graphene heterojunctions. ACS Appl. Mater. Interfaces, 5, 7072-7078(2013).

    [43] X. Zhang, Z. Dai, S. Si, X. Zhang, W. Wu, H. Deng, F. Wang, X. Xiao, C. Jiang. Ultrasensitive SERS substrate integrated with uniform subnanometer scale ‘hot spots’ created by a graphene spacer for the detection of mercury ions. Small, 13, 1603347(2017).

    [44] L. Chen, N. Qi, X. Wang, L. Chen, H. You, J. Li. Ultrasensitive surface-enhanced Raman scattering nanosensor for mercury ion detection based on functionalized silver nanoparticles. RSC Adv., 4, 15055-15060(2014).

    [45] X. Xu, J. Wang, K. Jiao, X. Yang. Colorimetric detection of mercury ion (Hg2+) based on DNA oligonucleotides and unmodified gold nanoparticles sensing system with a tunable detection range. Biosens. Bioelectron., 24, 3153-3158(2009).

    [46] J. Du, Z. Wang, J. Fan, X. Peng. Gold nanoparticle-based colorimetric detection of mercury ion via coordination chemistry. Sens. Actuators B, 212, 481-486(2015).

    [47] Y. Wang, M. Gao, C. Liao, F. Yu, L. Chen. A sulfydryl-based near-infrared ratiometic fluorescent probe for assessment of acute/chronic mercury exposure via associated determination of superoxide anion and mercury ion in cells and in vivo. Sens. Actuators B, 301, 127038(2019).

    [48] X. Guo, J. Huang, Y. Wei, Q. Zeng, L. Wang. Fast and selective detection of mercury ions in environmental water by paper-based fluorescent sensor using boronic acid functionalized MoS2 quantum dots. J. Hazard. Mater., 381, 120969(2020).

    [49] L. Wang, B. Li, F. Xu, X. Shi, D. Feng, D. Wei, Y. Li, Y. Feng, Y. Wang, D. Jia, Y. Zhou. High-yield synthesis of strong photoluminescent N-doped carbon nanodots derived from hydrosoluble chitosan for mercury ion sensing via smartphone APP. Biosens. Bioelectron., 79, 1-8(2016).

    [50] L. Zhang, T. Li, B. Li, J. Li, E. Wang. Carbon nanotube–DNA hybrid fluorescent sensor for sensitive and selective detection of mercury(II) ion. Chem. Commun., 46, 1476-1478(2010).

    [51] M. Fayazi, M. A. Taher, D. Afzali, A. Mostafavi. Fe3O4 and MnO2 assembled on halloysite nanotubes: a highly efficient solid-phase extractant for electrochemical detection of mercury(II) ions. Sens. Actuators B, 228, 1-9(2016).

    [52] Z. Zhu, Y. Su, J. Li, D. Li, J. Zhang, S. Song, Y. Zhao, G. Li, C. Fan. Highly sensitive electrochemical sensor for mercury(II) ions by using a mercury-specific oligonucleotide probe and gold nanoparticle-based amplification. Anal. Chem., 81, 7660-7666(2009).

    [53] S.-J. Liu, H.-G. Nie, J.-H. Jiang, G.-L. Shen, R.-Q. Yu. Electrochemical sensor for mercury(II) based on conformational switch mediated by interstrand cooperative coordination. Anal. Chem., 81, 5724-5730(2009).

    [54] Y. Zhu, L. Xu, W. Ma, Z. Xu, H. Kuang, L. Wang, C. Xu. A one-step homogeneous plasmonic circular dichroism detection of aqueous mercury ions using nucleic acid functionalized gold nanorods. Chem. Commun., 48, 11889-11891(2012).

    [55] J. Nan, X.-P. Yan. Facile fabrication of chiral hybrid organic–inorganic nanomaterial with large optical activity for selective and sensitive detection of trace Hg2+. Chem. Commun., 46, 4396-4398(2010).

    [56] S. Kacmaz, K. Ertekin, D. Mercan, O. Oter, E. Cetinkaya, E. Celik. An ultra sensitive fluorescent nanosensor for detection of ionic copper. Spectrochim. Acta A, 135, 551-559(2015).

    [57] X. Wang, Y. Lv, X. Hou. A potential visual fluorescence probe for ultratrace arsenic (III) detection by using glutathione-capped CdTe quantum dots. Talanta, 84, 382-386(2011).

    [58] T. Gong, J. Liu, X. Liu, J. Liu, J. Xiang, Y. Wu. A sensitive and selective sensing platform based on CdTe QDs in the presence of L-cysteine for detection of silver, mercury and copper ions in water and various drinks. Food Chem., 213, 306-312(2016).

    Xiongbin Wang, Qiushi Wang, Yulong Chen, Jiagen Li, Ruikun Pan, Xing Cheng, Kar Wei Ng, Xi Zhu, Tingchao He, Jiaji Cheng, Zikang Tang, Rui Chen. Metal-to-ligand charge transfer chirality-based sensing of mercury ions[J]. Photonics Research, 2021, 9(2): 213
    Download Citation