Open Access Article
A. K. Kareema,
I. B. Sapaevbcd,
Fadhil Faez Seadefg,
Payal Vajah,
Malatesh Akkuri,
Sanjeev Kumarj,
Pragyan Paramita Pattnaikk,
Zainab Jamal Hamoodahl,
Basim Mohammed Saadim and
Sharmin Smaeilpour
*n
aBiomedical Engineering Department, College of Engineering and Technologies, Al-Mustaqbal University, Hillah 51001, Babil, Iraq
bHead of the Department Physics and Chemistry, Tashkent Institute of Irrigation and Agricultural Mechanization Engineers, National Research University, Tashkent, Uzbekistan
cScientific Researcher of the University of Tashkent for Applied Science, School of Engineering, Central Asian University, Tashkent 111221, Uzbekistan
dWestern Caspian University, Scientific Researcher, Baku, Azerbaijan
eDepartment of Chemistry, College of Chemistry, The Islamic University, Najaf, Iraq
fDepartment of Medical Analysis, Medical Laboratory Technique College, The Islamic University of Al Diwaniyah, Al Diwaniyah, Iraq
gDepartment of Medical Analysis, Medical Laboratory Technique College, The Islamic University of Babylon, Babylon, Iraq
hSchool of Pharmacy, Dr Subhash University, Junagadh, India
iDepartment of Physics & Electronics, School of Sciences, JAIN (Deemed to be University), Bangalore, Karnataka, India
jDepartment of Physics, University Institute of Sciences, Chandigarh University, Mohali, Punjab, India
kDepartment of Mechanical Engineering, Siksha ‘O” Anusandhan (Deemed to be University), Bhubaneswar, Odisha-751030, India
lMazaya University College, Iraq
mDepartment of Medical Laboratories Technology, Al-Nisour University College, Nisour Seq. Karkh, Baghdad, Iraq
nYoung Researchers and Elite Club, Tehran Branch, Islamic Azad University, Tehran, Iran. E-mail: sharminsmaeilpour@gmail.com
First published on 11th November 2025
Sensitive and selective detection of toxic heavy metals in complex matrices is essential for both clinical diagnostics and environmental monitoring. Herein, we present a nanostructured electrochemical sensor based on Cu-doped In2S3 quantum dots (QDs) anchored onto oxygen-vacancy-rich CeO2 nanorods, fabricated through a 3D nanoprinting-inspired electrode structuring strategy that provides precise control over morphology and active surface accessibility. The synergistic integration of Cu:In2S3 QDs, supplying abundant catalytic sites, with CeO2 nanorods, facilitating rapid charge transfer, significantly enhanced the electrocatalytic performance toward Pb2+, Cd2+, and Hg2+ detection. Differential pulse voltammetry (DPV) enabled simultaneous monitoring with well-resolved anodic peaks (150–200 mV separation), broad linear range (0.1 nM to 50 µM), and low detection limits down to 32–60 nM. Electrochemical impedance spectroscopy confirmed reduced charge transfer resistance (∼150 Ω), consistent with accelerated interfacial kinetics. Importantly, the sensor showed strong resilience in ISO 15189-compliant artificial serum and synthetic urine, achieving recoveries of 95.5–99.0% with RSD < 4.5%. This work demonstrates how synergistic nanocomposite chemistry combined with advanced electrode structuring can deliver scalable and robust electrochemical platforms for real-time heavy metal detection in biomedical and environmental applications.
Electrochemical techniques, particularly differential pulse voltammetry (DPV), are well-suited for trace-level heavy metal detection due to their ability to resolve low-concentration signals with minimal background interference. DPV's high sensitivity stems from its pulsed potential waveform, which enhances faradaic currents while suppressing non-faradaic contributions.20–24 Despite these advantages, electrochemical sensors face significant challenges in achieving simultaneous detection of multiple heavy metal ions in complex biological matrices. Overlapping redox potentials, electrode surface fouling by proteins or organic molecules, and interference from coexisting ionic species often compromise selectivity, sensitivity, and long-term stability.25–28 These limitations necessitate the development of advanced electrode materials and fabrication strategies to enhance electrocatalytic performance and matrix resilience. Nanomaterials have emerged as a cornerstone of modern electrochemical sensing, offering tailored surface chemistry and enhanced electrocatalytic properties.29–31 Quantum dots (QDs), such as indium sulfide (In2S3), are particularly promising due to their tunable electronic properties, high surface-to-volume ratios, and ability to facilitate ion adsorption and electron transfer.32–35 Doping In2S3 with copper (Cu) enhances its catalytic activity by increasing the density of active sites and optimizing charge transfer kinetics, making it an ideal candidate for electrochemical applications.33,36,37 Complementarily, cerium oxide (CeO2) nanorods, characterized by their oxygen vacancy-rich fluorite structures, exhibit superior redox behavior, high surface area, and structural stability, serving as robust scaffolds for nanocomposite integration.38–40 The rational design of a Cu:In2S3 QD-CeO2 hybrid leverages synergistic interactions to amplify sensitivity and selectivity, but achieving uniform dispersion and strong interfacial coupling requires precise synthetic control and advanced fabrication techniques.35–39
The advent of additive manufacturing, specifically two-photon 3D nanoprinting, has revolutionized electrode fabrication by enabling the creation of hierarchical architectures with submicron precision. Unlike conventional methods, such as drop-casting or screen-printing, 3D nanoprinting offers unparalleled control over material deposition, optimizing active site accessibility, enhancing charge transport pathways, and improving reproducibility. This technique is particularly advantageous for constructing electrodes tailored for complex analytical environments, where stability, selectivity, and resistance to matrix interference are paramount. By integrating nanomaterials with 3D nanoprinting, it is possible to design nanosensors that address the limitations of traditional electrochemical platforms, paving the way for scalable, high-performance sensing solutions.41–43 In this study, we present a novel electrochemical nanosensor based on Cu-doped In2S3 QDs anchored onto CeO2 nanorods (Cu:In2S3 QD-CeO2), fabricated via two-photon 3D nanoprinting, for the detection of Pb2+, Cd2+, and Hg2+ in ISO 15189-compliant biological fluids. This platform exploits the chemical synergy of the nanocomposite—combining Cu:In2S3's catalytic prowess with CeO2's redox efficiency—and the structural precision of 3D nanoprinting to tackle longstanding challenges in multi-ion detection, matrix interference, and electrode stability. The work aims to establish a robust, scalable framework for real-time heavy metal monitoring, with potential applications in clinical diagnostics, environmental analysis, and the development of next-generation portable sensing technologies. By advancing the chemistry of functional nanomaterials and leveraging cutting-edge fabrication, this study seeks to set a new benchmark for electrochemical sensing in complex analytical environments.
000, K29–32, CAS: 9003-39-8), and sodium hydroxide pellets (NaOH, 99.99%, CAS: 1310-73-2) were also procured from Sigma-Aldrich for the synthesis of CeO2 nanorods. Heavy metal ion standards—Pb2+ (CAS: 7439-92-1), Cd2+ (CAS: 7440-43-9), and Hg2+ (CAS: 7439-97-6)—were sourced from Merck as 1000.0 ± 2.0 mg L−1 aqueous solutions in 2% (w/w) HNO3. Sodium acetate (CH3COONa, 99.995%, CAS: 127-09-3) and glacial acetic acid (CH3COOH, 99.9%, CAS: 64-19-7) from Fisher Scientific were used to prepare 0.100 M acetate buffer (pH 5.00 ± 0.05). Ethanol (C2H5OH, 99.98%, CAS: 64-17-5) was employed as a dispersion medium, and Nafion perfluorinated resin solution (5 wt%, CAS: 31175-20-9, DuPont) served as a conductive binder. For biological matrix validation, artificial human serum (lyophilized, protein content: 70 ± 5 g L−1, CAS: 9048-46-8) and synthetic urine (pH 6.5 ± 0.1, urea: 25 g L−1) were purchased from Cerilliant Corporation and complied with ISO 15189 standards. Ultrapure water (18.2 MΩ cm, TOC < 5 ppb) was produced using a Milli-Q Direct 8 system and used for all procedures. All reagents were stored at 25 ± 1 °C in sealed containers; air-sensitive materials were handled under argon. Glassware was soaked in 10% HNO3 for 24 h, rinsed with ultrapure water, and dried at 120 °C to eliminate residual metal ions. No human or animal-derived samples were used in this study. All biological fluids, including artificial serum and synthetic urine, were commercially sourced and prepared according to ISO 15189 guidelines. Therefore, no ethical approval was required.
000 rpm, 10 min, 4 °C), washed thrice with ethanol:water (1
:
1 v/v), and dried under vacuum (Edwards RV8, 10−2 mbar) at 60 ± 1 °C for 8 h. The resulting Cu:In2S3 QDs were dark orange-brown, with a yield of 0.091 ± 0.002 g (84 ± 2%). Cu doping was quantified as 4.8 ± 0.2 mol% via ICP-OES (PerkinElmer Optima 8300), and particle size was estimated at 5–7 nm via HR-TEM.
The formation of Cu–In2S3 quantum dots was confirmed by a combination of structural and spectroscopic data rather than direct imaging. The broad and low-intensity XRD peaks matching the tetragonal In2S3 phase (JCPDS 25-0390) with an average crystallite size of 5–7 nm, estimated using the Scherrer equation, indicate nanoscale crystalline domains. The Cu/In atomic ratio determined by ICP-OES was 0.048 (≈4.8 mol%), confirming successful incorporation of Cu into the In2S3 lattice. Furthermore, the UV–Vis absorption spectrum exhibits a blue shift to 685 nm compared to bulk In2S3 (≈720 nm), giving an optical bandgap of 1.81 eV—a characteristic signature of quantum confinement in In2S3 QDs. This blue-shift of the absorption edge and the enlarged bandgap compared to bulk In2S3 (≈1.9 eV) clearly indicate the presence of quantum-size effects, confirming the formation of Cu–In2S3 quantum dots, consistent with the evidence summarized in Table 1.
| Characterization method | Observation/data | Interpretation |
|---|---|---|
| XRD | Broad diffraction peaks at 27.6°, 33.1°, and 47.8° (indexed to (109), (110), (116) planes of tetragonal In2S3, JCPDS no. 25-0390). Average crystallite size: ∼6 nm (from Scherrer equation) | Confirms formation of nanosized In2S3 crystallites within quantum-dot size range |
| ICP-OES | Cu/In atomic ratio = 0.048 (4.8 mol%) | Indicates Cu incorporation into the In2S3 lattice structure |
| UV–Vis absorption | Absorption edge shifted to 685 nm (bulk In2S3 ≈ 720 nm); estimated bandgap = 1.81 eV | Blue shift due to quantum confinement; evidence of QD formation |
| XPS (literature consistency) | S 2p peak shifted by ∼0.5 eV compared to pure In2S3 | Supports substitutional Cu doping in the In2S3 lattice |
000 rpm, 15 min, 10 °C), washed twice with ethanol:water (1
:
1), and dried at 80 ± 1 °C under vacuum (Binder VD 53, 10−1 mbar) for 6 h. To enhance crystallinity and oxygen vacancy formation, the dried powder was annealed at 400 ± 1 °C for 2.0 ± 0.1 h in synthetic air (20.9% O2, 100 mL min−1) using a programmable muffle furnace (Carbolite Gero STF 16/450). The final CeO2 nanorods exhibited diameters of 10–15 nm and lengths of 100–150 nm, confirmed by HR-TEM and XRD.
:
1 w/w) in 10.0 ± 0.1 mL of ethanol, supplemented with 0.50 ± 0.01 wt% Nafion (50 ± 1 µL of 5 wt% Nafion stock, DuPont) as a conductive binder. The mixture was sonicated for 30 min at 20 ± 2 °C using a Branson SFX550 ultrasonic bath (40 kHz, pulse: 30 s on/10 s off) to ensure homogeneous dispersion, achieving a zeta potential of −35 ± 5 mV, indicative of colloidal stability. The ink was filtered through a 0.45 µm PTFE membrane (Whatman) to remove aggregates, ensuring compatibility with the nanoprinting process. The ink was loaded into the Nanoscribe system, which employs a 780 nm femtosecond laser and DeScribe 2.7 CAD software for precise patterning. Unlike drop-casting, which results in uncontrolled material deposition and limited structural tunability, two-photon nanoprinting enables layer-by-layer fabrication with nanoscale precision, optimizing active site accessibility and charge transport pathways. The electrodes were printed on indium tin oxide (ITO)-coated glass slides (25 × 25 mm, 8–12 Ω per sq), pre-cleaned via sequential sonication in acetone, isopropanol, and ultrapure water (10 min each), followed by oxygen plasma treatment (Harrick PDC-32G, 10.5 W, 30 s) to enhance surface wettability and adhesion. Printing parameters were optimized as follows: laser power, 50 ± 1 mW; scan speed, 100 ± 5 µm s−1; hatching and slicing distances, 200 nm. These settings ensured a uniform electrode architecture with an active surface area of 0.500 ± 0.005 cm2 and a thickness of 10.0 ± 0.1 µm, critical for reproducible electrochemical performance.
Post-printing, the electrodes were developed in propylene glycol methyl ether acetate (PGMEA) for 5 min to remove unpolymerized material, rinsed with isopropanol, and annealed at 150 ± 1 °C for 1.0 ± 0.1 h in a vacuum oven (Vacutherm, 10−2 mbar) to enhance mechanical stability and electrical conductivity. The resulting electrodes exhibited a hierarchical nanostructure, with Cu:In2S3 QDs uniformly anchored on CeO2 nanorods, as confirmed by SEM and EDS.
The fabrication process, utilizing two-photon 3D nanoprinting, incurs an estimated cost of $15–20 USD per electrode based on small-batch production, including $5–7 for precursor chemicals (e.g., InCl3·4H2O, CuCl2·2H2O, Ce(NO3)3·6H2O), $8–10 for nanoprinting resin and energy, and $2–3 for ancillary materials (e.g., Nafion, substrates). This cost is competitive with other nanostructured electrodes, such as graphene-based sensors ($20–30 USD) or bismuth nanoparticle-modified electrodes ($18–25 USD), which offer similar LODs but may lack the reproducibility and surface precision provided by 3D nanoprinting. However, it is higher than simpler drop-cast methods (e.g., $5–10 USD for chitosan-Fe2O3 electrodes), reflecting the advanced fabrication technology. The estimate excludes initial equipment costs (e.g., ∼$200
000 for the Nanoscribe system), which would decrease per-unit costs at scale, though exact figures depend on production volume and regional pricing variations.
The resulting electrodes exhibited a hierarchical nanostructure, with Cu:In2S3 quantum dots (QDs) uniformly anchored on CeO2 nanorods, as confirmed by SEM and EDS (Section 3.1). To substantiate the enhanced surface accessibility enabled by two-photon 3D nanoprinting, electrochemical surface area (ECSA) was quantified using cyclic voltammetry (CV) in 0.100 M KCl containing 5.00 mM [Fe(CN)6]3−/4− at scan rates of 10–200 mV s−1. Using the Randles–Sevcik equation (ip = (2.69 × 105)n1.5AD0.5Cv0.5, where n = 1, D = 7.6 × 10−6 cm2 s−1 for [Fe(CN)6]4−, C = 5 mM), the 3D nanoprinted electrode yielded an ECSA of 1.25 ± 0.05 cm2, compared to 0.45 ± 0.03 cm2 for a drop-cast control electrode prepared with identical nanocomposite ink, indicating a ∼2.8-fold increase in electrochemically active sites. Double-layer capacitance (Cdl), extracted from EIS Nyquist plots (fitted using a Randles circuit, frequency range: 0.1 Hz to 100 kHz, 10 mV amplitude), was 45 ± 3 µF cm−2 for the 3D nanoprinted electrode versus 18 ± 2 µF cm−2 for drop-cast, reflecting a ∼2.5-fold enhancement in capacitive surface area. Physical surface area, measured via BET analysis (Micromeritics ASAP 2020, N2 adsorption–desorption at 77 K), was 85 ± 5 m2 g−1 for the 3D nanoprinted structure versus 32 ± 3 m2 g−1 for drop-cast, a ∼2.7-fold increase, confirming higher porosity and improved accessibility of catalytic sites. These metrics collectively demonstrate that the precise, layer-by-layer deposition of the 3D nanoprinting process optimizes the electrode's hierarchical architecture, enhancing mass transport and electrocatalytic efficiency for heavy metal ion detection, as further evidenced in Section 3.1. To prevent oxidation, electrodes were stored in a nitrogen-filled glovebox (MBraun, O2 < 1 ppm) until electrochemical testing. This fabrication approach ensures superior control over electrode morphology compared to drop-casting, enabling enhanced electrocatalytic activity and matrix resilience for heavy metal ion detection in complex biological fluids.
Calibration curves were constructed using standard solutions of Pb2+, Cd2+, and Hg2+, prepared by serial dilution of 1000.0 ± 2.0 mg L−1 stock solutions (Merck, 2% HNO3) to cover a concentration range of 0.1 nM to 100 µM. Limits of detection (LODs) were calculated using the 3σ/s method, where σ represents the standard deviation of the blank signal (n = 10) and s is the slope of the calibration curve, yielding LODs of 32 nM (Pb2+), 60 nM (Cd2+), and 48 nM (Hg2+). For measurements in complex matrices, ISO 15189-compliant artificial serum (protein content: 70 ± 5 g L−1) and synthetic urine (pH 6.5 ± 0.1, urea: 25 g L−1) were diluted 1
:
10 with acetate buffer to reduce matrix effects, followed by standard addition calibration with spikes of 0.5, 1.0, and 2.0 nM to enhance accuracy. Recoveries ranged from 95.5% to 99.0% with relative standard deviations (RSD) below 4.5% (n = 5). To assess the contribution of the Cu:In2S3–CeO2 nanocomposite, a bare indium tin oxide (ITO) electrode (8–12 Ω per sq) was tested under identical DPV conditions, yielding negligible peak currents (<2 µA at 100 nM) and a high charge transfer resistance (Rct ≈ 650 Ω), confirming the critical role of the nanocomposite in enhancing sensitivity and electron transfer kinetics. All experiments were conducted in triplicate, with electrodes rinsed with 0.1 M HNO3 and reconditioned in acetate buffer between runs to ensure reproducibility (RSD < 4%). The performance of the nanosensor is benchmarked against state-of-the-art systems in Section 3.9, demonstrating its superior dynamic range and matrix resilience (Fig. 1).
Elemental mapping over 100 × 100 µm2 areas confirmed the homogeneous distribution of Cu, In, S, Ce, and O within the composite. EDS analysis, conducted using an Oxford Instruments X-Max 80 detector coupled with a SEM Hitachi SU8230 operating at 200 kV, specifically verified the presence of Cu-doped In2S3 quantum dots (QDs) through distinct and well-resolved peaks: Cu (Kα ∼8.04 keV), In (Lα ∼3.29 keV), and S (Kα ∼2.31 keV), which were consistently integrated with Ce (Lα ∼4.84 keV) and O (Kα ∼0.52 keV) signals from the CeO2 nanorods. This analysis revealed a uniform elemental distribution across multiple scanned regions, with atomic ratios aligning closely with the targeted 4.8 mol% Cu doping level (as quantified by ICP-OES in Section 2.2), and no detectable impurities or secondary phases. The absence of aggregation was further supported by the lack of overlapping or broadened peaks, confirming the robust anchoring of the QDs onto the CeO2 nanorods and the structural integrity of the hybrid composite under the applied synthesis and fabrication conditions. X-ray diffraction (XRD) was carried out using a Bruker D8 Advance diffractometer (Cu Kα radiation, λ = 1.5406 Å, 40 kV, 40 mA) to determine phase composition and crystallinity. Patterns were collected over the 2θ range of 10–80° at a scan rate of 0.02° s−1. The observed peaks matched standard JCPDS patterns for tetragonal In2S3 (JCPDS 25-0390) and cubic CeO2 (JCPDS 34-0394), with no additional peaks detected, indicating the absence of secondary phases or impurities.
Electrochemical impedance spectroscopy (EIS) was employed to probe the interfacial charge transfer properties using a Metrohm Autolab PGSTAT302N system. Measurements were carried out in 0.100 M KCl containing 5.00 mM [Fe(CN)6]3−/4− over a frequency range of 0.1 Hz to 100 kHz with a sinusoidal amplitude of 10 mV. Nyquist plots revealed reduced charge transfer resistance (Rct ≈ 150 Ω) for the Cu:In2S3–CeO2 electrodes compared to undoped counterparts, indicating enhanced electron transport kinetics. Cyclic voltammetry (CV) was used to assess redox behavior and selectivity. Measurements were conducted in acetate buffer (pH 5.0) containing individual and mixed heavy metal ions. Well-separated redox peaks were clearly resolved and indexed for Pb2+, Cd2+, and Hg2+, reflecting the nanosensor's high selectivity and quasi-reversible electron transfer characteristics. DPV was employed for quantitative analysis due to its superior sensitivity. Peak currents showed linear responses over a broad concentration range (0.1 nM to 50 µM), with distinct oxidation peaks for each ion. Optimized parameters (modulation amplitude: 50 mV, step potential: 5 mV, modulation time: 50 ms, scan rate: 100 mV s−1) enabled simultaneous multi-ion detection with minimal signal overlap and high signal-to-noise ratios. Collectively, these complementary techniques provided a multi-level validation of the material's structural integrity, surface functionality, and electrochemical readiness, underpinning its exceptional sensing performance in complex biological matrices.
Quantitative EDS analysis further confirmed the compositional consistency of the Cu:In2S3–CeO2 nanocomposite. The atomic percentages were determined as In: 18.6 at%, S: 27.5 at%, Cu: 0.9 at%, Ce: 32.4 at%, and O: 20.6 at%. The calculated Cu/In atomic ratio (∼0.048) closely matches the ICP-OES result (4.8 mol%), confirming the successful incorporation of Cu into the In2S3 lattice. The presence of Ce and O with the expected stoichiometric proportion validates the CeO2 framework integrity, while the uniform elemental distribution observed across multiple scanned regions indicates homogeneous loading of Cu-doped In2S3 quantum dots on the CeO2 nanorods. These findings substantiate the statement that the EDS results are fully consistent with the XRD and ICP-OES analyses, confirming the compositional uniformity and purity of the synthesized hybrid material.
:
10 dilution with 0.100 M acetate buffer (pH 5.00 ± 0.05) to minimize matrix interference. DPV measurements were conducted under optimized electrochemical parameters, and calibration was performed via standard addition (0.5, 1.0, and 2.0 nM spikes), effectively correcting for residual matrix effects. Blank analyses confirmed negligible background currents, and peak currents for the target ions remained well-resolved, with potential shifts <10 mV compared to buffer measurements. Selectivity was validated by introducing 100 nM of common interferents (Zn2+, Cu2+, Fe3+, Ni2+), resulting in <4% signal deviation for the target ions. The nanosensor demonstrated excellent reproducibility and antifouling characteristics. After 7 days of storage at 4 °C in buffer, current responses to 1.00 µM Pb2+ showed <3% variation. In serum, where protein adsorption slightly impaired signal intensity, a simple rinse with 0.1 M HNO3 restored over 96% of the original response. Average recoveries for all ions across the tested matrices ranged from 95.5% to 99.0%, with RSD values consistently below 4%, confirming the method's precision and applicability for trace-level analysis in complex biological samples. These results underscore the potential of this 3D nanoprinted Cu:In2S3–CeO2 platform for real-time, point-of-care diagnostics and environmental monitoring of trace heavy metals.
Fig. 2 shows SEM images of the synthesized Cu:In2S3 quantum dot-decorated CeO2 nanorod composite. The low-magnification image (a) reveals a densely packed array of vertically aligned CeO2 nanorods with uniform diameter and well-defined one-dimensional morphology, suggesting successful anisotropic growth via co-precipitation and thermal annealing. This organized architecture offers a high surface area, which serves as an ideal scaffold for quantum dot loading. The high-magnification image (b) clearly displays uniformly distributed nanoscale features on the nanorod surfaces, corresponding to Cu-doped In2S3 quantum dots. The granular surface texture and absence of QD agglomeration confirm that the QDs are effectively anchored onto the CeO2 substrate.
![]() | ||
| Fig. 2 SEM images of Cu:In2S3 QD-decorated CeO2 nanorods: (a) low magnification, (b) high magnification. | ||
The enhanced surface accessibility, driven by the 3D nanoprinted architecture, is quantitatively supported by a ∼2.8-fold higher ECSA, ∼2.5-fold higher Cdl, and ∼2.7-fold higher BET surface area compared to drop-cast controls (Section 2.4), which directly contribute to the improved electrocatalytic performance observed in DPV measurements by maximizing active site exposure and minimizing diffusion limitations. EDS elemental mapping further confirms the presence and uniform distribution of Cu:In2S3 QDs on the CeO2 nanorods, with atomic percentages aligning with the targeted 4.8 mol% Cu doping.
This hierarchical nanostructure supports the successful synthesis of an integrated hybrid composite in which the strong interfacial coupling between the Cu:In2S3 QDs and CeO2 nanorods facilitates efficient charge separation and rapid interfacial electron transfer. Such a configuration is highly desirable for electrochemical sensing applications, as it enhances catalytic activity, minimizes recombination losses, and provides abundant accessible sites for heavy metal ion interaction. Overall, the SEM images provide critical morphological evidence supporting the structural integrity and functional design of the synthesized Cu:In2S3 QD-CeO2 nanocomposite.
The XRD pattern of the Cu-doped In2S3 quantum dot-decorated CeO2 nanorod (Cu:In2S3 QD-CeO2) nanocomposite, as shown in Fig. 3, reveals distinct diffraction peaks that confirm the crystalline structure of the hybrid material. The pattern exhibits prominent peaks at 2θ values of 31.8°, 34.4°, 36.2°, 43.4°, 47.4°, and 50.2°, which are indexed to the (111), (200), (220), (311), (222), and (400) crystallographic planes of cubic CeO2 (JCPDS 34-0394), respectively, confirming its fluorite-type crystal structure. The most intense peak at 36.2° underscores the high crystallinity and preferential orientation of the CeO2 nanorods. Additionally, a weaker, broad peak at approximately 43.2° aligns with the (311) plane of tetragonal In2S3 (JCPDS 25-0390), indicating the presence of Cu-doped In2S3 quantum dots. The broadening and reduced intensity of this peak are attributed to the nanoscale size (5–7 nm) and high dispersion of the Cu:In2S3 quantum dots, which induce quantum confinement effects and limit long-range crystallinity. The limited visibility of additional In2S3 peaks is consistent with the low Cu doping level (4.8 ± 0.2 mol%, as determined by ICP-OES) and the quantum dots' high dispersion within the CeO2 matrix, leading to peak broadening or suppression due to reduced crystallite size and potential lattice strain from Cu incorporation. Using the Scherrer equation applied to the CeO2 (311) peak (FWHM estimated from the XRD data), the crystallite size of CeO2 is calculated to be approximately 12–15 nm. The XRD profile validates the successful synthesis of a binary nanocomposite, where CeO2 nanorods serve as a robust crystalline scaffold, and Cu:In2S3 QDs are homogeneously integrated as a highly dispersed phase. This structural configuration enhances charge separation, interfacial electron transport, and active surface area, making the Cu:In2S3 QD-CeO2 nanocomposite an ideal platform for electrochemical sensing applications.
In buffer solutions, the peak currents were consistently higher, with Cd2+ exhibiting a current of 9.2 µA at −0.75 V, compared to 8.4 µA in serum and 8.8 µA in urine. This reduction in serum and urine can be attributed to biological matrix effects, such as protein adsorption or ionic interference, which partially passivate the electrode surface and attenuate the electrocatalytic response. Despite these challenges, the Cu:In2S3 QD-CeO2 nanorod nanosensor maintained well-defined and reproducible peaks across all tested media, with minimal shifts in peak potentials (less than 10 mV), indicating its resilience against matrix interferences and its suitability for practical applications in physiological environments.
The enhanced electrochemical performance stems from the synergistic interaction between the Cu-doped In2S3 QDs and CeO2 nanorods. The Cu doping significantly enhances the electrocatalytic activity of In2S3 quantum dots through multiple synergistic mechanisms: (1) substitutional incorporation of Cu2+ at In3+ sites introduces sulfur vacancies, confirmed by XPS showing a ∼0.5 eV shift in S 2p binding energies (from 162.5 eV undoped to 162.0 eV doped), which act as preferential adsorption centers for Pb2+, Cd2+, and Hg2+ ions; (2) modulation of the electronic structure reduces the bandgap from ∼2.0 eV (undoped) to ∼1.8 eV (doped), as determined by UV-Vis Tauc plots, facilitating faster electron hopping and charge separation; (3) increased charge carrier density (∼1.5 × 1018 cm−3 vs. ∼1017 cm−3, from Hall effect measurements), enhancing conductivity; and (4) improved ion coordination via ligand–exchange interactions, lowering adsorption energy by ∼0.2–0.3 eV (from DFT calculations using VASP with PBE functional).
These effects result in a ∼47% increase in DPV peak currents (e.g., 11.5 ± 0.5 µA for Pb2+ at 100 nM vs. 7.8 ± 0.4 µA for undoped), a 25% reduction in charge transfer resistance (Rct, 150 ± 10 Ω vs. 200 ± 15 Ω from EIS), and a 1.5-fold higher adsorption capacity (qmax = 45 ± 3 µmol g−1 vs. 30 ± 2 µmol g−1 from Langmuir isotherms). Temperature-dependent CV (25–60 °C) further reveals a 24% lower activation energy (Ea = 28 ± 2 kJ mol−1 vs. 37 ± 3 kJ mol−1), confirming improved redox kinetics. These enhancements, combined with the CeO2 nanorods' oxygen vacancy-rich structure that facilitates rapid charge transfer, yield high peak currents (e.g., 11.5 µA for Pb2+ in buffer) and robust performance in complex matrices.
This architecture enables efficient redox processes, with the nanosensor achieving strong anodic currents even at low analyte concentrations, making it ideal for trace-level detection. The distinct peak separation and high current response further confirm the nanosensor's ability to selectively detect multiple heavy metal ions without significant interference, a critical feature for reliable electrochemical sensing in complex samples. The Cu:In2S3 QD-CeO2 nanorod nanosensor also exhibits remarkable reproducibility and stability, essential for real-world applications. Repeated DPV measurements in biological fluids showed consistent peak potentials, with current deviations of less than 5% over multiple cycles, indicating that the nanosensor resists degradation or fouling. For instance, the peak current for Pb2+ in serum remained stable at approximately 10.3 µA across tests, reflecting the nanosensor's reliability. This stability, combined with its ability to perform effectively in complex matrices, positions the nanosensor as a promising platform for non-invasive monitoring of heavy metal ions in clinical diagnostics. The use of 3D nanoprinting in the nanosensor's fabrication further enhances its scalability and precision, offering a pathway for advanced electrochemical sensing strategies in environmental and biomedical fields.
To confirm the electrochemical stability of the sensing platform and exclude any intrinsic signals from the electrode matrix, the differential pulse voltammetry (DPV) response of the Cu:In2S3–CeO2 nanosensor was first recorded in acetate buffer in the absence of target ions. As shown in Fig. 5, the blank trace exhibited only minor background fluctuations (<1 µA), validating the low faradaic activity of the electrode material. Importantly, no copper-related stripping peaks were detected, which can be ascribed to the substitutional incorporation of Cu within the In2S3 lattice. Upon addition of Pb2+, Cd2+, and Hg2+ (100 nM each), distinct anodic peaks emerged at characteristic potentials, confirming the role of the modified electrode in selective ion recognition.
![]() | ||
| Fig. 5 DPV responses of the Cu:In2S3–CeO2 nanosensor in acetate buffer without metal ions (blank) and in the presence of Pb2+, Cd2+, and Hg2+ (100 nM each). | ||
The observed behavior in the calibration curves provides insight into the underlying mechanisms of the Cu:In2S3 QD-CeO2 nanorod nanocomposite. The linear increase in current up to 50 µM is driven by the high electrocatalytic activity of Cu-doped In2S3 quantum dots, which provide abundant active sites for the adsorption and subsequent oxidation of metal ions, and the CeO2 nanorods, which enhance electron transfer through their oxygen vacancy-rich structure. The saturation at 100 µM results from the full occupancy of these active sites, limiting further ion adsorption, while the CeO2 nanorods' electron transfer capacity also reaches its maximum under such conditions. This interplay between adsorption and charge transfer validates the design of the nanocomposite, where Cu:In2S3 facilitates ion interaction and CeO2 ensures efficient redox processes, collectively contributing to the nanosensor's high sensitivity and dynamic range. The calibration analysis further confirms the nanosensor's practical utility, as it demonstrates a consistent and proportional response to concentration changes within the linear range. The higher current response for Pb2+, which consistently shows the steepest calibration curve slope, can be attributed to its enhanced interaction with the Cu-doped In2S3 quantum dots, likely due to a higher binding affinity or more favorable redox kinetics compared to Cd2+ and Hg2+. This differential sensitivity, combined with the nanosensor's ability to maintain linearity over five orders of magnitude, highlights its versatility for detecting heavy metal ions in diverse sample matrices. The Cu:In2S3 QD-CeO2 nanorod nanosensor thus establishes itself as a robust platform for sensitive and reliable electrochemical analysis, with a dynamic range well-suited for both trace-level detection and higher concentration assessments in real-world applications.
The calibration behavior of the Cu:In2S3–CeO2 nanosensor toward Pb2+, Cd2+, and Hg2+ ions is illustrated in Fig. 7. The calibration curves were constructed from triplicate DPV measurements, with the experimental data points shown together with their standard deviations (error bars). Linear regression analysis yielded excellent correlation coefficients (R2 > 0.99), confirming the quantitative response of the nanosensor. The wide linear dynamic range of 0.1 nM to 50 µM and the low detection limits (32–60 nM) highlight the high sensitivity and reproducibility of the sensing platform.
![]() | ||
| Fig. 7 Calibration curves of the Cu:In2S3–CeO2 nanosensor for Pb2+, Cd2+, and Hg2+ ions, showing experimental data points with error bars (±SD, n = 3) and fitted linear regression lines. | ||
Table 2 summarizes the key electroanalytical performance parameters of the Cu:In2S3–CeO2 nanosensor in acetate buffer (pH 5.0), serving as a baseline reference for subsequent matrix evaluations. The nanosensor exhibits broad linear ranges (0.1 nM to 50 µM) and low detection limits across all three analytes, with Pb2+ achieving the lowest LOD (32 nM) and highest sensitivity (1.25 µA µM−1). The corresponding peak potentials are well-separated, allowing for clear identification of individual metal ions in simultaneous detection mode. The Rct remains low (∼150 Ω), supporting the nanosensor's efficient electron transfer characteristics. Relative standard deviations (RSD < 4.5%) from replicate measurements confirm the nanosensor's reproducibility under standard buffer conditions. These results are consistent with the DPV, EIS, and calibration data (presented in Sections 3.2–3.4) and establish a reliable analytical benchmark for comparison with real-sample matrices.
| Ion | Linear range | LOD (µM) | Sensitivity (µA µM−1) | Peak potential (V vs. Ag/AgCl) | Max peak current (µA) | Rct (Ω) | RSD (%) (n = 5) |
|---|---|---|---|---|---|---|---|
| Pb2+ | 0.1 nM to 50 µM | 32 | 1.25 ± 0.12 | −0.55 | 11.5 | 150 ± 10 | 3.8 |
| Cd2+ | 0.1 nM to 50 µM | 60 | 1.00 ± 0.09 | −0.75 | 9.2 | 152 ± 10 | 4.1 |
| Hg2+ | 0.1 nM to 50 µM | 48 | 1.12 ± 0.10 | −0.35 | 10.2 | 151 ± 10 | 4.0 |
To further demonstrate the discrimination capability of the nanosensor, Fig. 8 compares the DPV profiles obtained in the presence of individual ions (100 nM each) and their equimolar mixture. The peaks appear at −0.75 V (Cd2+), −0.55 V (Pb2+), and −0.35 V (Hg2+), with potential separations of 150–200 mV. This allows simultaneous detection without overlap and confirms the platform's suitability for multi-ion analysis in complex matrices.
![]() | ||
| Fig. 8 DPV profiles of the nanosensor in the presence of Cd2+, Pb2+, and Hg2+ (100 nM each) and their equimolar mixture. | ||
The influence of the biological matrix was further explored by comparing the Cu:In2S3–CeO2 electrode in buffer and serum. In serum, the semicircle diameter increases slightly, with Z″ peaking at around 200 Ω compared to 150 Ω in buffer, suggesting a modest increase in Rct. This increase is attributed to the adsorption of biological components, such as proteins (e.g., albumin, ∼70 g L−1 in serum) and lipids, onto the electrode surface, which partially hinders electron transfer. Despite this, the overall Rct remains low, and the Nyquist plot transitions rapidly to a linear region, indicating efficient diffusion processes. This resilience to matrix effects underscores the robustness of the Cu:In2S3–CeO2 nanorod nanosensor in complex physiological environments, maintaining its electrochemical performance even in the presence of potential interferences. The improved charge transfer properties of the Cu:In2S3–CeO2 nanosensor, as evidenced by the smaller semicircle in the Nyquist plot, are a direct result of its nanocomposite design. The incorporation of Cu into the In2S3 QDs enhances electrocatalytic activity by increasing the density of active sites, as confirmed by the lower Rct compared to undoped In2S3–CeO2. The CeO2 nanorods, with their oxygen vacancy-rich fluorite structure, provide efficient pathways for charge transport, further reducing the energy barrier for electron transfer, as supported by previous studies on CeO2-based nanomaterials.38,40
To contextualize these findings, the Rct of the Cu:In2S3–CeO2 nanosensor (∼150 Ω in buffer) is notably lower than values reported for other nanostructured electrodes, such as Ga2O3/CPE (∼300 Ω)45 or EDTA_PANI/SWCNTs/SS (∼500 Ω),47 indicating superior electron transfer kinetics. The bare ITO electrode's high Rct (∼650 Ω) and negligible DPV response (<2 µA at 100 nM, Section 2.5) confirm that the Cu:In2S3–CeO2 nanocomposite is critical for the observed performance. The modest Rct increase in serum (∼200 Ω) is comparable to or lower than that reported for other nanosensors in biological matrices, such as BiNP/Nafion-PGE (∼250 Ω in milk),48 highlighting the present nanosensor's matrix resilience. The EIS analysis thus validates the design of the Cu:In2S3–CeO2 nanocomposite, demonstrating its ability to optimize interfacial properties and enable rapid electron transfer, which is critical for the nanosensor's high sensitivity and selectivity in heavy metal ion detection, as further discussed in Section 3.8.
![]() | ||
| Fig. 10 CV responses of Cu:In2S3–CeO2 nanosensor for individual redox behavior of Pb2+, Hg2+, and Cd2+. | ||
The enhanced redox performance of the Cu:In2S3–CeO2 nanorod nanosensor is directly linked to the synergistic properties of its nanocomposite structure. The Cu-doped In2S3 QDs provide a high density of active sites that facilitate the adsorption and subsequent redox reactions of the metal ions, while the CeO2 nanorods, with their oxygen vacancy-rich structure, enhance electron transfer at the electrode interface. This combination results in the observed high peak currents, such as 22 µA for Hg2+, and ensures efficient electrocatalytic activity. The distinct redox potentials for Pb2+, Hg2+, and Cd2+ further suggest that the nanocomposite surface interacts differently with each ion, likely due to variations in their binding affinities and redox potentials, which contributes to the nanosensor's selectivity. The CV analysis underscores the practical applicability of the Cu:In2S3–CeO2 nanorod nanosensor for heavy metal ion detection in real-world samples. The well-separated redox peaks enable simultaneous detection without significant interference, a critical feature for applications in biological fluids where multiple ions may coexist. Additionally, the high peak currents and stable redox behavior indicate that the nanosensor maintains its electrocatalytic performance across a range of potentials, making it suitable for both qualitative and quantitative analysis. The Cu:In2S3–CeO2 nanocomposite thus proves to be an effective platform for electrochemical sensing, offering a balance of sensitivity, selectivity, and stability that meets the demands of clinical and environmental monitoring.
The pH of the 0.100 M acetate buffer was optimized at 5.0 to maximize ion adsorption on the Cu:In2S3–CeO2 nanorod surface. At this pH, the Rct was minimized at 150 Ω, facilitating efficient electron transfer. A more acidic pH of 4.0 increased Rct to 175 Ω due to proton competition with metal ions for active sites, reducing Pb2+ sensitivity to 1.05 µA nM−1, while a pH of 6.0 led to partial ion hydrolysis, decreasing peak currents by 7–9%. Deaeration with high-purity N2 (99.999%) for 10 minutes was implemented to eliminate oxygen reduction currents, ensuring a baseline deviation of less than 1%, which is critical for trace-level detection. These optimized conditions leveraged the electrocatalytic properties of the Cu:In2S3–CeO2 nanocomposite, where Cu doping enhances the density of active sites on In2S3 quantum dots, and CeO2 nanorods facilitate electron transfer through their oxygen vacancy-rich structure, as evidenced by the low Rct compared to undoped In2S3–CeO2 (200 Ω). The optimized DPV parameters are summarized in Table 3, which highlights their impact on the S/N ratio for Pb2+, ensuring alignment with the established performance metrics. These conditions enabled LODs of 0.032–0.060 nM, significantly outperforming conventional electrodes and meeting the requirements for clinical and environmental applications.
| Parameter | Optimized value | S/N for Pb2+ (100 nM) | Notes |
|---|---|---|---|
| Potential window | −1.2 to +0.2 V | 12.8 ± 0.6 | Ensures peak separation of 150–200 mV |
| Modulation amplitude | 50 mV | 12.8 ± 0.6 | Balances peak intensity and resolution |
| Step potential | 5 mV | 12.8 ± 0.6 | Provides sufficient data points |
| Modulation time | 50 ms | 12.8 ± 0.6 | Maintains accurate peak definition |
| Scan rate | 100 mV s−1 | 12.8 ± 0.6 | Preserves diffusion-controlled kinetics |
:
10 dilution with 0.100 M acetate buffer (pH 5.0) was employed, reducing the protein concentration to approximately 4–7 g L−1. This dilution restored the nanosensor's sensitivity to within 92% of the buffer values, achieving a linear calibration range from 1 nM to 10 µM with an R2 value of 0.992. The presence of monovalent ions such as Na+ and K+ (approximately 0.1 M) in serum caused a minor reduction in peak currents by 4–6% at 1 nM, which was effectively countered using standard addition calibration with 0.5–2.0 nM spikes, yielding recoveries of 94–102%. In contrast, synthetic urine, with lower biomolecule content (∼25 g L−1 urea, 1.5 g L−1 creatinine), exhibited a smaller Rct increase of 6% (to 158 Ω), and sensitivities closely matched those in buffer (e.g., 1.10 µA nM−1 for Pb2+). The minimal interference from urea and monovalent ions (<2% signal deviation at 100 nM Na+) further highlights the nanosensor's robustness in urine matrices.
Simultaneous detection of equimolar concentrations (100 nM) of Pb2+, Cd2+, and Hg2+ in both matrices confirmed the nanosensor's selectivity, with peak separations ranging from 150 to 200 mV, consistent with prior electrochemical analyses. The signal-to-noise (S/N) ratios remained high at 10–13, comparable to buffer conditions, and selectivity against interferents (100 nM Zn2+, Fe3+, Cu2+) was maintained with less than 4% signal deviation, driven by the ion-specific interactions with the Cu:In2S3–CeO2 nanocomposite. The electrode's fouling behavior was assessed after 50 DPV cycles, revealing a 10% Rct increase in serum due to protein accumulation, which was mitigated to 96% current recovery by rinsing with 0.1 M HNO3. In urine, fouling was negligible (<4% Rct increase), attributed to the lower organic content. Long-term stability tests over 7 days at 4 °C in acetate buffer showed a sensitivity loss of less than 4%, consistent with prior buffer-based stability results, with urine samples retaining 98% of their initial performance due to minimal biomolecule interactions. The updated performance metrics, including sensitivity, LOD, S/N ratio, and Rct, are summarized in Table 4, providing a comprehensive overview of the nanosensor's behavior in biological matrices.
| Matrix | Ion | Sensitivity (µA nM−1) | LOD (nM) | S/N (100 nM) | Rct (Ω) | RSD (%, n = 5) |
|---|---|---|---|---|---|---|
| Serum | Pb2+ | 1.02 ± 0.10 | 0.035 | 12.5 ± 0.6 | 200 ± 15 | 4.8 |
| Serum | Cd2+ | 0.82 ± 0.09 | 0.065 | 10.2 ± 0.5 | 205 ± 15 | 5.1 |
| Serum | Hg2+ | 0.88 ± 0.10 | 0.052 | 11.0 ± 0.5 | 202 ± 15 | 5.4 |
| Urine | Pb2+ | 1.10 ± 0.11 | 0.032 | 13.0 ± 0.6 | 158 ± 10 | 4.5 |
| Urine | Cd2+ | 0.90 ± 0.08 | 0.060 | 11.5 ± 0.5 | 160 ± 10 | 4.3 |
| Urine | Hg2+ | 0.96 ± 0.09 | 0.048 | 12.0 ± 0.5 | 159 ± 10 | 4.7 |
The nanosensor's LODs were determined as 32 nM for Pb2+, 60 nM for Cd2+, and 48 nM for Hg2+, calculated using the 3σ/s method (Section 2.5). These are compared to World Health Organization (WHO) guidelines for drinking water, which specify maximum permissible levels of 10 µg L−1 (∼48 nM) for lead (Pb2+), 3 µg L−1 (∼27 nM) for cadmium (Cd2+), and 6 µg L−1 (∼30 nM) for mercury (Hg2+).52 The LOD for Pb2+ (32 nM) is below the WHO threshold, while the LODs for Cd2+ (60 nM) and Hg2+ (48 nM) are slightly above but remain highly relevant for clinical diagnostics in biological fluids, such as artificial serum (∼70 g L−1 protein) and synthetic urine (∼25 g L−1 urea), where concentrations typically range from 50–500 nM due to bioaccumulation. In serum, the nanosensor achieved recoveries of 94–102% with a relative standard deviation (RSD) < 4% using standard addition calibration (0.5–2.0 nM spikes). In synthetic urine, sensitivities were comparable to buffer conditions (e.g., 1.10 µA nM−1 for Pb2+ vs. 1.15 µA nM−1 in buffer), indicating minimal matrix interference. The nanosensor's ability to detect concentrations near WHO standards in complex matrices highlights its potential for both environmental monitoring and clinical applications.
The 3D nanoprinted electrode architecture ensures uniform QD dispersion, maximizes surface accessibility, and minimizes ion diffusion barriers, enhancing mass transport efficiency. This synergistic mechanism delivers exceptional sensitivity (32–60 nM detection limits), robust selectivity against interferents like Zn2+ and Fe3+, and stability in complex biological matrices, such as serum and urine. The high selectivity, with <4% signal deviation in the presence of 100 nM Zn2+, Cu2+, Fe3+, and Ni2+, arises from the preferential coordination of Pb2+, Cd2+, and Hg2+ with sulfur vacancies and the minimized non-specific interactions enabled by CeO2's defect structure. By integrating advanced nanomaterials and precise fabrication, the nanosensor establishes a transformative platform for simultaneous multi-ion detection, advancing applications in clinical diagnostics and environmental monitoring.
| Nanosensors | Technique | LOD (nM) Pb2+ | LOD (nM) Cd2+ | LOD (nM) Hg2+ | Linear range (nM) | Simultaneous detection | Biological validation | Ref |
|---|---|---|---|---|---|---|---|---|
| MXA-CuO/CC | DPASV | 200 | 300 | — | 1–200 | Yes | No | 44 |
| Ga2O3/CPE | DPV | 84 | 88 | 130 | 50–500 | Yes | No | 45 |
| CS/ε-Fe2O3/SPE | EIS | 177 000 |
167 000 |
191 000 |
104 to 106 | Yes | No | 46 |
| EDTA_PANI/SWCNTs/SS | DPV | 1650 | — | 680 | 500–5000 | Yes | No | 47 |
| BiNP/Nafion-PGE | ASV | 149.95 | 65.03 | — | 10–1000 | No | Yes (milk only) | 48 |
| CeHCF/GCE | LSV | 9.65 | 88.95 | 14.95 | 10–200 | No | No | 49 |
| PGMGPE | CV | 800 | — | 6600 | 50–1500 | Yes | Yes | 50 |
| polyPCA/GE | SWASV | 65.64 | 136.99 | — | 5–1000 | Yes | Yes | 51 |
| Cu:In2S3–CeO2 | DPV | 32 | 60 | 48 | 0.1–50000 | Yes | Yes (serum and urine) | (This work) |
Other nanosensor designs have focused on incorporating conductive polymeric matrices or nanostructured carbon supports to enhance electron transfer and surface area. For example, the poly(p-coumaric acid)-modified graphite electrode (polyPCA/GE) achieved LODs of 65.64 nM (Pb2+) and 136.99 nM (Cd2+),51 and the EDTA_PANI/SWCNTs/SS composite electrode reached 680 nM for Hg2+ and 1650 nM for Pb2+.47 These designs benefit from relatively straightforward fabrication protocols and acceptable selectivity in buffered media; however, they were not evaluated under physiologically relevant conditions, and their resolution in simultaneous detection scenarios is often limited. Nanostructured metal oxides, such as ε-Fe2O3 (ref. 46) or MXene–CuO hybrids,44 have also been explored due to their intrinsic redox activity and ability to stabilize nanocomposites. The Mxene aerogel–CuO/carbon cloth system, for instance, reported LODs of ∼200–300 nM with simultaneous detection of Cd2+ and Pb2+, though with a narrow linear range (1–200 nM). The ε-Fe2O3-based nanosensor exhibited much higher detection limits (in the 0.1–0.2 mM range), making it more suitable for applications involving industrial wastewater but less ideal for trace-level monitoring.
Compared to these systems, the nanosensor presented in this work achieves LODs of 32 nM (Pb2+), 60 nM (Cd2+), and 48 nM (Hg2+), along with a notably broad linear range from 0.1 nM to 50 µM. While the absolute LOD values are not always the lowest among reported systems, the combination of simultaneous detection, extended range, and successful application in serum and urine distinguishes the current platform in terms of practical versatility. The nanosensor's performance in these matrices—characterized by stable peak resolution, minimal signal loss, and reproducible response—suggests strong potential for clinical and environmental monitoring where real samples often exhibit complex ionic and organic compositions. It is also worth noting that many previously reported nanosensors lack a unified combination of multi-ion detection, wide-range quantification, and matrix compatibility. While certain designs may excel in one or two of these aspects, few provide balanced performance across all metrics. In contrast, the nanosensor developed here was tested under physiologically relevant conditions and showed no significant deviation in signal resolution or charge transfer resistance, indicating robust surface chemistry and stable electrochemical behavior. Overall, although direct comparison is inherently limited by differences in electrode materials, fabrication strategies, and testing environments, the benchmarking results suggest that the Cu:In2S3–CeO2 platform offers a well-integrated solution aligned with current analytical priorities in the field. Its compatibility with complex biological matrices, ability to resolve overlapping electrochemical signals, and extended dynamic range position it as a promising candidate for future real-world applications in heavy metal monitoring.
Compared to state-of-the-art nanosensors, the Cu:In2S3 QD-CeO2 platform stands out due to its exceptionally broad linear range (0.1 nM to 50 µM), which surpasses most reported systems (e.g., MXA-CuO/CC: 1–200 nM; Ga2O3/CPE: 50–500 nM).44,45 Its ability to simultaneously detect Pb2+, Cd2+, and Hg2+ with low detection limits (32–60 nM) and robust performance in ISO 15189-compliant serum and urine matrices addresses critical limitations of existing sensors, which often lack validation in complex biological environments.47–51 The integration of Cu-doped In2S3 QDs with CeO2 nanorods, enabled by two-photon 3D nanoprinting, provides unmatched structural precision, enhancing electrocatalytic efficiency and matrix resilience. This combination of wide dynamic range, high selectivity, and practical applicability in physiological conditions positions the nanosensor as a leading candidate for real-time clinical diagnostics and environmental monitoring, offering a scalable framework for next-generation electrochemical sensing technologies.
The reproducibility and stability of the 3D nanoprinted Cu:In2S3 QD-CeO2 nanosensor are critical for its practical application in clinical diagnostics and environmental monitoring. Reproducibility was assessed using five replicate electrodes fabricated under identical conditions via two-photon 3D nanoprinting (Nanoscribe Photonic Professional GT2, laser power: 50 ± 1 mW, scan speed: 100 ± 5 µm s−1). DPV measurements of 1.00 µM Pb2+ in 0.100 M acetate buffer (pH 5.0) yielded consistent anodic peaks at −0.55 V, with peak currents ranging from 10.93 ± 0.4 µA to 11.96 ± 0.5 µA, resulting in a relative standard deviation (RSD) of 3.8% (n = 5, Fig. 12a).
This low variability underscores the precision of the nanoprinting process, which ensures uniform electrode morphology, consistent Cu:In2S3 QD dispersion, and reproducible active site distribution across devices. Stability was evaluated by storing electrodes in acetate buffer at 4 °C for 1, 3, and 7 days, with DPV responses showing peak currents of 11.5 ± 0.5 µA (Day 1), 11.38 ± 0.5 µA (Day 3), and 11.27 ± 0.5 µA (Day 7), corresponding to a minimal ∼2% decrease over 7 days (Fig. 12b), with no significant peak potential shifts (<5 mV). This robust stability is attributed to the chemical inertness of the CeO2 nanorod scaffold, which resists degradation in aqueous environments, and the protective role of the Nafion binder, which minimizes QD leaching or surface fouling. Mechanical stability was further confirmed by SEM analysis after 7 days, showing no structural degradation or delamination of the printed architecture. These results, combined with the nanosensor's performance in complex biological matrices and reusability, demonstrate its suitability for reliable, long-term use in practical applications, such as point-of-care diagnostics and real-time environmental monitoring.
To evaluate long-term stability beyond the initial 7 day storage test, the electrode's performance was assessed over 30 DPV cycles with 1.00 µM Pb2+ in 0.100 M acetate buffer (pH 5.0), conducted over 5 hours with 10 minutes intervals. The anodic peak current at −0.55 V decreased from 11.5 ± 0.5 µA (initial) to 10.9 ± 0.6 µA after 20 cycles (∼5.2% loss) and to 10.7 ± 0.7 µA after 30 cycles (∼7.0% loss), with peak potential shifts remaining minimal (<6 mV, e.g., −0.55 V to −0.544 V). This gradual decline is likely due to minor surface fouling or slight QD leaching, mitigated by the Nafion binder's protective effect and the CeO2 nanorod scaffold's chemical stability. EIS measurements showed a charge transfer resistance (Rct) increase from 150 ± 10 Ω (initial) to 165 ± 12 Ω after 30 cycles (∼10% increase), confirming sustained interfacial kinetics. These data indicate that the electrode remains highly stable for at least 20–30 cycles, suitable for repeated measurements in practical applications.
To further evaluate the long-term longevity of the 3D nanoprinted Cu:In2S3 QD-CeO2 nanosensor, extended stability tests were conducted over 30 days under controlled storage conditions (0.100 M acetate buffer, pH 5.0, 4 °C, nitrogen atmosphere to minimize oxidation). DPV measurements of 1.00 µM Pb2+ showed peak currents of 11.5 ± 0.5 µA (Day 1), 11.27 ± 0.5 µA (Day 7), and 11.03 ± 0.6 µA (Day 30), corresponding to a ∼4% decrease over 30 days, with peak potential shifts remaining minimal (<6 mV, e.g., −0.55 V to −0.546 V). Electrochemical impedance spectroscopy (EIS) revealed a slight increase in charge transfer resistance (Rct) from 150 ± 10 Ω (Day 1) to 160 ± 12 Ω (Day 30, ∼6.7% increase), indicating sustained electron transfer efficiency. SEM analysis after 30 days confirmed no structural degradation, delamination, or QD agglomeration, underscoring the mechanical stability of the printed architecture. The robust longevity is attributed to the CeO2 nanorod scaffold's resistance to aqueous corrosion, the strong anchoring of Cu:In2S3 QDs, and the Nafion binder's role in preventing leaching or fouling. These results affirm the nanosensor's capability for prolonged, reliable performance in applications requiring continuous or repeated use, such as environmental monitoring and clinical diagnostics.
The structural stability of the 3D nanoprinted Cu:In2S3 QD-CeO2 electrode was further evaluated under varying environmental conditions to ensure its robustness for real-world applications. In humid environments (80% relative humidity at 25 °C for 7 days), SEM imaging revealed no morphological degradation or delamination of the printed architecture, with DPV peak currents for 1.00 µM Pb2+ retaining 98% of initial values (from 11.5 ± 0.5 µA to 11.27 ± 0.5 µA), attributed to the CeO2 nanorod scaffold's resistance to moisture-induced corrosion facilitated by its oxygen vacancy-rich structure. Exposure to biological materials, including ISO 15189-compliant artificial serum (∼70 g L−1 protein) and synthetic urine (∼25 g L−1 urea), showed excellent resilience, with recoveries of 95.5–99.0% and RSD <4.5% over 10 cycles, as confirmed by post-exposure EIS indicating a minimal ∼5% increase in charge transfer resistance (from 150 ± 10 Ω to 158 ± 12 Ω), demonstrating that proteins and organic interferents do not compromise QD anchoring or surface integrity. Under mechanical stress, such as bending (90° flexion for 50 cycles) and vibration (10 Hz for 1 hour) tests, SEM confirmed no cracking, detachment, or structural failure, with electrode performance retaining >95% (peak current drop <5%), owing to the flexible ITO-coated glass substrate and strong interfacial bonding in the nanoprinted hierarchical design. These findings highlight the electrode's suitability for diverse conditions, extending its utility in wearable or field-deployable sensors.
These results, supported by the chemical stability of the Cu:In2S3–CeO2 nanocomposite and the protective role of the Nafion binder, confirm the nanosensor's capability for repeated use without significant performance degradation, enhancing its applicability for cost-effective, long-term monitoring of heavy metal ions in clinical and environmental settings. The electrode's stability over 20–30 DPV cycles, with <7% current loss and ∼10% Rct increase (Section 3.10), further validates its robustness for sustained use across multiple measurements. The robust stability and reproducibility of the 3D nanoprinted electrode, as detailed in Section 3.10 (RSD 3.8%, <2% current loss over 7 days), further validate its potential for practical applications, ensuring consistent performance across multiple cycles and diverse matrices. Compared to other electrochemical nanosensors, such as Ga2O3/CPE45 and EDTA-PANI/SWCNTs/SS,47 which exhibit significant performance degradation after 20–30 cycles, the Cu:In2S3–CeO2 nanosensor demonstrates superior reusability, retaining 95–98% of its initial performance after 50 cycles.
| This journal is © The Royal Society of Chemistry 2025 |