Zainab Ishfaqa,
Layla A. Almutairib,
M. Yasir Alia,
Salhah Hamed Alrefaeec,
Mohamed Abdelsabour Fahmyde,
Elsammani Ali Shokrallaf,
Lamiaa G. Alharbeg,
Adnan Ali*a,
Arslan Ashfaq*a and
A. R. Abd-Elwahedh
aDepartment of Physics, Government College University, Faisalabad, 38000, Pakistan. E-mail: adnnan_1982@yahoo.com; arslan.ashfaq201@gmail.com
bDepartment of Biology, College of Science Princess Nourah bint Abdulrahman University, P. O. Box 84428, Riyadh 11671, Saudi Arabia
cDepartment of Chemistry, Faculty of Science, Taibah University, Yanbu 30799, Saudi Arabia
dDepartment of Mathematics, Adham University College, Umm Al-Qura University, Adham, 28653, Makkah, Saudi Arabia
eDepartment of Basic Sciences, Faculty of Computers and Informatics, Suez Canal University, New Campus, 41522 Ismailia, Egypt
fDepartment of Physics, Faculty of Science, Al-Baha University, Alaqiq, 65779-7738, Saudi Arabia
gDepartment of Physics, Aljamoum University College, Umm Al-Qura University, Makkah, Saudi Arabia
hDepartment of Physics, College of Science, Qassim University, Buraydah 51452, Saudi Arabia
First published on 4th November 2024
Surface Enhanced Raman Spectroscopy (SERS) is a highly sensitive analytical technique used for fingerprint recognition of molecular samples. The SERS effect, which enhances Raman scattering signals, has been the subject of extensive research over the past few decades. More recently, the commercialization of portable Raman spectrometers has brought SERS closer to real-world applications. The aim of the study was to enhance their performance, properties, and biocompatibility for potential use as SERS substrates. The synthesis and characterization of MoS2 and SnS2 nanoparticles are described, along with the functionalization process using L-cysteine. The detection and identification of Escherichia coli (E. coli) bacteria using MoS2 and SnS2 as SERS substrates are also investigated. The results demonstrate the successful functionalization and characterization of the nanostructures, indicating their potential as SERS substrates. The abstract highlights the importance of developing cost-effective and environmentally friendly disposable analysis chips with high accuracy and specificity for practical SERS applications.
In order to expand the practical utilization of SERS in real-life scenarios, it is imperative to develop cost-effective, environmentally friendly disposable analysis chips that offer both high accuracy and specificity.6,7 Additionally, it is crucial to mitigate the risks associated with cross-contamination and false positives. These factors play a crucial role in ensuring the reliability and effectiveness of SERS in various applications. Numerous techniques have been documented in the literature for fabricating structured SERS-active substrates, including dip coating, spin-coating, electrochemical synthesis, chemical vapor deposition, soft lithography, etching and electron beam lithography.8–10 However, it is important to note that these methods possess certain limitations with regards to either throughput volume or cost implications.11,12 In addition, achieving consistent signal intensity across different regions poses a significant challenge when it comes to the mass production of SERS-active nanostructures.13,14
The remarkable physical and chemical properties exhibited by 2D materials have sparked growing interest among researchers and scientists. 2D nanomaterials offer distinctive advantages, including facile synthesis, significant specific surface areas, remarkable mechanical properties, excellent optical properties, and favorable biocompatibility.15,16 These advantages contribute to their practical application in enhancing SERS and provide a viable solution to overcome challenges associated with metal substrates, such as high costs, catalytic effects, strong metal–adsorbate interactions, and photobleaching.17,18 As a result, numerous 2D materials have been extensively explored as potential SERS substrates.
Molybdenum disulfide (MoS2) has garnered significant attention in the field of materials science over the past decade due to its layered structure, resembling graphite, and exhibiting distinct anisotropic electrochemical, electronic, and optical properties. These properties make it highly relevant in various applications such as biology, physicochemistry, optics, imaging, sensing, therapy, and intercalation agents.19–21 Surface functionalization of MoS2 involves modifying its properties through the covalent bonding of particles to the single-layer nanosheets. Scanning tunneling microscopy (STM) serves as direct evidence for the covalent functionalization of transition metal dichalcogenides (TMDs), including MoS2.22,23 The functionalized MoS2 with a large surface area can increase its surface area and improve its effectiveness in interacting with other materials by adding functional groups or nanostructures to its surface. Its exceptional hydrophilicity provides a number of advantages, making it a promising material in a variety of biomedical applications.24 Overall, functionalized MoS2 nanostructures are important for SERS detection because they increase sensitivity, provide a large surface area for molecule adsorption, maintain chemical stability, have tunable plasmonic properties, are biocompatible, and can be easily integrated into sensing platforms.25,26
SnS2 is a layered, n-type semiconducting material with a hexagonal structure and an indirect bandgap of 2.23 eV. Similar to other TMDs, SnS2 consists of tin atoms sandwiched between two sulfur layers with covalent bonding, while each monolayer is held together by van der Waals forces.27 The structural properties of SnS2, including interlayer distances, binding energies, and in-plane lattice parameters, exhibit minimal variation across layers. The layer-dependent Raman spectra of SnS2 exhibit a slight increase in the frequencies of the Raman-active modes as the number of layers increases, while their intensities display a significant enhancement. The investigation of electronic, excitonic, and vibrational properties of SnS2 materials opens up new avenues for understanding the key characteristics of 2D materials.28 Due to its impressive performance, SnS2 finds applications in environmental remediation. Biofunctionalized SnS2 nanoparticles possess a large surface area and exceptional hydrophilicity. These attributes improve the loading efficiency and capacity of antibodies in the bio-detection stage and ensure the functionality of immobilized protein biomolecules. The functionalization of SnS2 is characterized using scanning electron microscopy, electrochemical impedance spectroscopy, static water contact angle measurements, and cyclic voltammetry.29 Because of their distinct surface properties, MoS2 and SnS2 make ideal SERS substrates. When functionalized with the MoS2 and SnS2 nanoparticles, they can greatly boost the Raman signals of molecules adsorbed on their surfaces. Charge transfer between the analytes molecule and the substrate as well as the localized surface plasmon resonance (LSPR) phenomenon are the sources of this increase. MoS2 and SnS2 nanostructures are chemically stable, which is critical for ensuring the accuracy and reliability of SERS observations throughout time. Functionalization can improve their stability and avoid degradation, resulting in more consistent and precise detection. For the sensitive and targeted identification of biomolecules, pathogens, and other analytes in complicated biological samples, functionalized MoS2 and SnS2 nanostructures can be employed.30
Functionalized 2D MoS2/SnS2 nanostructures for SERS have been the subject of ongoing research, particularly in nanotechnology and materials science. The functionalization of MoS2 economical SERS substrate for label-free bilirubin detection in clinical diagnosis. A one-pot hydrothermal synthesis of Fe-doped MoS2 was created as SERS substrate. Fe-MoS2 NFs were employed to detect bilirubin in serum. The Fe-MoS2 NF SERS substrate has a linear detection range of 10−3–10−9 M and a low limit of detection (LOD) of 10−8 M.31 The improvement the SERS sensitivity was also investigated of the π-conjugated fluorinated 7,7,8,8-tetracyanoquinodimethane derivatives by using the charge-localization effect caused by 2D MoS2 flakes. A significant Raman signal amplification in SERS was achieved using a 2D hetero structure made of FnTCNQ nanostructures grown on a 2D MoS2 flake. The SERS enhancement factor of MB molecules on the ideal F4TCNQ/MoS2 nanocomposite substrate can reach up to 2.531 × 106, with a limit of detection (LOD) of 10−10 M. The SERS results for MB, Rhodamine 6G (R6G), and 4-aminothiophenol (4-ATP) molecules suggest that the FnTCNQ/MoS2 SERS platform is promising for the detection of trace molecules.32 The photo-assisted decorating of silver nanoparticles (Ag-NPs) by hydrothermally produced hexagonal-like tin disulfide (SnS2-NHs) for ultrasensitive detection of synthetic dyes. The Ag-NPs/SnS2 NHs nanostructure exhibits both a local electromagnetic effect from the Ag-NPs and an effective charge-transfer effect from the SnS2 NHs. Methylene Blue (MB) and tartrazine (TZ) were used to test the SERS performance of Ag-NPs/SnS2 NHs. The produced nanostructure has a large linear range (MB (10−3–10−10 M) and TZ (10−2–10−9 M), low limit of detection (MB (4.12 × 10−10 M) and TZ (3.01 × 10−9 M), and excellent enhancement factor (108 and 107 for MB and TZ, respectively).33 The simple SERS active gold functionalized SnS2 quantum dots (Au/SnS2 QDs) that can detect and photodegrade Hg2+ ions under visible light irradiation. Crystal violet (CV) dye is employed as the indirect Raman probe for SERS detection at an excitation laser of 532 nm. The Au hybrids had higher SERS activity than pristine-SnS2 due to a combination of electromagnetic and chemical enhancements. The detection limit of Au/SnS2 toward Hg2+ was determined to be 1.05 ng ml−1.34,35
In this study, we focused on investigating the surface functionalization of 2D molybdenum/tin chalcogenide nanostructures through covalent bonding. The aim was to enhance their performance, properties, and bio-compatibility. Furthermore, we explored the potential of these nanostructures as SERS substrates for the detection and identification of various analytes, including microorganisms like Escherichia coli (E. coli) Methylene Blue (MB).
CFU ml−1 = no. of colonies × dilution factor/volume of cultured plated |
CFU ml−1 = 61 × 105/0.1 ml |
CFU ml−1 = 6.1 × 107 ml−1 |
This selected diluted bacteria solution is then carried for the measurements of SERS using MoS2-L-Cys and SnS2-L-Cys.
Fig. 2 (a) Raman spectroscopy, (b) XRD, and (c) SEM analysis of annealed MoS2 nanoparticles at various temperatures ranging from 200 to 400 °C. |
In Fig. 2(b), the XRD pattern of MoS2 exhibits four diffraction peaks located at 2θ angles of 26.3°, 33.3°, 35.5°, and 38.9°, corresponding to the (004), (001), (103), and (102) planes, respectively. All of these peaks are in agreement with the JCPDS card no. 1010993. The observed broadening of the peaks confirms the pure phase and hexagonal structure of MoS2.
In Fig. 2(c), the SEM images of the MoS2 nanoparticles reveal their spherical morphology and high porosity. The agglomerations of the MoS2 nanoparticles were shown in the SEM images. The grain size of the nanoparticles was increased due to increase the post-annealing temperature.
For the testing of synthesized nano materials SERS substrate, we have used 0.1 M of MB on the prepared sample of MoS2 in Fig. 3. The sample of MoS2 has shown high signal to noise ratio with the lower signal intensity. To overcome this issue, we functionalized our prepared samples of MoS2 with L-Cys.
Fig. 4(a) to (c) display the Raman, XRD, and SEM images of the MoS2-L-Cys NPs after functionalization, annealed at different temperatures: 200, 300, and 400 °C, respectively. In Fig. 4(a), the MoS2-L-Cys structure exhibits five prominent vibrational modes.38 Functionalization leads to an increased intensity of the E1g vibrational plane, with a slight shift in position. The second vibrational mode, E12g, remains unaffected by functionalization. Before functionalization, the results showed an intensity increase in the 2LA mode with increasing thickness. However, for MoS2-L-Cys NPs, the relative intensity of 2LA decreases with increasing thickness. The interaction between 2H-MoS2 and L-cysteine induces changes in defect density or electronic properties on the MoS2 surface, resulting in different band behaviors. Notably, compared to bulk 2H-MoS2, the MoS2-L-Cys structure exhibits a significant increase in the intensity of E1g.
Fig. 4 (a)–(c) Raman spectroscopy data, XRD patterns, and SEM images of the MoS2-L-Cys NPs annealed at temperatures ranging from 200 to 400 °C. (a) Raman spectrum (b) XRD pattern (c) SEM images. |
In Fig. 4(b), the structural analysis of MoS2-L-Cys is observed using XRD. Four prominent Bragg's peaks are observed at (004), (100), (102), and (103) in addition to the peaks corresponding to L-cysteine at different angles. Upon functionalization with L-cysteine, the intensity of the MoS2 peaks increases; however, the sharpness of the peaks decreases.
In Fig. 4(c), the SEM analysis of MoS2-L-Cys NPs reveals a uniform distribution of particle sizes without any signs of aggregation.
Fig. 5 shows the SERS spectrum of 0.1 M of Methylene Blue for the testing of synthesized MoS2-L-Cys functionalized as a SERS substrate, which was annealed at temperatures of 200 °C, 300 °C and 400 °C. The samples annealed at 300 °C and 400 °C have shown significant enhancement in the SERS spectrum as compared to the sample annealed at 200 °C. Table 1 shows the detailed analysis of the Raman spectrum with corresponding bonds of MoS2.
Fig. 5 The Raman spectra of MB using MoS2-L-Cys as SERS substrate annealed at temperatures 200 °C, 300 °C and 400 °C. |
Raman spectra of MB (cm−1)this work | Raman spectra of MB (cm−1)39,43–45 | Corresponding bonds |
---|---|---|
445–479 | 449 | C–N–C |
500 | 502 | C–N–C |
671 | 670 | C–H |
769–953 | 768 | C–H |
1035 | 1030 | C–H |
1185 | 1184 | C–N |
1304 | 1301 | C–H |
1395 | 1396 | C–H |
1440 | 1442 | C–N |
1502 | 1513 | C–C |
1622 | 1618 | C–C |
Fig. 6 shows the different concentrations of MB ranging from 0.1–10−7 M for the least concentration of any organic molecule to be used for the SERS. Up to 1 μm concentration of MB we have observed a detectable high intensity.
Fig. 6 Different concentration of MB ranging from 0.1–10−7 M measured on MoS2 sample annealed at 300 °C. |
Fig. 7 presents the Raman spectra of E. coli, which was annealed at temperatures 300 °C and 400 °C. The Raman spectra were obtained using an excitation laser of 633 nm, and eight active Raman modes were measured at wavenumbers of 652, 828, 958, 1129, 1169, 1240, 1300, 1499, and 1580 cm−1.
Fig. 7 The Raman spectra of E. coli using MoS2-L-Cys as SERS substrate annealed at temperatures 300 °C and 400 °C. |
Table 2 shows the detailed analysis of the Raman spectra of E. coli using MoS2-L-Cys. The E. coli SERS peaks were examined for the substrates annealed at 300 °C and 400 °C. Strong peaks were observed at 828 cm−1 (O–P–O stretching), 1129 cm−1 (CS), and 1499 cm−1 (carbohydrates modes). Other peaks were observed at 652 cm−1 (C–H), 958 cm−1 (CH bonding), 1169 cm−1 (C–C), 1240 cm−1 (C–N), 1300 cm−1 (lipids), and 1580 cm−1 (lipids). These results demonstrate the potential of the SERS substrate for the detection of E. coli bacteria even at low concentrations. However, the substrate annealed at 300 °C did not show E. coli signals due to a high signal-to-noise ratio. The functionalization resulted in an increase in peak intensity, indicating a change in defect density and electronic properties on the surface, as well as a slight shift in the Raman spectra. The MoS2-L-Cys nanoparticles were utilized as SERS substrates for the detection of E. coli bacterial cells, even at low concentrations. Raman analysis revealed that these substrates have the potential to detect various pathogens.
Raman spectra of E. coli (cm−1)this work | Raman spectra of E. coli (cm−1)11,23,39 | Corresponding bonds |
---|---|---|
652 | 658 | C–H |
828 | 830 | O–P–O stretch |
958 | 960 | CH bending mode |
1129 | 1130 | CS |
1169 | 1161 | C–C |
1240 | 1245 | C–N |
1300 | 1330 | CH2 |
1499 | 1513 | Carbohydrates |
1580 | 1587 | Lipids |
Fig. 8 (a)–(c) XRD, Raman spectroscopy data, and SEM images of the SnS2 NPs annealed at different temperatures ranging from 300 to 400 °C. |
In Fig. 8(b), the purity and phase of the SnS2 samples are determined through the XRD pattern. The XRD pattern of SnS2 at different temperature ranges reveals distinct diffraction peaks. At 300 °C, the peak is consistently indexed as a hexagonal SnS2 phase. As the temperature decreases, the formation of both SnS and SnS2 is observed. The XRD analysis of SnS2 shows seven reflection peaks corresponding to (100), (002), (011), (012), (110), (111), (103), and (200) planes.
In Fig. 8(c), the morphology of the prepared SnS2 samples was examined using SEM analysis. The SEM investigation revealed that SnS2 exhibits nanoflakes with a hexagonal stacking structure.
We have used 0.1 M of MB on the prepared sample of MoS2 in Fig. 9 for the testing of synthesized nano materials SERS substrate. Fig. 9 shows that the sample of SnS2 has shown a higher signal-to-noise ratio with a lower signal intensity. To overcome this issue we functionalized our prepared samples of SnS2 with L-Cys.
Fig. 10(a) to (c) depict the Raman, XRD, and SEM images of the SnS2-L-Cys NPs annealed at different temperatures ranging from 300 to 400 °C. In Fig. 10(a), the Raman spectra of SnS2 exhibit a slight shift in their peaks. The interaction between SnS2 and L-cysteine is evident from the presence of the Eg peak at 204 cm−1, which is still weak due to the nano size effect. The Raman spectra at 312 cm−1 indicate the presence of the stretched out-plane peak mode, but a shift in position is observed after functionalization. L-Cysteine peaks at 720 cm−1, 843 cm−1, and 1020 cm−1 were observed at different temperatures, respectively. SnS2 with L-cysteine displayed Raman spectra that were not significantly different from the SnS2 spectra, but the interaction with L-cysteine affected the peak positions.
Fig. 10 (a)–(c) XRD, Raman spectroscopy data, and SEM images of the SnS2-L-Cys NPs annealed at different temperatures ranging from 300 to 400 °C. |
In Fig. 10(b), after functionalization of SnS2 with L-cysteine, five diffraction peaks corresponding to (100), (002), (011), (012), and (110) were observed, along with L-cysteine peaks at different angles and with varying intensities. The X-ray diffraction (XRD) pattern was used to determine the purity and phase of SnS2-L-cysteine. The intensity of the peaks increased after functionalization.
The SEM micro-images in Fig. 10(c) display the SnS2-L-Cys NPs annealed at different temperatures ranging from 300 to 400 °C. Following functionalization, the original morphology of both SnS2 and L-Cys elements is no longer apparent, and instead, a uniform distribution of particle sizes is observed.
Fig. 11 illustrates the SERS spectrum of 0.1 M of Methylene Blue for the testing of synthesized SnS2-L-Cys functionalized as a SERS substrate, which was annealed at the temperatures of 300 °C, 350 °C and 400 °C. The samples annealed at 350 °C and 400 °C have shown significant enhancement in the SERS spectrum as compared to the sample annealed at 300 °C. Table 3 shows the Raman analysis of MB using SnS2-L-Cys.
Raman spectra of MB (cm−1)this work | Raman spectra of MB (cm−1)39,43–45 | Corresponding bonds |
---|---|---|
448 | 449 | C–N–C |
500 | 502 | C–N–C |
675 | 670 | C–H |
773–949 | 768 | C–H |
1036 | 1030 | C–H |
1185 | 1184 | C–N |
1302 | 1301 | C–H |
1392 | 1396 | C–H |
1440 | 1442 | C–N |
1505 | 1513 | C–C |
1626 | 1618 | C–C |
Fig. 12 illustrates the tested results for different concentrations of MB ranging from 0.1–10−7 M for the least concentration of any organic molecule to be used for the SERS. The maximum 1 μm concentration of MB we have observed a detectable enhancement in the intensity of Raman signal.
Fig. 12 Different concentrations of MB ranging from 0.1–10−7 M measured on SnS2 sample annealed at 400 °C. |
In Fig. 13, the Raman spectra of E. coli using SnS2–L-Cys as a SERS substrate annealed at high temperatures, specifically 350 °C and 400 °C, are presented. Eleven active Raman modes were identified by utilizing a 633 nm excitation laser. These modes correspond to vibrations at 520, 582, 803, 983, 1034, 1098, 1137, 1290, 1340, 1400, and 1540 cm−1, which are associated with C–N–C, S–S, tyrosine, C–O–O, C–H, CS, CH2, NH2, and ring stretching of adenine, respectively. The results demonstrate that this substrate can be used for the detection of E. coli bacteria at low concentrations.
Fig. 13 The Raman spectra of E. coli using SnS2-L-Cys as SERS substrate annealed at different temperatures ranging from 300 to 400 °C. |
The functionalization increased in peak intensity, indicating a change in defect density and electronic properties on the surface, as well as a slight shift in the Raman spectra. The annealing temperature has a considerable impact on the characteristics of functionalized SnS2-L-cysteine SERS substrates. Higher annealing temperatures can improve the homogeneity and reproducibility of SERS signals and enhancement factors can also be affected by its annealing temperature which are critical for detecting low-concentration analytes in SERS applications. The adsorption and binding of molecules, such as the L-cysteine–SnS2 substrate surface are influenced by the annealing temperature. The higher temperatures of 400 °C can enable greater adsorption or binding interactions, improving the functionalized substrate stability, and increasing the intensity of the SERS signal.
The SnS2-L-Cys nanoparticles were utilized as SERS substrates for the detection of E. coli bacterial cells, even at low concentrations. Raman analysis revealed that these substrates have the potential to detect various pathogens. Table 4 shows the detailed analysis of Raman spectra of E. coli using SnS2–L-Cys as an SERS substrate.
Raman spectra of E. coli (cm−1)this work | Raman spectra of E. coli (cm−1)39,43,44 | Corresponding bonds |
---|---|---|
520 | 525 | C–N–C |
582 | 580 | S–S |
803 | 796–809 | Tyrosine |
983 | 960–996 | C–O–O |
1034 | 1030 | C–H |
1098 | 1100 | Aromatic CS |
1137 | 1126–1144 | CS |
1290 | 1300 | CH2 |
1340 | 1336–1342 | γ(NH2) |
1400 | 1396 | C–H |
1540 | 1551–1569 | Ring stretching (adenine) |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra05315j |
This journal is © The Royal Society of Chemistry 2024 |