Abstract
Metal nanoclusters (NCs) are novel materials with low cytotoxicity, high chemical stability, intense luminescence, etc. These characteristics are in great demand during biomarker detection and bioimaging. These properties of metal NCs are exploited by colorimetric, luminescent, and Raman tiny sensors. Neopterin (Nep) is used in medicine as a biomarker of inflammation and immune system activation, cancer, COVID-19, etc. The clusters of Au, Ag, and Cu with magic atom numbers m* equal to 2 and 10 were studied. Gibbs energy of complexation (E
b) has been established using density functional theory (DFT). The highest E
b was determined for the complexes of
1 Introduction
Nanoclusters (NCs) of coinage metals are specific functional nanoobjects. They are largely exploited in multiple studies and applied technologies due to their unique physicochemical characteristics. These characteristics lead to the efficient production of signals when interplaying with bio-analytes [1].
Coinage metal clusters are exploited nowadays in biocatalysis, diagnostics, bioimaging, and biosensors due to their chemical stability, high biocompatibility, sensitivity to the molecular environment, low photobleaching, and tunable optical properties [2,3]. The characteristics of gold (Au) NCs are exceptionally unique; for this reason, they are frequently used in nanotechnology as compared to other metals and non-metal materials [4]. Silver (Ag) clusters are significant mostly because of their intense luminescence [5]. Copper (Cu) NCs have several advantages over Au and Ag clusters; they are high yield, raw abundance, low cost, and presence as a trace element in biosystems [6].
Coinage metal NCs can be obtained using DNA templates [7,8] and protein matrices [9], and their physicochemical properties and photophysics are intensively studied [10]. Metal clusters are used in nanosensor fabrication because of their tunable luminescence and efficiency in the detection of various substances [11,12], including organic molecules, ions, etc.
The metal NCs, which follow conventional magic atom numbers (2, 10, 18, 36, 54, and 86), possess the most attractive properties. The “superatom” concept explaining the emergence of magic atom numbers was developed in 2008 [13]. This concept says that the metal core is a “superatom” stabilized when its amount of electrons (e−) and e−-accepting groups conditions specific magic numbers. The equation for a magic number m* is as follows:
where N Me is the amount of metal atoms, which is also the number of unpaired electrons, A is the amount of e−-accepting groups, and q is the charge of the cluster [14]. The e−-accepting groups may be various organic ligands including pterins.
Unconjugated pterins (Ptrs) are low-molecular-weight organic compounds with a natural origin; in living species, reduced pterins are important coenzymes. Their decay products, oxidized pterins, can serve as biochemical markers of mechanical trauma [15], oxidative stress [16], inflammation [17], gastrointestinal diseases [18], cancer [19], SARS-CoV-2 [20,21], etc. High-performance liquid chromatography and capillary electrophoresis are often used to detect pterins in biological fluids [22,23]. However, new techniques of pterin detection with improved precision, sensitivity, and cost-effectiveness are also being developed, for example, Raman spectroscopy with adsorption of carboxypterin on carbon nanopillars coated with Au [24]. Ag colloid can be used for the Raman detection of xanthopterin, isoxanthopterin, and 7,8-dihydrobiopterin [25].
Neopterin (Nep) is a degradation product of 7,8-dihydroneopterin (H2Nep), which is a strong antioxidant produced by macrophages [26]. Serum and urinary Nep levels are determined by the macrophage activity [17]. Urinary Nep and total Nep (Nep plus H2Nep) are biomarkers of oxidative stress [16] and cancer: it is established that preactivation of cell-mediated immunity is related to poor prognosis in cancer-affected organisms, as well as human immunodeficiency virus infected patients [27]. Monocytes and macrophages make Nep when stimulated by interferon-gamma released by T cells [28]. As a whole, Nep levels can be a sign of various diseases and pathologies, including atherosclerosis, arthritis, and severe cardiovascular diseases [29,30].
Vargas and Martínez studied Cu, Ag, and Au complexes with pterins and determined that metal anions, as well as ultra-small metal clusters, form non-conventional H-bonds with pterins, their dimers, and tetramers [31]. We have already simulated the interactions of Ag and Au clusters with pterin using density functional theory (DFT) [32,33]. We established that at alkaline pH > 8, the Raman spectrum of singly deprotonated Ptr−1 undergoes dramatic changes upon the addition of Ag. Moreover, triatomic clusters of both Ag (
The detection of Nep in serum, urine, and cerebrospinal can be performed cost-effectively using coinage metal NCs. Nep–Me NC complexes can be quantitively determined with UV–vis, luminescent, infrared, and SERS spectra. Therefore, quantum-chemical simulations for Nep–Me NC complexes are necessary to establish the potential benefits of quantitative determination of Nep in biological fluids.
The aim of the study was to establish the interactions between coinage metal ultra-small NCs with magic atom numbers m* = 2, 10, and Nep. First, the binding energies to establish the most stable complexes were found. Second, the nature of metal–organic bonds was determined using quantum-topological analysis. Third, UV–vis and Raman spectra of the complexes were obtained to find the systems with significant potential for experimental detection of Nep.
2 Materials and methods
Geometry optimization and binding energy evaluation in Orca 5.0.2 [35] have been done using PBE functional [36]. Dispersion corrections were made with the Grimme method (D3) [37]. The 6–31 G(d,p) basis set was used along with the LANLTZ pseudopotential for the electrons of Au, Ag, and Cu since this method showed good results for metal complexes of DNA bases [38]. The CPCM polarizable continuum model has been used to take the influence of H2O into account [39]. The geometries of Nep were treated by putting metal atoms near the atoms of pterin with e− lone pairs: nitrogen and oxygen. Thus, the complexation reaction between Nep and metal NC looked as follows:
The binding energy between Nep and metal NCs (E b) was established as follows: E b = E Nep + E Me – E Nep–Me. The calculation has been done for both the protonated form Nep0 and the deprotonated Nep−1 (pK a = 8 [40]).
The simulations of UV–vis and Raman spectra for the most energetically favorable complexes have been done. The absorption spectra were obtained using the time-dependent density functional theory, in particular, M062X functional [41], along with the def2-TZVP basis set [42]. This functional was shown to be accurate enough when simulating the spectra of the complexes for nucleobases bound to transition metals [43]. Finally, the Raman spectra were determined with B3LYP-D3/def2-TZVP. B3LYP shows sufficient results when compared to the experimental Raman spectra [44,45].
The natural bond orbital (NBO) analysis has been performed with B3LYP in Gaussian 16 [46]. The E (2) stabilization energy of the i → j ground state transitions has been obtained in accordance with the following equation:
where
Equation (4) is used to quantify the amount of charge transferred between
The NBO and Quantum theory “Atoms-in-Molecules” (QTAIM) analysis have been performed previously to investigate Ag and Au NC complexes of pterin with a similar methodology [32,33].
3 Results and discussion
3.1 Geometry of the naked clusters
Structure and Gibbs free energies of individual metal atoms and isolated clusters (

Geometry of isolated metal clusters, according to the PBE-D3/def2-TZVP method.
For the cationic triatomic clusters, interatomic distances equal to 2.63, 2.68, and 2.36 Å have been established for
Naked Au9 is a planar 2D cluster with a C2v symmetry point group. Ag9 and Cu9 possess Cs and C2v symmetry, respectively, yet with a non-planar 3D geometry (Figure 1). This is in agreement with the latest research on Ag and Cu clusters [48,49]. Cartesian coordinates of the naked clusters are presented in Supplementary Materials.
The Au10 cluster is planar; however, it possesses a D3h symmetry, which is in agreement with previous research [50]. Ag10 and Cu10 are D2d geometries, which does not contradict the literature data [51,52]. Upon detachment of an electron, D2d symmetry does not change, forming D2d
Therefore, the symmetry of the naked clusters depends on metal type, charge, and atom number. The geometries of Au, Ag, and Cu clusters are similar only in the case of
3.2 Complexes with magic atom number m* = 2
Optimization of the
Binding energies of metal clusters with Nep (in kcal mol−1) forming complexes with magic atom number 2
| Au 0 |
|
Au 2 |
|
|
| Nep0 | 17.3 | 35.9 | ||
| Nep−1 | 13.1 | 61.4 | ||
| Ag 0 |
|
Ag 2 |
|
|
| Nep0 | 7.1 | 18.6 | ||
| Nep−1 | 3.9 | 29.2 | ||
| Cu 0 |
|
Cu 2 |
|
|
| Nep0 | 19.6 | 29.7 | ||
| Nep−1 | 25.4 | 33.8 |
The neutral Me0 atom attaches Nep−1 with the following priority: Cu > Au > Ag. For

Nep–Me clusters with magic atom number m* = 2.
The N5 site is the most typical for Me attraction since Me forms anchoring Me–O and non-conventional hydrogen bonds with the side-substituent of Nep (the only exception is Nep−1–Au0 system where the bonding occurs through the N3 atom of Nep); obviously, the character of Au–N3 binding is different as compared to Au–N5 binding, probably the Au–N3 interaction is more electrostatic. Interestingly, Nep−1–
Therefore, for each complex
3.3 Complexes with magic atom number m* = 10
All the m* = 10 Au and Ag complexes attach metal atoms through the N5 site and are monodentate (Figure 3). The only exceptions are Nep−1–

Geometries of
Binding energies of metal clusters with Nep (in kcal mol−1) forming complexes with magic atom number m* = 10
| Au 9 |
|
Au 10 |
|
|
| Nep0 | 12.3 | 19.2 | ||
| Nep−1 | 20.5 | 23.2 | ||
| Ag 9 |
|
Ag 10 |
|
|
| Nep0 | 7.9 | 10.7 | ||
| Nep−1 | 12.1 | 17.6 | ||
| Cu 9 |
|
Cu 10 |
|
|
| Nep0 | 17.1 | 18.1 | ||
| Nep−1 | 29.5 | 25.8 |
As it was stated previously, some NCs change symmetry upon the interplay with Nep. Thus, the most stable complex of the Nep−1–
Summation of clusters changing (+) and not changing (−) their symmetry upon the binding with Nep
| Au 0 |
|
Au 2 |
|
|
| Nep0 | − | − | ||
| Nep−1 | − | − | ||
| Ag 0 |
|
Ag 2 |
|
|
| Nep0 | — | — | ||
| Nep−1 | − | − | ||
| Cu 0 |
|
Cu 2 |
|
|
| Nep0 | — | — | ||
| Nep−1 | − | — | ||
| Au 9 |
|
Au 10 |
|
|
| Nep0 | + | + | ||
| Nep−1 | − | + | ||
| Ag 9 |
|
Ag 10 |
|
|
| Nep0 | − | + | ||
| Nep−1 | + | + | ||
| Cu 9 |
|
Cu 10 |
|
|
| Nep0 | + | + | ||
| Nep−1 | − | + |
3.4 QTAIM analysis
QTAIM [54,55] was used to study the properties of metal bonds with Nep. The most stable complexes of Au and Cu with Nep−1 have been regarded. Previously, we have applied QTAIM to study the interactions of amino acids and pterin with Au [33,56] and Ag [32,57]. We have calculated five QTAIM parameters (Table 4): (1) density of all electrons ρ(r), (2) Laplacian of electron density ∇2 ρ(r), (3) Lagrangian kinetic energy G(r), (4) potential energy density V(r), and (5) energy density H(r).
Bader’s QTAIM theory BCP properties (all in Hartree, whereas bond energy E bond is in kcal mol−1)
| Complex | BCP | P(r) | ∇2 ρ(r) | G(r) | V(r) | H(r) | E bond |
|---|---|---|---|---|---|---|---|
| Nep−1–
|
Au–N5 | 0.1112 | 0.3433 | 0.1216 | −0.1574 | −0.3580 | 49.4 |
| Nep−1–
|
Au–O | 0.0698 | 0.2202 | 0.0671 | −0.0791 | −0.0120 | 24.8 |
| Nep−1–
|
Au–O1′ | 0.0373 | 0.11730 | 0.0318 | −0.0342 | −0.0024 | 10.7 |
| Nep−1–
|
Cu–N5 | 0.0816 | 0.3273 | 0.1036 | −0.1253 | −0.0217 | 39.3 |
| Nep−1–
|
Cu–O | 0.0671 | 0.3071 | 0.0908 | −0.1047 | −0.0140 | 32.8 |
| Nep−1–
|
Cu–O3′ | 0.0641 | 0.2938 | 0.0861 | −0.0988 | −0.0127 | 31.0 |
| Nep−1–
|
Au–N5 | 0.0762 | 0.2377 | 0.0750 | −0.0906 | −0.0156 | 28.4 |
| Nep−1–
|
Au–O | 0.0706 | 0.2507 | 0.0753 | −0.0880 | −0.0126 | 27.6 |
| Nep−1–
|
Au–O3′ | 0.0555 | 0.2104 | 0.0593 | −0.0660 | −0.0067 | 20.7 |
| Nep−1–Cu9 | Cu–N5 | 0.0953 | 0.3771 | 0.1241 | −0.1540 | −0.0298 | 48.3 |
| Nep−1–Cu9 | Cu–O | 0.0926 | 0.4860 | 0.1457 | −0.1699 | −0.0242 | 53.3 |
| Nep−1–Cu9 | Cu–O3′ | 0.0607 | 0.2905 | 0.0833 | −0.0939 | −0.0106 | 29.5 |
All bond critical points (BCPs) have a positive Laplacian, which signifies the following: all Me–X (X = O, N, C or H) interactions have an electrostatic nature. Moreover, all BCPs have negative H(r), which signifies that electrostatic bonding stabilizes Me–X interactions. Positive ∇2 ρ(r) and negative energy density H(r) for each BCP mean that all Me–X bonds are partially electrostatic and partially covalent. Therefore, QTAIM has shown that the interplay between Nep and Me clusters is partially electrostatic and partially covalent.
The Au–N5 bond (49.4 kcal mol−1) of the Nep−1–
Contrary to Gibbs binding energy (Tables 1 and 2), according to QTAIM, the bonding of ultra-small clusters with m* = 2 to Nep is not stronger than the bonding of m* = 10 NCs. The strongest bonding in terms of QTAIM bonding energy is observed for Cu–O in the Nep−1–Cu9 complex. On average, the bonding of O and N with Cu is stronger than Au–X bonding.
3.5 Absorption spectra of Nep–Me complexes
We have analyzed the first 20 transitions of the naked metal clusters and Nep–Me complexes. As one can see in Figure 4, the

UV–vis electronic spectra of
The spectrum of the naked
Cu9 has a major peak located in the visible region at 437 nm (0.237) and long-wave maximum with extremely low intensity: 992 nm (0.004). Upon attachment of Nep−1, these main maxima are shifted to 477 nm (0.055) and 797 nm (0.014); as a whole, the UV–vis spectrum is significantly transformed.
Upon the complexation of a metal NC and an organic molecule, the absorbance peaks should red-shift, since the electronic systems of a molecule and metal atoms unite to form a common entity. This is a well-known effect for similar systems regarding both absorption and fluorescence spectra: for example, Ag and Au clusters in complex with cysteine [58]. Similar effects (bathochromic shift of absorption spectra) were observed for phenylalanine and Au [59]. The larger system of electrons results in a red-shift of the spectra, both absorbance and fluorescence [58]. Indeed, in our case, the spectra of Nep−1–
All the above Nep–Me systems possess S0 → S1 transitions with zero or minimal intensity, which should them barely legal for luminescent detection of Nep, whereas colorimetric detection seems to be meaningful in each case.
3.6 Raman spectra of the most stable complexes
The Raman spectra of Nep−1 and the most stable complexes with magic atom number m* equal 2 and 10 have been simulated: Nep−1–
The naked Nep−1 Raman spectrum has four major bands with nearly equal intensity (Figure 5): at 1,330 cm−1 (664 a.u.), 1,601 cm−1 (631 a.u.), 3,088 cm−1 (686 a.u.), and 3,582 cm−1 (725 a.u.). The first band corresponds to “breathing” of the whole pteridine ring system. The 1,601 cm−1 band is responsible for C–NH2 bond stretching. The 3,088 and 3,582 cm−1 bands are about the C1′–H bond stretching and NH2 group symmetric stretching, respectively.

Raman spectra of Nep−1–
Three major peaks of Nep−1–
The spectrum of Nep−1–
The main transformations of the Nep−1 spectrum with the attachment of
The predominant peak of Nep−1–Cu9 is located at 3,387 cm−1 (10,283 a.u.). It relates to the O1′–H bond stretching. In the spectrum of the bare Nep−1, this transition is located at 3,671 cm−1 (110 a.u.). This means that upon the attachment of Cu9, this maximum is bathochromically shifted by 286 cm−1, whereas its intensity is enhanced by 93 times. The second highest maximum is at 1,588 cm−1 (2,654 a.u.), which refers to C6–C7 bond stretching. Cu9 interacts with all the functional groups of Nep: pyrimidine, pyrazine, and side substituent (Figure 3), which apparently makes Cu9 a selective and effective tool for Nep in vitro detection.
In real experimental conditions, the spectra are affected by other molecules, both low-molecular-weight compounds and biopolymers. Usually, performing the detection of pterin in aqueous solutions, we determine the optimal experimental conditions for cluster synthesis and pterin detection: this includes concentrations, pH, temperature, time of preparation, etc. [60]. When a protocol for pterin detection is determined, one may start to perform the experiments in biological liquids, for example, in human serum. Specific experimental conditions allow us to determine specific compounds (pterins) in the presence of other molecules. Moreover, when one utilizes Raman spectroscopy, it allows one to determine particular fingerprints of specific biomolecules [61], yielding unique Raman maxima and signal shifts. That is why SERS has attracted the attention of the researchers in recent years [62]. SERS allows thdetermination of specific substances in biological liquids even with a femtomole level limit of detection [63].
Analysis of the Raman spectra shows that complexes with m* = 10 are more promising tools for the experimental detection of Nep at alkaline pH than m* = 2 clusters. For example, the Cu9 cluster causes more than 90 times enhancement of the Raman signal.
3.7 NBO analysis
The chemical enhancement of the SERS signal can be expressed in terms of ground state charge transfer (CT) [64]. Thus, NBO analysis was used to evaluate ground-state CTs. Since NBO analysis was used to study SERS effects, the NBO parameters with the B3LYP-D3/def2-TZVP method was obtained. The B3LYP functional has already demonstrated its accuracy for the NBO analysis of similar systems [32,65,66,67].
The two most stable Nep–Me complexes with m* = 2 (Nep−1–
Second-order perturbation theory study of the Fock matrix in the NBO basis for the most stable Nep–Me systems
| Complex | Donor (i) | Orbital type | Occupancy | Acceptor (j) | Orbital type | Occupancy | E (2), kcal mol−1 | q CT |
|---|---|---|---|---|---|---|---|---|
| Nep−1–
|
N5 | n | 0.846 | Au | n* | 0.065 | 15.26 | 0.053 |
| Nep−1–
|
N5 | n | 0.899 | Cu | n* | 0.120 | 22.36 | 0.162 |
| Nep−1–
|
O3′ | n | 0.931 | Cu | n* | 0.065 | 15.44 | 0.106 |
| Nep−1–
|
N5 | n | 0.900 | Au | n* | 0.062 | 5.66 | 0.016 |
| Nep−1–Cu9 | N5 | n | 0.885 | Cu | n* | 0.075 | 22.29 | 0.079 |
| Nep−1–Cu9 | O | n | 0.917 | Cu | n* | 0.051 | 19.51 | 0.066 |
| Nep−1–Cu9 | O3′ | n | 0.941 | Cu | n* | 0.047 | 12.57 | 0.041 |
Table 5 represents E
(2) and q
CT values for the maximal ground state CTs (the highest E
(2) values for concrete complexes) with the participation of metal. Most CTs occur between metal atoms and N5 of Nep. For example, the nN5
→
Another example is that the ligand–metal CT complex is a nO3′
→
Regarding complexes with m* = 10, small CTs are equally distributed among various atoms and orbitals in the Nep−1–

Selected natural bond orbitals for ligand–metal CTs of the Nep−1–Cu9 complex.
4 Conclusion
The interactions between Nep and coinage metal (Au, Ag, and Cu) NCs were studied using the methods of quantum chemistry and DFT. In particular, the geometries and symmetry of the most stable naked clusters have been obtained. The symmetry of an isolated cluster depends on metal type, charge, and number of atoms. The most stable geometries of the naked clusters and clusters in complex with Nep are mostly different: upon complexation with Nep, the geometries change more significantly in the case of q = +1 clusters than in the case of neutral ones.
Gibbs free energies of complexation for complexes of
UV–vis absorption and Raman spectra have been calculated for a set of the most stable complexes with magic atom numbers m* = 2, 10: Nep−1–
Nep levels in blood serum are in the nM range [69]. These levels can be barely detected using absorption spectroscopy. However, Raman spectroscopy and SERS open many possibilities in detecting pterins and other biomarkers: SERS permits a 102–106-times chemical/electromagnetic signal enhancement [64]. Moreover, one can use aptamer-based [70] technology to develop nanosensors for Nep detection. Thus, this study opens many perspectives for pterin sensing in the future. Detection of Nep using SERS is the most promising using copper since the Nep spectrum undergoes significant chemical enhancement and maxima shift upon complexation with Cu9.
Nep detection in biological samples is of great necessity in medicine, and improvement of experimental techniques for Nep determination is much needed. The theoretical analysis has shown that the usage of copper, Au, and Ag NCs and nanoparticles is a promising tool for the detection of Nep biomarkers. It shows the possibility of NCs application in vitro using Raman spectroscopy and tunable colorimetry.
-
Funding information: The research was funded by the Russian Science Foundation, grant number 20-73-10029, https://rscf.ru/project/20-73-10029/.
-
Author contributions: Andrey A. Buglak is responsible for the entire work.
-
Conflict of interest: The author states no conflict of interest.
-
Data availability statement: The datasets generated during and/or analysed during the current study are available from the corresponding author on reasonable request.
References
[1] Zhang L, Wang E. Metal nanoclusters: New fluorescent probes for sensors and bioimaging. Nano Today. 2014;9(1):132–57.10.1016/j.nantod.2014.02.010Search in Google Scholar
[2] Du X, Jin R. Atomically precise metal nanoclusters for catalysis. ACS Nano. 2019;13(7):7383–7.10.1021/acsnano.9b04533Search in Google Scholar PubMed
[3] Bai Y, Shu T, Su L, Zhang X. Fluorescent gold nanoclusters for biosensor and bioimaging application. Crystals. 2020;10(5):357.10.3390/cryst10050357Search in Google Scholar
[4] Bai L, Zhu L, Ang CY, Li X, Wu S, Zeng Y, et al. Iron(III)-quantity-dependent aggregation–dispersion conversion of functionalized gold nanoparticles. Chem Eur J. 2014;20(14):4032–7.10.1002/chem.201303958Search in Google Scholar PubMed
[5] Xie Y-P, Shen Y-L, Duan G-X, Han J, Zhang L-P, Lu X. Silver nanoclusters: Synthesis, structures and photoluminescence. Mater Chem Front. 2020;4(8):2205–22.10.1039/D0QM00117ASearch in Google Scholar
[6] Babu Busi K, Palanivel M, Kanta Ghosh K, Basu Ball W, Gulyás B, Padmanabhan P, et al. The multifarious applications of copper nanoclusters in biosensing and bioimaging and their translational role in early disease detection. Nanomaterials (Basel, Switz). 2022;12(3):301.10.3390/nano12030301Search in Google Scholar PubMed PubMed Central
[7] Lee KX, Shameli K, Yew YP, Teow SY, Jahangirian H, Rafiee-Moghaddam R, et al. Recent developments in the facile bio-synthesis of gold nanoparticles (AuNPs) and their biomedical applications. Int J Nanomed. 2020;15:275–300.10.2147/IJN.S233789Search in Google Scholar PubMed PubMed Central
[8] Chen Y, Phipps ML, Werner JH, Chakraborty S, Martinez JS. DNA templated metal nanoclusters: From emergent properties to unique applications. Acc Chem Res. 2018;51(11):2756–63.10.1021/acs.accounts.8b00366Search in Google Scholar PubMed
[9] Tan SC, He Z, Wang G, Yu Y, Yang L. Protein-templated metal nanoclusters: Molecular-like hybrids for biosensing, diagnostics and pharmaceutics. Molecules. 2023;28(14):5531.10.3390/molecules28145531Search in Google Scholar PubMed PubMed Central
[10] Maity S, Bain D, Patra A. An overview on the current understanding of the photophysical properties of metal nanoclusters and their potential applications. Nanoscale. 2019;11(47):22685–723.10.1039/C9NR07963GSearch in Google Scholar PubMed
[11] Romeo MV, López-Martínez E, Berganza-Granda J, Goñi-de-Cerio F, Cortajarena AL. Biomarker sensing platforms based on fluorescent metal nanoclusters. Nanoscale Adv. 2021;3(5):1331–41.10.1039/D0NA00796JSearch in Google Scholar
[12] Qian S, Wang Z, Zuo Z, Wang X, Wang Q, Yuan X. Engineering luminescent metal nanoclusters for sensing applications. Coord Chem Rev. 2022;451:214268.10.1016/j.ccr.2021.214268Search in Google Scholar
[13] Häkkinen H. Atomic and electronic structure of gold clusters: Understanding flakes, cages and superatoms from simple concepts. Chem Soc Rev. 2008;37(9):1847–59.10.1039/b717686bSearch in Google Scholar PubMed
[14] Muniz-miranda F. Computational approaches to the electronic properties of noble metal nanoclusters protected by organic ligands. Nanomaterials. 2021;11(9):2409.10.3390/nano11092409Search in Google Scholar PubMed PubMed Central
[15] Lindsay A, Gieseg SP. Pterins as diagnostic markers of exercise-induced stress: A systematic review. J Sci Med Sport. 2020;23(1):53–62.10.1016/j.jsams.2019.08.018Search in Google Scholar PubMed
[16] Baxter-Parker G, Roffe L, Moltchanova E, Jefferies J, Raajasekar S, Hooper G, et al. Urinary neopterin and total neopterin measurements allow monitoring of oxidative stress and inflammation levels of knee and hip arthroplasty patients. PLoS One. 2021;16(8):e0256072.10.1371/journal.pone.0256072Search in Google Scholar PubMed PubMed Central
[17] Gieseg SP, Baxter-Parker G, Lindsay A. Neopterin, inflammation, and oxidative stress: What could we be missing? Antioxidants. 2018;7(7):80.10.3390/antiox7070080Search in Google Scholar PubMed PubMed Central
[18] Zhang L, Wang X, Ji X, Zou S. Changes of serum neopterin and its significance as biomarker in prediction the prognosis of patients with acute pancreatitis. J Lab Med. 2020;44(4):205–9.10.1515/labmed-2020-0013Search in Google Scholar
[19] Kośliński P, Daghir-Wojtkowiak E, Szatkowska-Wandas P, Markuszewski M, Markuszewski MJ. The metabolic profiles of pterin compounds as potential biomarkers of bladder cancer—integration of analytical-based approach with biostatistical methodology. J Pharm Biomed Anal. 2016;127:256–62.10.1016/j.jpba.2016.02.038Search in Google Scholar PubMed
[20] Hailemichael W, Kiros M, Akelew Y, Getu S, Andualem H. Neopterin: A promising candidate biomarker for severe COVID-19. J Inflamm Res. 2021;14:245–51.10.2147/JIR.S290264Search in Google Scholar PubMed PubMed Central
[21] Fuchs D, Gisslen M. Laboratory diagnostic value of neopterin measurements in patients with COVID-19 infection. Pteridines. 2021;32(1):1–4.10.1515/pteridines-2021-0001Search in Google Scholar
[22] Burton C, Shi H, Ma Y. Development of a high-performance liquid chromatography - tandem mass spectrometry urinary pterinomics workflow. Anal Chim Acta. 2016;927:72–81.10.1016/j.aca.2016.05.005Search in Google Scholar PubMed
[23] Grochocki W, Buszewska-Forajta M, Macioszek S, Markuszewski MJ. Determination of urinary pterins by capillary electrophoresis coupled with LED-induced fluorescence detector. Molecules. 2019;24(6):1166.10.3390/molecules24061166Search in Google Scholar PubMed PubMed Central
[24] Castillo J, Rozo C, Bertel L, Rindzevicius T, Mendez S, Martíneza F, et al. Orientation of pterin-6-carboxylic acid on gold capped silicon nanopillars platforms: Surface enhanced Raman spectroscopy and density functional theory studies. J Braz Chem Soc. 2015;27:971–7.10.5935/0103-5053.20150352Search in Google Scholar
[25] Smyth C, Mehigan S, Rakovich YP, McCabe EM, Bell SEJ. Pterin detection using surface-enhanced Raman spectroscopy incorporating a straightforward silver colloid-based synthesis technique. J Biomed Opt. 2011;16(7):1–6.10.1117/1.3600658Search in Google Scholar PubMed
[26] Eisenhut M. Neopterin in diagnosis and monitoring of infectious diseases. J Biomarkers. 2013;2013:196432.10.1155/2013/196432Search in Google Scholar PubMed PubMed Central
[27] Fuchs D, Hausen A, Reibnegger G, Werner ER, Dierich MP, Wachter H. Neopterin as a marker for activated cell-mediated immunity: Application in HIV infection. Immunol Today. 1988;9(5):150–5.10.1016/0167-5699(88)91203-0Search in Google Scholar PubMed
[28] Wachter H, Fuchs D, Hausen A, Reibnegger G, Werner ER. Neopterin as marker for activation of cellular immunity: Immunologic basis and clinical application. In: Spiegel HEBT-A in CC (ed). Vol. 27. Amsterdam, Netherlands: Elsevier; 1989. p. 81–141.10.1016/S0065-2423(08)60182-1Search in Google Scholar PubMed
[29] Reibnegger G, Egg D, Fuchs D, Günther R, Hausen A, Werner ER, et al. Urinary neopterin reflects clinical activity in patients with rheumatoid arthritis. Arthritis Rheum. 1986;29(9):1063–70.10.1002/art.1780290902Search in Google Scholar PubMed
[30] Watanabe T. Neopterin derivatives – a novel therapeutic target rather than biomarker for atherosclerosis and related diseases. Vasa. 2020;50(3):165–73.10.1024/0301-1526/a000903Search in Google Scholar PubMed
[31] Vargas R, Martínez A. Non-conventional hydrogen bonds: Pterins-metal anions. Phys Chem Chem Phys. 2011;13(28):12775–84.10.1039/c1cp20134dSearch in Google Scholar PubMed
[32] Buglak AA, Kononov AI. Silver cluster interactions with pterin: Complex structure, binding energies and spectroscopy. Spectrochim Acta Part A Mol Biomol Spectrosc. 2022;279:121467.10.1016/j.saa.2022.121467Search in Google Scholar PubMed
[33] Chebotaev PP, Plavskii VY, Kononov AI, Buglak AA. Pterin interactions with gold clusters: A theoretical study. Dyes Pigm. 2023;216:111323.10.1016/j.dyepig.2023.111323Search in Google Scholar
[34] Sharma B, Frontiera RR, Henry A-I, Ringe E, Van Duyne RP. SERS: Materials, applications, and the future. Mater Today. 2012;15(1):16–25.10.1016/S1369-7021(12)70017-2Search in Google Scholar
[35] Neese F, Wennmohs F, Becker U, Riplinger C. The ORCA quantum chemistry program package. J Chem Phys. 2020;152(22):224108.10.1063/5.0004608Search in Google Scholar PubMed
[36] Perdew JP, Burke K, Ernzerhof M. Generalized gradient approximation made simple. Phys Rev Lett. 1996;77(18):3865–8.10.1103/PhysRevLett.77.3865Search in Google Scholar PubMed
[37] Grimme S, Ehrlich S, Goerigk L. Effect of the damping function in dispersion corrected density functional theory. J Comput Chem. 2011;32(7):1456–65.10.1002/jcc.21759Search in Google Scholar PubMed
[38] Espinosa Leal LA, Lopez-Acevedo O. On the interaction between gold and silver metal atoms and DNA/RNA nucleobases – a comprehensive computational study of ground state properties. Nanotechnol Rev. 2015;4(2):173–91.10.1515/ntrev-2012-0047Search in Google Scholar
[39] Barone V, Cossi M. Quantum calculation of molecular energies and energy gradients in solution by a conductor solvent model. J Phys Chem A. 1998;102(11):1995–2001.10.1021/jp9716997Search in Google Scholar
[40] Lorente C, Thomas AH. Photophysics and photochemistry of pterins in aqueous solution. Acc Chem Res. 2006;39(6):395–402.10.1021/ar050151cSearch in Google Scholar PubMed
[41] Zhao Y, Truhlar DG. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other function. Theor Chem Acc. 2008;120(1):215–41.10.1007/s00214-007-0310-xSearch in Google Scholar
[42] Weigend F, Ahlrichs R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys Chem Chem Phys. 2005;7(18):3297–305.10.1039/b508541aSearch in Google Scholar PubMed
[43] Maksimov DA, Pomogaev VA, Kononov AI. Excitation spectra of Ag3–DNA bases complexes: A benchmark study. Chem Phys Lett. 2017;673:11–8.10.1016/j.cplett.2017.01.074Search in Google Scholar
[44] Birke RL, Lombardi JR. Simulation of SERS by a DFT study: A comparison of static and near-resonance Raman for 4-mercaptopyridine on small Ag clusters. J Opt. 2015;17(11):114004.10.1088/2040-8978/17/11/114004Search in Google Scholar
[45] Birke RL, Lombardi JR. TDDFT study of charge-transfer Raman spectra of 4-mercaptopyridine on various ZnSe nanoclusters as a model for the SERS of 4-Mpy on semiconductors. J Phys Chem C. 2018;122(9):4908–27.10.1021/acs.jpcc.7b12392Search in Google Scholar
[46] Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, et al. Gaussian16 Revision C.01. 2016: Gaussian 16, Revision C.01. Wallin: Gaussian, Inc.; 2016.Search in Google Scholar
[47] Lushchikova OV, Huitema DMM, López-Tarifa P, Visscher L, Jamshidi Z, Bakker JM. Structures of Cun + (n = 3–10) clusters obtained by infrared action spectroscopy. J Phys Chem Lett. 2019;10(9):2151–5.10.1021/acs.jpclett.9b00539Search in Google Scholar PubMed PubMed Central
[48] Jaque P, Toro-Labbé A. Polarizability of neutral copper clusters. J Mol Model. 2014;20:2410.10.1007/s00894-014-2410-6Search in Google Scholar PubMed
[49] Garg S, Kaur N, Goel N, Molayem M, Grigoryan VG, Springborg M. Properties of naked silver clusters with up to 100 atoms as found with embedded-atom and density-functional calculations. Molecules. 2023;28(7):3266.10.3390/molecules28073266Search in Google Scholar PubMed PubMed Central
[50] Shi HX, Sun WG, Kuang XY, Lu C, Xia XX, Chen BL, et al. Probing the interactions of O2 with small gold cluster AunQ (n = 2–10, Q = 0, −1): A neutral chemisorbed complex Au5O2 cluster predicted. J Phys Chem C. 2017;121(44):24886–93.10.1021/acs.jpcc.7b09022Search in Google Scholar
[51] Harb M, Rabilloud F, Simon D, Rydlo A, Lecoultre S, Conus F, et al. Optical absorption of small silver clusters: Agn, (N = 4–22). J Chem Phys. 2008;129(19):194108.10.1063/1.3013557Search in Google Scholar PubMed
[52] Die D, Zheng B-X, Zhao L-Q, Zhu Q-W, Zhao Z-Q. Insights into the structural, electronic and magnetic properties of V-doped copper clusters: Comparison with pure copper clusters. Sci Rep. 2016;6(1):31978.10.1038/srep31978Search in Google Scholar PubMed PubMed Central
[53] van der Tol J, Jia D, Li Y, Chernyy V, Bakker JM, Nguyen MT, et al. Structural assignment of small cationic silver clusters by far-infrared spectroscopy and DFT calculations. Phys Chem Chem Phys. 2017;19(29):19360–8.10.1039/C7CP03335DSearch in Google Scholar
[54] Bader RFW. A quantum theory of molecular structure and its applications. Chem Rev. 1991;91(5):893–928.10.1021/cr00005a013Search in Google Scholar
[55] Kumar PSV, Raghavendra V, Subramanian V. Bader’s theory of atoms in molecules (AIM) and its applications to chemical bonding. J Chem Sci. 2016;128(10):1527–36.10.1007/s12039-016-1172-3Search in Google Scholar
[56] Buglak AA, Kononov AI. Comparative study of gold and silver interactions with amino acids and nucleobases. RSC Adv. 2020;10(56):34149–60.10.1039/D0RA06486FSearch in Google Scholar
[57] Buglak AA, Ramazanov RR, Kononov AI. Silver cluster–amino acid interactions: A quantum-chemical study. Amino Acids. 2019;51(5):855–64.10.1007/s00726-019-02728-zSearch in Google Scholar PubMed
[58] Feng T, Chen Y, Feng B, Yan J, Di J. Fluorescence red-shift of gold-silver nanoclusters upon interaction with cysteine and its application. Spectrochim Acta Part A Mol Biomol Spectrosc. 2019;206:97–103.10.1016/j.saa.2018.07.087Search in Google Scholar PubMed
[59] Buglak AA, Kononov AI. Interactions of deprotonated phenylalanine with gold clusters: Theoretical study with prospects for amino acid detection. Spectrochim Acta Part A Mol Biomol Spectrosc. 2024;311:124004.10.1016/j.saa.2024.124004Search in Google Scholar PubMed
[60] Sych TS, Shekhovtsov NV, Buglak AA, Kononov AI. Amino acid-stabilized luminescent gold clusters for sensing pterin and its analogues. Anal Methods. 2024;16(27):4607–18.10.1039/D4AY00700JSearch in Google Scholar PubMed
[61] Cialla D, Pollok S, Steinbrücker C, Weber K, Popp J. SERS-based detection of biomolecules. Nanophotonics. 2014;3(6):383–411.10.1515/nanoph-2013-0024Search in Google Scholar
[62] Tran VA, Tran TTV, Le VT, Doan VD, Vo GNL, Tran VH, et al. Advanced nano engineering of surface-enhanced Raman scattering technologies for sensing applications. Appl Mater Today. 2024;38:102217.10.1016/j.apmt.2024.102217Search in Google Scholar
[63] Issatayeva A, Farnesi E, Cialla-May D, Schmitt M, Rizzi FMA, Milanese D, et al. SERS-based methods for the detection of genomic biomarkers of cancer. Talanta. 2024;267:125198.10.1016/j.talanta.2023.125198Search in Google Scholar PubMed
[64] Cong S, Liu X, Jiang Y, Zhang W, Zhao Z. Surface enhanced Raman scattering revealed by interfacial charge-transfer transitions. Innovation. 2020;1(3):100051.10.1016/j.xinn.2020.100051Search in Google Scholar PubMed PubMed Central
[65] Gangadharana RP, Krishnanb SS. Natural bond orbital (NBO) population analysis of 1-azanapthalene-8-Ol. Acta Phys Pol A. 2014;125(1):18–22.10.12693/APhysPolA.125.18Search in Google Scholar
[66] Abbaz T, Bendjeddou A, Villemin D. Molecular structure, HOMO, LUMO, MEP, natural bond orbital analysis of benzo and anthraquinodimethane derivatives. Pharm Biol Eval. 2018;5(2):27.10.26510/2394-0859.pbe.2018.04Search in Google Scholar
[67] Buglak AA, Kononov AI. Silver cluster interactions with tyrosine: Towards amino acid detection. Int J Mol Sci. 2022;23(2):634.10.3390/ijms23020634Search in Google Scholar PubMed PubMed Central
[68] Weinhold F, Landis CR. Natural bond orbitals and extensions of localized bonding concepts. Chem Educ Res Pract. 2001;2(2):91–104.10.1039/B1RP90011KSearch in Google Scholar
[69] Tatzber F, Rabl H, Koriska K, Erhart U, Puhl H, Waeg G, et al. Elevated serum neopterin levels in atherosclerosis. Atherosclerosis. 1991;89(2–3):203–8.10.1016/0021-9150(91)90061-7Search in Google Scholar PubMed
[70] Trausch JJ, Ceres P, Reyes FE, Batey RT. The structure of a tetrahydrofolate-sensing riboswitch reveals two ligand binding sites in a single aptamer. Structure. 2011;19(10):1413–23.10.1016/j.str.2011.06.019Search in Google Scholar PubMed PubMed Central
© 2025 the author(s), published by De Gruyter
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Research Articles
- Neopterin interactions with magic atom number coinage metal nanoclusters: A theoretical study
- High expression of folate metabolic pathway gene MTHFD2 is related to the poor prognosis of patients and may apply as a potential new target for therapy of NSCLC
- Changes and imbalance of Th1 and Th2 immune response in pediatric patients with seasonal allergic conjunctivitis
- Extracellular spermidine attenuates tryptophan breakdown in mitogen-stimulated peripheral human mononuclear blood cells
- Plasma total neopterin and neopterin levels are significantly elevated in stroke patients before carotid endarterectomy surgery
Articles in the same Issue
- Research Articles
- Neopterin interactions with magic atom number coinage metal nanoclusters: A theoretical study
- High expression of folate metabolic pathway gene MTHFD2 is related to the poor prognosis of patients and may apply as a potential new target for therapy of NSCLC
- Changes and imbalance of Th1 and Th2 immune response in pediatric patients with seasonal allergic conjunctivitis
- Extracellular spermidine attenuates tryptophan breakdown in mitogen-stimulated peripheral human mononuclear blood cells
- Plasma total neopterin and neopterin levels are significantly elevated in stroke patients before carotid endarterectomy surgery