You are on page 1of 12

Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 270 (2022) 120816

Contents lists available at ScienceDirect

Spectrochimica Acta Part A: Molecular and


Biomolecular Spectroscopy
journal homepage: www.elsevier.com/locate/saa

A copper nanoclusters probe for dual detection of microalbumin and


creatinine
Supitcha Thammajinno a,b,c, Chittanon Buranachai a,b,d, Proespichaya Kanatharana a,b,c,
Panote Thavarungkul a,b,c,d, Chongdee Thammakhet-Buranachai a,b,c,⇑
a
Division of Physical Science, Faculty of Science, Prince of Songkla University, Hat Yai, Songkhla 90110, Thailand
b
Center of Excellence for Trace Analysis and Biosensor, Prince of Songkla University, Hat Yai, Songkhla 90110, Thailand
c
Center of Excellence for Innovation in Chemistry, Faculty of Science, Prince of Songkla University, Hat Yai, Songkhla 90110, Thailand
d
Thailand Center of Excellence in Physics, Commission on Higher Education, 328 Si Ayutthaya Road, Bangkok 10400, Thailand

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 Glutathione capped copper


nanoclusters probe was synthesized
with ease.
 The single probe is capable of dual
detection to obtain albumin to
creatinine ratio.
 The probe shows excellent
performance in real-urine analysis.

a r t i c l e i n f o a b s t r a c t

Article history: A fluorescent probe based on glutathione-capped copper nanoclusters (GSH-CuNCs) was developed for
Received 20 September 2021 the detection of dual targets, human serum albumin (HSA) and creatinine, in human urine. The GSH-
Received in revised form 21 December 2021 CuNCs were synthesized by a one-pot green method using ascorbic acid as a reducing agent. The detec-
Accepted 23 December 2021
tion of HSA was in a turn-on mode via electrostatic interaction in a basic condition while the detection of
Available online 29 December 2021
creatinine was in a turn-off mode via non-covalent bonding in an acidic condition. Under optimal condi-
tions, the linear range and detection limit of HSA were 5.0 nM to 150 nM and 1.510 ± 0.041 nM, while
Keywords:
those of creatinine were 30 lM to 1000 lM and 13.0 ± 1.0 lM. This easily fabricated nanocluster probe
Glutathione
Copper nanoclusters
provided a fast response with high sensitivity, and good selectivity. Recoveries from urine samples were
Human serum albumin in the range of 81.44 ± 0.25 to 109.22 ± 0.57% for HSA and 80.57 ± 0.16 to 109.0 ± 0.10% for creatinine. The
Creatinine urinary analytical results from the fluorescent probe were in good agreement (P > 0.05) to those obtained
Albumin to creatinine ratio from immunoturbidimetric and enzymatic methods, signifying the excellent performance of this sensing
Fluorescence detection system.
Ó 2021 Elsevier B.V. All rights reserved.

1. Introduction applications involve detecting more than one type of target ana-
lyte, which requires more than one sensor construct to handle. A
A common practice in biosensor and chemical sensor develop- well-known example is human serum albumin (HSA) and crea-
ment is to focus on detecting one analyte with the highest sensitiv- tinine detection in urine samples to detect early kidney damage
ity, good selectivity and the lowest detection limit. However, some [1]. Urinary HSA, or microalbumin, is an indicator of kidney dis-
eases [2]. However, the amount of secreted microalbumin also
depends on food uptake [3] and this complicates the diagnosis. A
⇑ Corresponding author at: Division of Physical Science, Faculty of Science, Prince
solution to this problem is to measure creatinine concomitantly
of Songkla University, Hat Yai, Songkhla 90110, Thailand.
with microalbumin [4]. Creatinine is another biomarker for renal
E-mail address: chongdee.t@psu.ac.th (C. Thammakhet-Buranachai).

https://doi.org/10.1016/j.saa.2021.120816
1386-1425/Ó 2021 Elsevier B.V. All rights reserved.
S. Thammajinno, C. Buranachai, P. Kanatharana et al. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 270 (2022) 120816

dysfunction. It is the final product of creatine metabolism in skele- with HSA via electrostatic force inducing an aggregation of the
tal muscles [5] and is excreted by kidney at a relatively constant clusters and the enhancement of fluorescence intensity (‘‘turn-
rate [6] in healthy people independent of food uptake [7]; there- on”) as seen in some prior works [34,35]. The phenomenon of
fore, biomarkers in urine are often measured against the concen- aggregation-induced emission was occurred under certain condi-
tration of creatinine. Renal dysfunction gives rise to a reduction tions such as solvent, ions, pH, and special substances and leads
in creatinine excretion rate [8]. Accordingly, measuring the con- to aggregation of CuNCs with a strong fluorescence because of
centrations of microalbumin and creatinine concomitantly as albu- the intramolecular rotation and vibration of NCs are inhibited.
min creatinine ratio (ACR) provides a more reliable diagnosis for Therefore, the non-radiative pathways were blocked, and mean-
kidney malfunction. For example, urine ACR has been widely used while, the radiative decay was activated in the molecular aggregate
as an indicator to identify kidney disease that can occur as a com- state, resulting by increase in fluorescence intensity of NCs [36]. On
plication of diabetes [4] or chronic kidney disease [9]. the other hand, at an acidic pH, creatinine is neutral and could
Microalbumin and creatinine can be measured by various con- interact with GSH-CuNCs via non-covalent bonding [37]. The com-
ventional analytical methods, such as high-performance liquid plex formation could then quench CuNCs fluorescence (‘‘turn-off”)
chromatography (HPLC) [10,11] and liquid chromatography (LC) because of, e.g. electron transfer between creatinine and nanoclus-
[12,13]. These methods provide high selectivity but they need high ters [38].
cost of equipment, require well-trained professionals with involve We synthesized the GSH-CuNCs and characterized the interac-
several environmentally toxic solvents. In clinical laboratories gen- tions between the nanoclusters and the two target analytes (HSA
erally use immunoassays, among which most are immunoturbidi- and creatinine). Afterwards, the affecting parameters of the pro-
metric assays (ITA) for HSA detection [14] and enzymatic method posed sensing method were optimized to obtain the highest sensi-
applied with creatinine detection [15]. These methods required tivity and specificity before validating the sensing performance
the labeling of antibody-based assays and enzymes. Although the and applying for the determination of the two target-analytes in
introduction of antibody and enzymes have improved the sensitiv- real urine samples. The results were also compared with those
ity and selectivity, but it is a high cost and instable of enzymes obtained from conventional methods used by the hospital.
[2,16]. Among them, the electrochemical and optical sensors are
considered the two most popular types. Electrochemical sensors
2. Experimental section
for microalbumin and creatinine detections are highly sensitive
with extra low detection limit and exceptional selectivity [17,18],
2.1. Materials
however, they are prone to electrical interference and often require
complicate electrode preparation to avoid interference. Optical
Chemicals were obtained from manufacturers as follows. Cop-
sensors offer comparable sensitivity and selectivity but are less
per (ΙΙ) chloride dehydrate (CuCl22H2O) was from Merck (Darm-
susceptible to external interference and are easier to construct
stadt, Germany). L-glutathione (GSH) and L-cysteine were from
when compared with the electrochemical type [19,20]. Neverthe-
Sigma-Aldrich (Tokyo, Japan). L-ascorbic acid, dopamine
less, most of the existing sensors is still capable of detecting only
hydrochloride, creatinine and human serum albumin were from
either microalbumin or creatinine but rarely both. In this work
Sigma-Aldrich (Darmstadt, Germany). Sodium dihydrogen phos-
we address the problem by developing a single fluorescence sensor
phate dihydrate (NaH2PO42H2O) and sodium phosphate dibasic
for a dual detection of microalbumin and creatinine based on cop-
dodecahydrate (Na2HPO412H2O) were from Ajax Finechem (Auck-
per nanoclusters.
land, New Zealand). Sodium acetate trihydrate (C2H3NaO23H2O)
Metallic nanoclusters (NCs) contain a few to several hundred
was from AnalaR NORMAPUR (Haasrode, Belgium). Sodium chlo-
metal atoms [21]. Therefore, the electrons are well confined and
ride (NaCl) was from Merck (Copenhagen, Denmark). Uric acid
the clusters become fluorescent [22]. Similar to quantum dots,
was from Sigma-Aldrich (Budapest, Hungary) and urea was from
metallic NCs have high extinction coefficient and are highly resis-
Riedel-de Haën (Bucharest, Romania). All aqueous solutions were
tant to photobleaching but they are less toxic and easier to synthe-
prepared with deionized water (18.2 MXcm resistivity) from a
size [23]. Among various types, copper NC is interesting because it
water purification system (Milli-Q water Direct, Millipore Corp.,
is relatively inexpensive and easy to synthesize [24,25] and after
Billerica, MA, USA).
stabilized with proper capping agents [21,26,27], the interaction
of CuNCs with certain target analytes can either increase or
decrease their fluorescence intensity [28,29]. For example, a prior 2.2. Preparation of GSH-CuNCs
publication reports that xanthine (XA) can bind to the surface of
glutathione-stabilized copper nanoclusters (GSH-CuNCs) via NACu The preparation procedure was adapted from the method of
bond and increase the surface electron density resulting in an Huang and co-workers [39]. Briefly, 100 mL of 0.20 mM GSH solu-
enhancement of fluorescence intensity (turn-on mode) [30], while tion were mixed under vigorous stirring with 312 lL of a 0.050 M
rifampicin (RFP), after binding, quenches the fluorescence of solution of CuCl2 in DI water. The colorless solution immediately
glutathione-stabilized copper nanoclusters (GSH-CuNCs) via the changed to a white emulsion. Then, 2.25 mL of 0.10 mM ascorbic
inner-filter effect (turn-off mode) [31]. Therefore, it is possible to acid were added dropwise to the white emulsion, which became
engineer a CuNCs probe using a proper capping agent, such as a clear solution after 30 min. The clear solution was continuously
GSH, under suitable sensing conditions to detect more than one stirred for 4 h at 65 °C to obtain a yellow solution of CuNCs, which
target analyte using different detection modes (turn-on vs turn- was stored in darkness at 4 °C without further purification. To ver-
off) without the complications of using separate detection methods ify that the synthesis was successful, the UV–vis absorption spectra
or different sensor development. of the nanocluster solution was measured by an absorption spec-
In this work, we report the development of a single GSH-CuNCs trometer (Avaspec 2048, Apeldoorn, The Netherlands).
fluorescent probe for the detection of dual target analytes, microal-
bumin and creatinine, in order to obtain the ACR values. Our 2.3. Characterization of CuNCs
hypothesis was based on a fact that at a pH in the range of 7.0–
8.0, the GSH is positively charged (pKa of amine group on The morphology and size of CuNCs were studied under trans-
GSH = 9.49) [32] while HSA is negatively charged (isoelectric mission electron microscopy (TEM) (JEOL, JEM 2010 system, Tokyo,
points (pI) of HSA = 4.7) [33]. Therefore, GSH-CuNCs could interact Japan) operating at 200 kV. The specimens were prepared by
2
S. Thammajinno, C. Buranachai, P. Kanatharana et al. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 270 (2022) 120816

dropping 3.50 lL of the CuNCs solution onto a carbon coated cop- tions in the absence and presence of creatinine were measured
per grid and then dried at room temperature prior to imaging. In under a time-resolved fluorometer using a 370 nm pulsed laser
addition, a portion of the nanocluster solution was lyophilized as the excitation light source (FLS 980, Edinburgh Instruments, Liv-
and the solid sample was dispersed in KBr discs prior to character- ingston, United Kingdom).
izing the chemical composition using a Fourier transform infrared
(FT-IR) spectrometer (VERTEX 70, Bruker, Germany). Moreover, the 2.5. Optimization of operational conditions
zeta potential of the GSH-CuNCs in the presence and absence of
HSA and creatinine were evaluated by a particle analyzer (Zeta- Optimization was performed by changing a single parameter at
PALS system, Brookhaven Instruments, New York, USA) to verify a time while other parameters were kept constant. Standard solu-
the surface charge. tions of HSA (5.0–150 nM) and creatinine (30–1000 lM) were pre-
To confirm the success of the synthesis, cyclic voltammetry (CV) pared and analyzed (three replicates for each tested condition). The
was used to verify that Cu2+ were reduced to Cu+ and Cu0 to pro- F/F0 and F0/F of HSA and creatinine standards were plotted against
duce CuNCs. The CV was performed using an lAUTOLAB Type III their respective concentrations. The optimal value was the one that
potentiostat galvanostat (Metrohm Autolab B.V., Utrecht, The provided the steepest slope (sensitivity) of the calibration plot, the
Netherlands) controlled by a Nova 2.1.1 software using a standard shortest analysis time and the lowest limit of detection (LOD) and
three-electrode cell at room temperature. A glassy carbon elec- limit of quantification (LOQ).
trode (3.0 mm in diameter) as a working electrode was first pol-
ished with 1.0 lm and then 0.30 lm alumina slurries before
2.6. Analysis of real samples
being sonicated successively in ethanol and water. A platinum wire
as a counter electrode and Ag/AgCl (in saturated KCl) as a reference
Urine samples were kindly provided by Songklanagarind Hospi-
electrode were dipped in 0.10 M NaOH as an electrolyte.
tal (Hat Yai, Songkhla, Thailand). The samples collecting and han-
In the part of modification on glassy carbon electrode (GCE),
dling protocol met the criteria of the Exemption Determination
chitosan cryogel (prepared by freezing and thawing of chitosan
as was acknowledged by Human Research Ethics Committee of
[40]) was used to entrap the CuNCs. Briefly, 6.0 mg of the lyophi-
Faculty of Medicine, Prince of Songkla University (REC.63-510-
lized CuNCs was added into 500 lL of chitosan solution (2.0% w/
19-2). They were diluted with an appropriate dilution factor for
v) in an amber glass bottle (4.0 mL). The mixture was stirred over-
HSA and creatinine detection using 0.15 M phosphate buffer at
night at room temperature. Next, 25.0 lL of this mixture was com-
pH 7.4 and 0.10 M acetate buffer at pH 5.0, respectively, then ana-
bined with 1.0 lL of 2.5% (v/v) of glutaraldehyde, used as the cross-
lyzed under the optimal conditions. The obtained concentrations of
linking reagent. Five lL of this mixture was then drop-casted onto
HSA and creatinine were statistically compared by the Wilcoxon
a glassy carbon electrode and stored at 20 °C overnight to achieve
signed rank test to those obtained from the immunoturbidimetric
cryogelation and then thawed at 4.0 °C for 30 min to obtain a
and enzymatic standard methods used by the hospital.
CuNCs-CS cryogel/GCE [41,42]. In addition, a chitosan cryogel
modified glassy carbon electrode without CuNCs (CS cryogel/
GCE) was also prepared for comparison with the CuNCs-CS cryo- 3. Results and discussion
gel/GCE.
3.1. Characterization of GSH-CuNCs
2.4. Steady state and lifetime-resolved fluorescence measurement
The UV–vis absorption spectrum of GSH in DI water (Fig. 1b)
All steady state fluorescence measurements were done in a showed no absorption peak while the spectrum of GSH-CuNCs in
spectrofluorometer (RF-5301, Shimadzu, Tokyo, Japan). For HSA DI water showed a sharp peak at 292 nm that indicated the suc-
detection, the GSH-CuNCs stock solution was diluted and mixed cessful synthesis of GSH-CuNCs. The absence of a plasmon band
with HSA standard stock solution at a desired concentration in of Cu nanoparticles at 507 nm in this spectrum indicated that the
phosphate buffer at a basic pH. The dilution was done with a factor synthesized GSH-CuNCs were of a high purity [25,39]. Under UV
that kept the fluorescence intensity of the blank (i.e. no HSA) low light at 365 nm, the as-prepared GSH-CuNCs solution emitted an
(70 a.u. out of the maximum intensity limit of the instrument intense blue fluorescence (Fig. 1b cuvette I) which was compared
at 1000 a.u.), because HSA detection was done in a turn-on mode; with the non-fluorescence of DI water (cuvette II). When the exci-
the fluorescence intensity increases with increasing the HSA con- tation wavelength was varied from 300 to 370 nm at 10 nm incre-
centration. For creatinine detection, the GSH-CuNCs stock solution ments, the successive peaks of fluorescence emission occurred at
was diluted with acetate buffer and mixed with creatinine stan- the same wavelength of 428 nm. The maximum intensity of fluo-
dard stock solution at a desired concentration at an acidic pH with rescence emissions occurred at the excitation wavelength of
a different factor from that used in HSA detection to ensure that 360 nm (Fig. 1c) and this was used as the excitation wavelength
the initial fluorescence intensity from blank solution (i.e. no crea- for the remaining experiments.
tinine) was high (165 a.u.). This is because creatinine detection The molecular structures of GSH and GSH-CuNCs were charac-
was done in a turn-off mode. Then, the mixture was transferred terized by FT-IR spectroscopy. In the FTIR spectrum of GSH
into a quartz microcuvette for fluorescence measurement (Starna (Fig. 1d), absorption peaks of the ANH2 group were observed at
Scientific Ltd., Essex, U.K.) and its fluorescence spectra was 3348 and 3250 cm 1 for NAH, and 3026 cm 1 for CAH stretching
recorded with the excitation wavelength at 360 nm and the emis- vibration bands. Peaks around 3000 cm 1 were attributed to the
sion wavelengths ranging from 380 to 700 nm and slid widths of absorption bands of OAH stretching. Characteristic peaks were also
5 nm for excitation and 10 nm for emission. Finally, the integrated presented at 2525 cm 1 for the SAH stretching vibration of the
fluorescence intensity of GSH-CuNCs in the presence (F) and ASH group while the peak at 1715 cm 1 was attributed to the
absence (F0) of the target analytes (either HSA or creatinine) were C@O stretching vibration of the ACOOH group. The sharp NAH
calculated. F/F0 was used as the signal for HSA detection, whereas stretching bands, present in the spectrum of GSH at 3348 and
F0/F was used as the signal for creatinine detection. 3250 cm 1, were not presented in the spectrum of GSH-CuNCs.
To clarify the quenching mechanism, the fluorescence lifetimes Instead, a wide band presented at about 3406 cm 1, which should
was employed to explore the mechanism of the fluorescence be attributed to the absorption of the associated hydroxyl (AOH)
response of GSH-CuNCs towards creatinine. The GSH-CuNCs solu- and amino (ANH2) groups. The SAH stretching peak of GSH at
3
S. Thammajinno, C. Buranachai, P. Kanatharana et al. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 270 (2022) 120816

Fig. 1. (a) UV–vis absorption spectra of GSH and GSH-CuNCs were produced at wavelengths from 200 to 700 nm: (b) Photographs show the fluorescence of (I) CuNCs in
solution compared to (II) DI water under UV light at 365 nm: (c) Fluorescence emission spectra of CuNCs were produced at excitation wavelengths from 300 to 370 nm:
(d) FT-IR spectra of GSH and GSH-capped CuNCs.

2525 cm 1 was also absent in the spectrum of GSH-CuNCs. These To confirm this possible mechanism, TEM image and zeta
results indicated the attachment of GSH on the surface of CuNCs potential were obtained. TEM images showed spherical GSH-
via the amine and thiol groups [43]. CuNCs with an average diameter of 1.70 ± 0.66 nm (n = 100) (Sup-
To further characterize CuNCs, cyclic voltammetry was carried plementary information, Fig. S2a). Aggregation of GSH-CuNCs in
out to study their electrochemical behavior. Owing to the small the presence of HSA had an average diameter of 62 ± 37 nm
size of CuNCs, chitosan cryogel (CS-cryogel), which can be very (n = 55) (Supplementary information, Fig. S2b). The size of GSH-
easily made by freezing and thawing of chitosan was employed CuNCs aggregations in the presence of HSA was similar to those
to entrap these nanomaterials [40,44]. reported in earlier research [35], confirming that HSA induces the
The electrochemical tests were carried out in 0.10 M NaOH nanoclusters to form aggregations, resulting in a change in the size
aqueous solution at a scanning rate of 10 mV s 1. The forward scan of the nanoclusters.
range was from the potential of 1.20 to + 0.80 V, followed by the Besides TEM images, zeta potential was used further to confirm
reverse scan back to the onset potential ( 1.20 V). The CV of the interaction between GSH-CuNCs and HSA. The zeta potentials
CuNCs-CS cryogel/GCE exhibited characteristic anodic current of GSH-CuNCs were measured before and after the addition of
peaks at 0.36 and 0.048 V, corresponding to Cu (0)/Cu (I) and HSA. The GSH-CuNCs provided a zeta potential value of
Cu (I)/Cu (II), respectively (Supplementary information, Fig. S1). 25.1 mV, indicated that the negative charge was exposed on the
The cathodic peaks for Cu (II)/Cu (I) and Cu (I)/Cu (0) is at 0.37 surfaces of the CuNCs. The addition of 50, 100 and 150 nM HSA
and 0.74 V, respectively. These peaks were not observed with changed the value to 23.78 mV, 15.36 mV and 9.81 mV,
the CS cryogel/GCE. These CVs further demonstrated the successful respectively. The decrease of the negatively charge value with
preparation of CuNCs, and confirmed that the synthesized CuNCs higher concentration of HSA indicating that the negatively charge
had the same properties as previous works [44,45]. surface of GSH-CuNCs was neutralized and the interaction between
HSA and GSH-CuNCs occurred most likely via electrostatic force,
3.2. Sensing mechanism for HSA and creatinine detection which induce the nanoclusters to come closer together.
In an acidic condition at pH 5.0, creatinine is electrostatically
The possible detection mechanism of HSA and creatinine by the neutral and could interact with GSH-CuNCs probe via non-
GSH-CuNCs is proposed in Fig. 2. In the presence of HSA, the iso- covalent bonding to form non-fluorescent complexes, resulting in
electric point (pI) of HSA is 4.7 [33]. At pH 7.4, that is higher than electron transfer from creatinine to the nanoclusters, causing a
the pI, HSA is negatively charged. Since the amine group of GSH- reduction in fluorescence emission as a ‘‘turn-off” mode (Fig. 2b)
CuNCs is positively charged [32], HSA induces the GSH-CuNCs to [37,38]. Fluorescence lifetime time measurement was employed
come closer and interact via electrostatic attraction, which affects to explore the possible mechanism of the GSH-CuNCs probe for
the aggregate of nanoclusters. The aggregation increases fluores- the detection of creatinine. From fitting the fluorescence decay
cence intensity and the GSH-CuNCs are working in the turn-on curves, the average fluorescence lifetimes of the GSH-CuNCs in
mode (Fig. 2a). the absence and presence of creatinine were found to be
4
S. Thammajinno, C. Buranachai, P. Kanatharana et al. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 270 (2022) 120816

Fig. 2. Sensing mechanism in the presence of (a) human serum albumin (HSA) at pH 7.4 induces the electrostatic interaction between HSA and GSH-CuNCs following the
aggregation-induced emission enhancement (turn-on mode). Inset: a fluorescence spectra of GSH-CuNCs with concentration of HSA increase the FL intensity also increase
and (b) the detection of creatinine at pH 5.0 was interacted via non-covalent bonding with GSH-CuNCs based on static quenching via formation of ground state complex
(turn-off mode). Inset: a fluorescence spectra of GSH-CuNCs with concentration of creatinine increase the FL intensity also decrease.

1.858 ± 0.059 ns and 1.833 ± 0.063 ns, respectively (Supplemen- Interestingly, Fig. 3B shows a downward deviation from a pure
tary information, Fig. S3). The unchanged average fluorescence life- linear Stern-Volmer plot. This is unlikely due to the presence of
time after the addition of creatinine indicates that creatinine may both static and dynamic quenching because, in that case, the devi-
quench the fluorescence of the GSH-CuNCs via a static (non- ation would have been upward [47]. Instead, this downward
fluorescent complex formation) quenching mechanism. The slight behavior was reported in the case of buried fluorophores inacces-
increase in the average diameter of GSH-CuNCs from sible to the quencher, such as tryptophan’s fluorescence quenching
1.70 ± 0.66 nm to 2.22 ± 0.70 nm (n = 85) in the presence of crea- by trifluoroacetamide [48]. We then hypothesize that some of our
tinine observed via the TEM images (Supplementary data, CuNCs are easily accessed and statically quenched by creatinine
Fig. S2c) was likely due to the formation of ground state complex whereas the others are not due to heterogenous surface function-
between creatinine molecules and copper atoms [46]. alization. This gave rise to the initial decrease in the nanocluster’s

5
S. Thammajinno, C. Buranachai, P. Kanatharana et al. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 270 (2022) 120816

fluorescence intensity with increasing creatinine concentration. At a good buffer capacity that could cover the range of pH from
the concentration higher than 1000 lM, however, all nanoclusters 4.0 to 8.0 [49], two common buffers were applied together:
that are easily accessed are all quenched while the rest are still flu- phosphate buffer in a pH range from 5.80 to 8.0 for HSA detec-
orescent even when the concentration of the quencher is raised tion and acetate buffer in a pH range from 3.40 to 5.90 for
further. Despite this heterogeneity, we are confident in the repro- creatinine detection.
ducibility of the synthesis and as far as creatinine sensing is con- The effect of basic pH was investigated for HSA at 7.0, 7.2, 7.4,
cerned, we are using only the linear portion of the plot.” 7.6 and 8.0. The sensitivity of the response was not much different
On the other hand, zeta potential of GSH-CuNCs in acetate buf- at each pH (Supplementary information, Fig. S5a). This was possi-
fer pH 5.0 exhibited a value of 15.05 mV, suggested that there is bly because the isoelectric point (pI) of HSA is 4.7 [33] and within
an electrostatic repulsion among the GSH-CuNCs helping to stabi- the applied pH range, HSA remained negatively charged and could
lize CuNCs from aggregations. Upon the addition of 30, 50 and always interact with GSH via electrostatic bonding. Since the sen-
300 lM creatinine, the apparent zeta potential of the CuNCs sitivity at pH 7.4 was slightly higher than at other pH values, this
decreased from 15.05 to 12.51 mV, 12.51 to 10.50 mV and pH was used for further studies.
10.50 to 8.77 mV, respectively. The reduction in the zeta poten- Creatinine detection was tested in acetate buffer at pH 4.0, 4.5
tial value in the presence of creatinine suggested that surface and 5.0, and in phosphate buffer at pH 6.0 and 7.0 (Supplementary
charge neutralization involved between GSH-CuNCs and creatinine information, Fig. S5b). Detection sensitivity increased from pH 4.0
via non-covalent bonding and this led to the formation of non- to 5.0 and decreased at higher pH. This was likely because at pH
fluorescent complexes between the creatinine and GSH-CuNCs 5.0, creatinine (pKa  5.02) [50] remains in a neutral form, and
probe [38]. It is unlikely that creatinine causes a destruction of GSH which contains amide, carboxylic acid and amino groups
the nanocluster that results in the decrease in fluorescence inten- [32] can induce creatinine towards the CuNCs via hydrogen bond-
sity because the extinction peaks of the nanocluster do not signif- ing, enhancing the likelihood of interaction between creatinine and
icantly change in the presence of various concentrations of copper atoms on the CuNCs surface. At pH higher than 5.0, crea-
creatinine (Supplementary information, Fig. S4). Therefore, an effi- tinine is in a negatively charged form, and repulsion will occur
cient electron transfer between creatinine and nanoclusters was between GSH and creatinine, lowering the possibility of the inter-
occurred in acetate buffer pH 5.0 which quenches the fluorescence action. So, acetate buffer at pH 5.0 was used in the remaining
emission. studies.
The concentration of phosphate buffer at pH 7.4 was optimized
3.3. Optimization of conditions for HSA and creatinine detection at 0.050, 0.10, 0.15, 0.20 and 0.25 M while the acetate buffer at pH
5.0 was optimized at 0.010, 0.050, 0.10 and 0.15 M. Sensitivity
3.3.1. pH and concentration of buffer toward HSA increased in the phosphate buffer from 0.050 to
pH was the most important parameter in this study because 0.15 M (Supplementary information, Fig. S6a) and sensitivity
it directly affected the binding between GSH-CuNCs and both toward creatinine increased in the acetate buffer from 0.010 to
target analytes. For the proposed sensing mechanism to work, 0.10 M (Supplementary information, Fig. S6b). In both cases, sensi-
the moieties in the reactions required the appropriate basic tivity decreased at higher concentrations. It is likely that at too
and acidic buffer ranges. Since there was no single buffer with high a concentration, ions might block access to the GSH-CuNCs

Fig. 3. Fluorescent spectra and calibration plots were produced from (a) HSA and (b) creatinine detection under the optimum 15 min reaction time. Calibration plots mapped
F/F0 for HSA (turn-on mode), and F0/F for creatinine (turn-off mode) against concentration at an excitation wavelength of 360 nm. Insets show zooms of the fluorescence
spectra peaks of GSH-CuNCs in the presence of different concentrations of the analytes.

6
S. Thammajinno, C. Buranachai, P. Kanatharana et al. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 270 (2022) 120816

surface, making difficult paths for the targets to interact [51]. 3.4.3. Precision
Therefore, 0.15 M phosphate buffer at pH 7.4 and 0.10 M acetate The precision of the proposed sensor was evaluated by spiking
buffer at pH 5.0 were the optimal buffer conditions for HSA and standard solutions of HSA and creatinine in urine samples obtained
creatinine detection. from Songklanagarind Hospital. The final concentrations of HSA in
the spiked urine samples were 5.0, 10.0, 50.0, 100.0 and 150.0 nM,
3.3.2. Reaction time whereas those of creatinine were 30.0, 70.0, 300.0, 500.0 and
The effect of reaction time on HSA and creatinine detection was 1000.0 lM. Six replications of each spiked sample were analyzed
evaluated at 10, 15, 20 and 25 min. For both targets, detection sen- under the optimal conditions. The relative standard deviations (%
sitivities increased from 10 to 15 min and became relatively con- RSD) varied between 0.10 and 0.60 for the determination of HSA
stant thereafter (Supplementary information, Fig. S7a and S7b). and between 0.12 and 0.50 for the determination of creatinine.
Sensitivity was higher at 15 min, so this was the time chosen as These values were better than the AOAC guideline RSDs, which
the optimal value. are 7.3% at the concentration of 10 mg L 1 and 5.3% at the concen-
tration of 100 mg L 1 for HSA and creatinine, respectively [52].
Therefore, the analytical precision achieved with the developed
3.4. Analytical performances sensor was good for HSA and creatinine detection.

3.4.1. Linear dynamic range, limit of detection and limit of 3.4.4. Stability
quantification The stability of the GSH-CuNCs was tested to determine how
Under optimal conditions, the system was tested in the turn-on long the fluorescent sensing probe remained usable after it was
mode, detecting HSA concentrations from 1.0 to 300 nM and in the synthesized. Calibration curves were plotted using freshly pre-
turn-off mode, detecting creatinine concentrations from 10 to pared GSH-CuNCs to determine HSA and creatinine in their linear
3000 lM. The response to HSA concentration was linear in the ranges (up to 7 concentrations, n = 3) and the sensitivities obtained
range of 5.0 nM to 150 nM (Fig. 3a) and to creatinine concentration were set as 100% at week zero. The GSH-CuNCs solution was kept
in the range of 30 lM to 1000 lM (Fig. 3b). The limits of detection in darkness at 4 °C and used to detect HSA and creatinine in the
(LODs) and limits of quantification (LOQs) were calculated based same concentration ranges on a weekly basis. During the first
on LOD = 3Sa/b and LOQ = 10Sa/b, where Sa is the standard devia- 14 weeks, relative sensitivity remained stable at an average of
tion of the intercept and b is the slope of the calibration curve. 99.64 ± 0.67% for HSA and 99.84 ± 0.50% for creatinine. Sensitivity
The LODs were 1.510 ± 0.041 nM for HSA and 13.0 ± 1.0 lM for cre- then slowly declined and dropped below 90% after 20 weeks (Sup-
atinine while the LOQs were 5.04 ± 0.14 nM for HSA and plementary information, Fig. S9a and 9b). The probes were there-
44.0 ± 3.0 lM for creatinine. fore considered stable for up to 20 weeks [53].
It is reasonable to argue that the fluorescence enhancement
and quenching are not very obvious. So, another commercial 3.4.5. Selectivity of the sensor
spectrofluorometer (LS-55, PerkinElmer, Waltham, MA, USA) Human urine may present many interferences to the detection
was used to test and confirm the fluorescence enhancement of the target analytes. Selectivity was therefore tested toward rel-
and quenching of the reaction between GSH-CuNCs probe and evant biological molecules and salts at the highest level normally
the two targets. The fluorescence emission spectra were recorded found in human urine. The molecules tested and their levels were
from 380 nm to 700 nm with the excitation wavelength of uric acid at 4.46 mM [54], ascorbic acid at 3.40 mM [55], cysteine
360 nm, as has been used with the first spectrofluorometer (RF- at 2.50 mM [56], urea at 500 mM [57], Na+ at 260 mM [58] and
5301, Shimadzu, Tokyo, Japan). As expected, under the same opti- Cl at 170 mM [59]. Dopamine was tested at a higher than normal
mal condition for HSA detection, the probe’s fluorescence signal level, at 2.50 mM, because its level in human urine may depend on
increases with increasing concentration of HSA; likewise, under age [60]. Also, HSA was tested at a higher than normal level, at
the optimal condition for creatinine detection, the probe’s fluores- 0.0040 mM, [61] as an interference to creatinine detection and cre-
cence signal decreases with increasing creatinine concentration atinine was tested as an interference to HSA detection at 15.0 and
(Fig. S8 and Table S1 in supplementary information). This agrees 30.0 mM, its middle and maximum levels in human urine [62]. The
with those observed under the first fluorometer. There are slight concentrations to be determined in this study were from the mid-
differences between the two data sets taken from the two fluo- dle of the linear range of the calibration plots, at 0.10 lM for HSA
rometers, such as sensitivities of creatinine detection, but we and 0.50 mM for creatinine.
believe that those values are instrument dependent that have to Taking account, the level of the analytes of interest in real sam-
be considered when characterizing a sensor. Therefore, we can ples, detection should be carried out in a 6-time dilution of HSA
conclude that the increase and decrease of the probe’s fluores- and 40-time dilution of creatinine so that their concentrations
cence signal really come from the interaction between GSH- would fall within the calibration ranges. Therefore, the interfer-
CuNCs probe and HSA and creatinine, respectively, regardless of ences were diluted with buffer accordingly.
the fluorometers being used, providing that the detections are During HSA detection (Fig. 4a), negligible increases in the fluo-
done under the optimal conditions. rescence intensity of the biological molecules were observed:
dopamine 0.31%, Na+ 0.41%, Cl 0.32%, ascorbic acid 0.45%, uric acid
3.4.2. Reproducibility 0.68%, 15 mM creatinine 0.66% and 30.0 mM creatinine 0.98%.
The reproducibility of the synthesized GSH-CuNCs as a fluores- These were well below the acceptable 5% change of signal [63].
cent probe for the determination of HSA and creatinine was evalu- During creatinine detection (Fig. 4b), no interference effects were
ated from the performances of six different preparations of the observed. All interferences produced the same signals as the blank
sensor. Up to seven concentrations of HSA and creatinine in their solution (GSH-CuNCs in buffer), indicating that the relevant mole-
respective linear ranges (each concentration n = 3) were detected cules and salts had negligible effects on the detection of HSA and
and the sensitivities of the six preparations varied from creatinine.
(196.0 ± 1.4)  10 6 to (198.0 ± 2.0)  10 6 nM 1 (RSD = 0.41%) Each of the studied interferences was also mixed with HSA at
for HSA and from (96.0 ± 1.0)  10 6 to (98.0 ± 3.0)  10 6 lM 1 0.10 lM or creatinine at 0.50 mM. The fluorescence emission
(RSD = 0.65%) for creatinine. Thus, the synthesis and the detection intensities of all mixtures showed no significant change and were,
processes exhibited good reproducibility (P > 0.05). therefore, within the acceptable limit of signal change of less than
7
S. Thammajinno, C. Buranachai, P. Kanatharana et al. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 270 (2022) 120816

Fig. 4. The charts show the selectivity of the proposed sensor toward (a) HSA and (b) creatinine in the presence of other biological molecules and salts found in human urine
compared with a blank (GSH-CuNCs). Selectivity was also tested using solutions of (c) HSA at 0.10 lM and (d) creatinine at 0.50 mM mixed with interfering biological
molecules and salts.

5% (Fig. 4c and 4d). Thus, detection of HSA and creatinine with this samples and spiked buffer solutions were analyzed and the
method was highly selective. Even though HSA enhances while obtained integrated fluorescence intensities were plotted against
creatinine quenches the probe’s fluorescence, HSA does not inter- their respective concentrations. The slopes of the calibration plots
fere with creatinine detection because HSA does not increase the obtained from the spiked samples and spiked buffer solutions were
probe’s fluorescence under the optimum condition of creatinine compared by two-way ANOVA. The two slopes were significantly
detection (i.e., acidic solution). Likewise, creatinine does not inter- different at the confidence level of 95% (P < 0.05), indicating the
fere with HSA detection because creatinine does not quench the presence of a matrix effect. Therefore, the standard addition
probe’s fluorescence under the optimum condition of HSA detec- method was used to quantify the concentration of both HSA and
tion (i.e., slightly basic solution). creatinine in urine samples.
To detect these two compounds, the same sample was divided Fifteen samples were analyzed and the results were compared
into two portions; each of which was pipetted into a different tube to the results obtained from the immunoturbidimetric and enzy-
containing different buffer solution: 0.15 M phosphate buffer at pH matic methods used by the hospital. The results obtained from
7.4 for HSA and 0.10 M acetate buffer at pH 5.0 for creatinine. Then all urine samples with the developed fluorescence detection
the fluorescence signal from each reaction was measured sepa- method were in good agreement with the concentrations of target
rately before analyzing the concentration of the two analytes. Since analytes obtained from the conventional methods used by the hos-
the probe binds selectively to HSA or creatinine, depending on the pital (P > 0.05) (Fig. 5a, 5b and 5c).
conditions of the solutions, as shown in Fig. 4A-D, we are able to To further validate the developed fluorescence sensor, the accu-
detect both HSA and creatinine from the same samples with the racy of the method was also evaluated in terms of recoveries,
same probe. which were in the range of 81.44 ± 0.25–109.22 ± 0.57% for HSA
detection and 80.57 ± 0.16–109.0 ± 0.10% for creatinine detection
3.5. Determination of HSA and creatinine in human urine samples (Table S2). These values were within the acceptable limits between
80 and 110% (AOAC, 2011). Therefore, the GSH-CuNCs provided
Urine samples were diluted into the appropriate linear ranges good recovery, good precision, good accuracy, and could detect
before detection. Samples were diluted 6 or 20 times (depending HSA and creatinine in urine samples.
on their original fluorescence intensity) with 0.15 M phosphate The analytical performances of the GSH-CuNCs fluorescent
buffer at pH 7.4 for HSA determination and 40 times with 0.10 M probe were compared with performances reported for other sen-
acetate buffer at pH 5.0 for creatinine determination. sors for HSA and creatinine detection (Table 1). For HSA detection,
The matrix effect of urine samples was first studied by spiking the linear range, reaction time and LOD of the proposed method
with HSA and creatinine standard solutions to obtain final concen- were either comparable or better than the results of the reported
trations after dilution in the range of 5.0 nM to 150 nM and 30 to fluorescence and colorimetric methods [19,64]. Importantly, the
1000 lM, respectively. The different buffer solutions used for the LOD value (1.5 nM) was more than sufficient for the determina-
target analytes were also spiked in the same way. The spiked urine tion of urinary HSA (33.0–400 nM) [61]. In the case of creatinine
8
S. Thammajinno, C. Buranachai, P. Kanatharana et al. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 270 (2022) 120816

Fig. 5. The charts compare results from (a) HSA detection and (b) creatinine detection with GSH-CuNCs and standard methods and the chary (c) shows relative ACRs.

detection, the developed method provided a good linear range the dual targets, human serum albumin and creatinine in
with a comparable reaction time. Although, the LOD for creatinine human urine. The detection principle was based on electrostatic
in this work was higher than in other works [20,38,64,65], it was interaction between the nanoclusters and human serum albu-
good enough for the determination of creatinine in human urine min to turn on fluorescence and non-covalent bonding between
(3.6–27.0 mM in males and 3.30–22.50 mM in females) [62]. The the nanoclusters and creatinine to quench, or turn off, fluores-
great advantage of the developed sensor is probably the storage cence. Under optimal conditions, the developed method pro-
stability of 20 weeks. In addition, the preparation reproducibility vided good linear ranges and low limits of detection for both
of the GSH-CuNCs showed a good % RSD (n = 3). Values of % targets. Furthermore, when applied to the analysis of human
recovery in urine samples for HSA detection and creatinine were serum albumin and creatinine in human urine samples, the
also acceptable and comparable with the recoveries of other developed method produced highly accurate and precise analyt-
methods. ical results which were consistent with those obtained from the
immunoturbidimetric and enzymatic methods used by the col-
4. Conclusions laborating hospital. In addition, the fluorescence probe was also
highly selective toward human serum albumin and creatinine in
A fluorescent probe utilizing glutathione-capped copper nan- the presence of biological molecules and salts typically found in
oclusters was successfully developed for the determination of urine.

9
S. Thammajinno, C. Buranachai, P. Kanatharana et al. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 270 (2022) 120816

Table 1
Comparison of the performances of the GSH-CuNCs fluorescence sensor with the sensor performances of other fluorescence and colorimetric methods for HSA and creatinine
detection.

Sensing probe Method Linear range LOD Reaction Reproducibility Stability Sample Recovery Ref.
time (%)
HSA detection
(nM) (nM)
poly(thymin)-CuNPs Fluorescence 150–2500 82 15 min – – Serum 97.0–101.0 [19]
Ex 340 nm
Em 610 nm
CIZS QDs Fluorescence (0.075– 45 – – – Serum 88.14–94.39 [64]
Ex 380 nm 100)  103
Em 570 nm
citrate-AuNPs Colorimetric 20–100 1.4 15 min – – Artificial 96.7–103.8 [67]
A650/A520 urine
GSH-CuNCs Fluorescence 5.0–150 1.5 15 min RSD 0.41% 20 weeks Urine 81.44 ± 0.25– This
Ex 360 nm (n = 3) 109.22 ± 0.57 work
Em 428 nm
Creatinine detection
(lM) (lM)
AuQC@gluten Fluorescence 20–520 0.0020 15 min – – Blood 98.16–99.5 [65]
Ex 380 nm
Em 680 nm
BSA-CuNCs Fluorescence 5.0–60 0.050 12 min – >18 days Urine 97.63–102.43 [38]
Ex 525 nm
Em 643 nm
6
CC-AuNCs Fluorescence 0.00010–0.10 7.54  10 30 min RSD 4.2% >1 month Serum and 96.00–107.0 [20]
Ex 365 nm (n = 5) urine
Em 481 nm
TDA-AgNPs Colorimetric 0.010–1.0 0.0030 5 min – – Serum and 89.4–102.1 [66]
A560/A390 urine
GSH-CuNCs Fluorescence 30–1000 13.0 15 min RSD 0.65% 20 weeks Urine 80.57 ± 0.16– This
Ex 360 nm (n = 3) 109.0 ± 0.10 work
Em 428 nm

CRediT authorship contribution statement Compliance with ethical standards

Supitcha Thammajinno: Methodology, Investigation, Formal Urine samples in experiments were performed with the
analysis, Writing – original draft, Visualization. Chittanon Bur- approval of the Exemption Determination as was acknowledged
anachai: Writing – review & editing. Proespichaya Kanatha- by Human Research Ethics Committee of Faculty of Medicine,
rana: Resources, Validation, Writing – review & editing. Prince of Songkla University (REC.63-510-19-2).
Panote Thavarungkul: Conceptualization, Methodology, Valida-
tion, Formal analysis, Resources, Writing – review & editing.
Chongdee Thammakhet-Buranachai: Conceptualization, Publisher’s Note
Methodology, Validation, Formal analysis, Resources, Writing –
review & editing. Springer Nature remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

Declaration of Competing Interest


Appendix A. Supplementary material
The authors declare that they have no known competing finan-
cial interests or personal relationships that could have appeared Supplementary data to this article can be found online at
to influence the work reported in this paper. https://doi.org/10.1016/j.saa.2021.120816.

Acknowledgements References

[1] T. Chen, N. Xie, L. Viglianti, Y. Zhou, H. Tan, B.Z. Tang, Y. Tang, Quantitative
This work was financially supported by the Thailand Center of urinalysis using aggregation-induced emission bioprobes for monitoring
Excellence in Physics (ThEP), Prince of Songkla University and chronic kidney disease, Faraday Discuss. 196 (2017) 351–362, https://doi.
Faculty of Science, Prince of Songkla University (grant no. ThEP- org/10.1039/C6FD 00153J.
[2] A. Fatoni, A. Numnuam, P. Kanatharana, W. Limbut, P. Thavarungkul, A novel
61-PHM-PSU2), the Research Assistantship, Faculty of Science, molecularly imprinted chitosan–acrylamide, graphene, ferrocene composite
Prince of Songkla University (contract no. 1-2559-02-010) and cryogel biosensor used to detect microalbumin, Analyst 139 (23) (2014) 6160–
the Center of Excellence for Innovation in Chemistry (PERCH- 6167, https://doi.org/10.1039/C4AN01000K.
[3] E.M. Wrone, M.R. Carnethon, L. Palaniappan, S.P. Fortmann, Association of
CIC). Partial support is also gratefully acknowledged from the
dietary protein intake and microalbuminuria in healthy adults: Third National
Graduate School Dissertation Funding for Thesis, the Center of Health and Nutrition Examination Survey, Am. J. Kidney Dis. 41 (3) (2003)
Excellence for Trace Analysis and Biosensor (TAB-CoE), Division 580–587, https://doi.org/10.1053/ajkd.2003.50119.
of Physical Science, Department of Chemistry, Department of Phy- [4] K/DOQI clinical practice guidelines for chronic kidney disease: evaluation,
classification, and stratification, Am. J. Kidney Dis. 39(2 Suppl 1) (2002) S1-
sics, Faculty of Science, Prince of Songkla University, Hat Yai, 266. PMID: 11904577. https://pubmed.ncbi.nlm.nih.gov/11904577/ (accessed
Songkhla, Thailand. 29 December 2021).

10
S. Thammajinno, C. Buranachai, P. Kanatharana et al. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 270 (2022) 120816

[5] E. Mohabbati-Kalejahi, V. Azimirad, M. Bahrami, A. Ganbari, A review on bridge, Microchim. Acta 186 (6) (2019) 364, https://doi.org/10.1007/s00604-
creatinine measurement techniques, Talanta 97 (2012) 1–8, https://doi.org/ 019-3468-8.
10.1016/j.talanta. 2012.04.005. [28] Q. Tan, J. Qiao, R. Zhang, L. Qi, Copper nanoclusters-modified with papaya juice
[6] B. Gao, Y. Li, Z. Zhang, Preparation and recognition performance of creatinine- for fluorescence turn-on detection of serum l-histidine, Microchem. J. 153
imprinted material prepared with novel surface-imprinting technique, J. (2020), https://doi.org/10.1016/j.microc.2019.104333 104333.
Chromatogr. B 878 (23) (2010) 2077–2086, https://doi.org/10.1016/j. [29] Z. Cai, H. Li, J. Wu, L. Zhu, X. Ma, C. Zhang, Ascorbic acid stabilised copper
jchromb.2010. 06.007. nanoclusters as fluorescent sensors for detection of quercetin, RSC Adv. 10 (15)
[7] C.S. Pundir, S. Yadav, A. Kumar, Creatinine sensors, TrAC, Trends Anal. Chem. 50 (2020) 8989–8993, https://doi.org/10.1039/D0RA01265C.
(2013) 42–52, https://doi.org/10.1016/j.trac.2013.04.013. [30] M.R. Mathew, S.K. Anand, J. Radecki, H. Radecka, K. Girish Kumar, Simple and
[8] S. Gowda, P.B. Desai, S.S. Kulkarni, V.V. Hull, A.A.K. Math, S.N. Vernekar, cost-effective ‘‘turn-on” fluorescence sensor for the determination of xanthine,
Markers of renal function tests, North Am. J. Med. Sci. 2 (4) (2010) 170–173. J. Fluorescence 30 (3) (2020) 695–702, https://doi.org/10.1007/s10895-020-
PMCID: PMC3354405. https://www.ncbi.nlm.nih.gov/pmc/articles/ 02543-w.
PMC3354405/ (accessed 29 December 2021). [31] X.M. Wu, J.H. Zhang, Z.S. Feng, W.X. Chen, F. Zhang, Y. Li, An ultra-sensitive
[9] C.A. Peralta, M.G. Shlipak, S. Judd, M. Cushman, W. McClellan, N.A. Zakai, M.M. ‘‘turn-off” fluorescent sensor for the trace detection of rifampicin based on
Safford, X. Zhang, P. Muntner, D. Warnock, Detection of chronic kidney disease glutathione-stabilized copper nanoclusters, Analyst 145 (4) (2020) 1227–
with creatinine, cystatin C, and urine albumin-to-creatinine ratio and 1235, https://doi.org/10.1039/c9an01994d.
association with progression to end-stage renal disease and mortality, JAMA [32] Z. Aliakbar Tehrani, Z. Jamshidi, M. Jebeli Javan, A. Fattahi, Interactions of
305 (15) (2011) 1545–1552, https://doi.org/10.1001/jama.2011.468. glutathione tripeptide with gold cluster: influence of intramolecular hydrogen
[10] J.H. Contois, C. Hartigan, L.V. Rao, L.M. Snyder, M.J. Thompson, Analytical bond on complexation behavior, J. Phys. Chem. A 116 (17) (2012) 4338–4347,
validation of an HPLC assay for urinary albumin, Clin. Chim. Acta 367 (1) https://doi.org/10.1021/jp2080226.
(2006) 150–155, https://doi.org/10.1016/j.cca.2005.12.002. [33] I.M. Vlasova, A.M. Saletsky, Study of the denaturation of human serum
[11] E.M.K. Leung, W. Chan, A novel reversed-phase HPLC method for the albumin by sodium dodecyl sulfate using the intrinsic fluorescence of albumin,
determination of urinary creatinine by pre-column derivatization with ethyl J. Appl. Spectrosc. 76 (4) (2009) 536–541, https://doi.org/10.1007/s10812-
chloroformate: comparative studies with the standard Jaffé and isotope- 009-9227-6.
dilution mass spectrometric assays, Anal. Bioanal. Chem. 406 (6) (2014) 1807– [34] X. Hu, X. Mao, X. Zhang, Y. Huang, One-step synthesis of orange fluorescent
1812, https://doi.org/10.1007/s00216-013-7592-8. copper nanoclusters for sensitive and selective sensing of Al3+ ions in food
[12] R. Singh, F.W. Crow, N. Babic, W.H. Lutz, J.C. Lieske, T.S. Larson, R. Kumar, A samples, Sens. Actuators, B 247 (2017) 312–318, https://doi.org/10.1016/j.
liquid chromatography-mass spectrometry method for the quantification of snb.2017.03.050.
urinary albumin using a novel 15N-isotopically labeled albumin internal [35] Z. Huang, M. Wang, Z. Guo, H. Wang, H.e. Dong, W. Yang, Aggregation-
standard, Clin. Chem. 53 (3) (2007) 540–542, https://doi.org/ enhanced emission of gold nanoclusters induced by serum albumin and its
10.1373/clinchem.2006.078832. application to protein detection and fabrication of molecular logic gates, ACS
[13] H. Hou, W. Xiong, X. Zhang, D. Song, G. Tang, Q. Hu, LC-MS-MS Measurements Omega 3 (10) (2018) 12763–12769, https://doi.org/10.1021/
of urinary creatinine and the application of creatinine normalization technique acsomega.8b0187510.1021/acsomega.8b01875.s001.
on cotinine in smokers’ 24 hour urine 245415–245415 J. Analyt. Methods [36] S. Shahsavari, S. Hadian-Ghazvini, F. Hooriabad Saboor, I. Menbari Oskouie, M.
Chem. 2012 (2012), https://doi.org/10.1155/2012/245415. Hasany, A. Simchi, A.L. Rogach, Ligand functionalized copper nanoclusters for
[14] V.C. Hernández, M.D.P. Ruiz, R.T. Carrión, Y. Douhal, T. Duran, Albumin by versatile applications in catalysis, sensing, bioimaging, and optoelectronics,
immunonephelometry or immunoturbidimetry? That’s the question, Mater. Chem. Front. 3 (11) (2019) 2326–2356, https://doi.org/10.1039/
Clinica Chimica Acta 493 (2019) S469, https://doi.org/10.1016/j. C9QM00492K.
cca.2019.03.990. [37] S. Kalasin, P. Sangnuang, P. Khownarumit, I.M. Tang, W. Surareungchai,
[15] T. Küme, B. Sağlam, C. Ergon, A.R. Sisman, Evaluation and comparison of abbott Evidence of Cu(I) Coupling with Creatinine Using Cuprous Nanoparticles
jaffe and enzymatic creatinine methods: could the old method meet the new Encapsulated with Polyacrylic Acid Gel-Cu(II) in Facilitating the Determination
requirements?, J Clin. Lab. Anal. 32 (1) (2018), https://doi.org/10.1002/ of Advanced Kidney Dysfunctions, ACS Biomater. Sci. Eng. 6 (2) (2020) 1247–
jcla.22168. 1258.
[16] B. Babamiri, A. Salimi, R. Hallaj, M. Hasanzadeh, Nickel nanoclusters as a novel [38] R. Rajamanikandan, M. Ilanchelian, Protein-protected red emittive copper
emitter for molecularly imprinted electrochemiluminescence based sensor nanoclusters as a fluorometric probe for highly sensitive biosensing of
toward nanomolar detection of creatinine, Biosens. Bioelectron. 107 (2018) creatinine, Anal. Methods 10 (29) (2018) 3666–3674, https://doi.org/
272–279, https://doi.org/10.1016/j.bios.2018.02.022. 10.1039/C8AY00827B.
[17] M.O. Shaikh, P.-Y. Zhu, C.-C. Wang, Y.-C. Du, C.-H. Chuang, Electrochemical [39] H. Huang, H. Li, J.-J. Feng, H. Feng, A.-J. Wang, Z. Qian, One-pot green synthesis
immunosensor utilizing electrodeposited Au nanocrystals and of highly fluorescent glutathione-stabilized copper nanoclusters for Fe3+
dielectrophoretically trapped PS/Ag/ab-HSA nanoprobes for detection of sensing, Sens. Actuators, B 241 (2017) 292–297, https://doi.org/10.1016/j.
microalbuminuria at point of care, Biosens. Bioelectron. 126 (2019) 572– snb. 2016.10.086.
580, https://doi.org/10.1016/j.bios.2018. 11.035. [40] A. Fatoni, A. Numnuam, P. Kanatharana, W. Limbut, C. Thammakhet, P.
[18] A. Diouf, S. Motia, N. El Alami El Hassani, N. El Bari, B. Bouchikhi, Development Thavarungkul, A highly stable oxygen-independent glucose biosensor based
and characterization of an electrochemical biosensor for creatinine detection on a chitosan-albumin cryogel incorporated with carbon nanotubes and
in human urine based on functional molecularly imprinted polymer, J. ferrocene, Sens. Actuators, B 185 (2013) 725–734, https://doi.org/10.1016/j.
Electroanal. Chem. 788 (2017) 44–53, https://doi.org/10.1016/ snb.2013.05.056.
j.jelechem.2017.01.068. [41] T. Kangkamano, A. Numnuam, W. Limbut, P. Kanatharana, P. Thavarungkul,
[19] M. Chen, X. Xiang, K. Wu, H. He, H. Chen, C. Ma, A novel detection method of Chitosan cryogel with embedded gold nanoparticles decorated multiwalled
human serum albumin based on the poly(thymine)-templated copper carbon nanotubes modified electrode for highly sensitive flow based non-
nanoparticles, Sensors (Basel) 17 (11) (2017) 2684, https://doi.org/10.3390/ enzymatic glucose sensor, Sens. Actuators, B 246 (2017) 854–863, https://doi.
s17112684. org/10.1016/j.snb.2017.02.105.
[20] X. Jin, J. Shi, J. Guan, G. Ni, J. Peng, Microwave-assisted synthesis of cholic acid- [42] J. Choosang, S. Khumngern, P. Thavarungkul, P. Kanatharana, A. Numnuam, An
capped gold nanoclusters as fluorescence-enhanced probes for creatinine ultrasensitive label-free electrochemical immunosensor based on 3D porous
detection, NANO 12 (06) (2017) 1750070, https://doi.org/10.1142/ chitosan–graphene–ionic liquid–ferrocene nanocomposite cryogel decorated
s1793292017500709. with gold nanoparticles for prostate-specific antigen, Talanta 224 (2021)
[21] L. Shang, S. Dong, G.U. Nienhaus, Ultra-small fluorescent metal nanoclusters: 121787, https://doi.org/10.1016/j.talanta.2020.121787.
Synthesis and biological applications, NANOToday 6 (4) (2011) 401–418, [43] S.K. Balavandy, K. Shameli, D.R. Biak, Z.Z. Abidin, Stirring time effect of silver
https://doi.org/10.1016/j.nantod.2011.06.004. nanoparticles prepared in glutathione mediated by green method, Chem. Cent.
[22] J. Zheng, P.R. Nicovich, R.M. Dickson, Highly fluorescent noble-metal quantum J. 8 (1) (2014) 11, https://doi.org/10.1186/1752-153x-8-11.
dots, Annu. Rev. Phys. Chem. 58 (1) (2007) 409–431, https://doi.org/10.1146/ [44] X. Gao, C. Du, C. Zhang, W. Chen, Copper nanoclusters on carbon supports for
annurev.physchem.58.032806.104546. the electrochemical oxidation and detection of hydrazine, ChemElectroChem 3
[23] M. Cui, Y. Zhao, Q. Song, Synthesis, optical properties and applications of ultra- (8) (2016) 1266–1272, https://doi.org/10.1002/celc.201600036.
small luminescent gold nanoclusters, TrAC, Trends Anal. Chem. 57 (2014) 73– [45] B. Zhang, B. Chen, J. Wu, S. Hao, G. Yang, X. Cao, L. Jing, M. Zhu, S.H. Tsang, E.H.
82, https://doi.org/10.1016/j.trac.2014.02.005. T. Teo, Y. Huang, The electrochemical response of single crystalline copper
[24] X. Hu, T. Liu, Y. Zhuang, W. Wang, Y. Li, W. Fan, Y. Huang, Recent advances in nanowires to atmospheric air and aqueous solution, Small 13 (10) (2017)
the analytical applications of copper nanoclusters, TrAC, Trends Anal. Chem. 1603411, https://doi.org/10.1002/smll.201603411.
77 (2016) 66–75, https://doi.org/10.1016/j.trac.2015.12.013. [46] L. Meng, J.-H. Yin, Y. Yuan, N. Xu, 11-Mercaptoundecanoic acid capped gold
[25] X. Jia, J. Li, E. Wang, Cu nanoclusters with aggregation induced emission nanoclusters as a fluorescent probe for specific detection of folic acid via a
enhancement, Small 9 (22) (2013) 3873–3879, https://doi.org/10.1002/ ratiometric fluorescence strategy, RSC Adv. 8 (17) (2018) 9327–9333, https://
smll.201300896. doi.org/10.1039/C8RA00481A.
[26] R.S. Aparna, J.S. Anjali Devi, R.R. Anjana, J. Nebu, S. George, Reversible [47] J.R. Lakowicz, Principles of fluorescence spectroscopy, Springer, New York,
fluorescence modulation of BSA stabilised copper nanoclusters for the 2006.
selective detection of protamine and heparin, Analyst 144 (5) (2019) 1799– [48] P. Midoux, P. Wahl, J.-C. Auchet, M. Monsigny, Fluorescence quenching of
1808, https://doi.org/10.1039/C8AN01703D. tryptophan by trifluoroacetamide, Biochimica et Biophysica Acta (BBA) -
[27] J. Pang, Y. Lu, X. Gao, L. He, J. Sun, F. Yang, Z. Hao, Y. Liu, DNA-templated copper General Subjects 801 (1) (1984) 16–25, https://doi.org/10.1016/0304-4165
nanoclusters as a fluorescent probe for fluoride by using aluminum ions as a (84)90207-1.

11
S. Thammajinno, C. Buranachai, P. Kanatharana et al. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 270 (2022) 120816

[49] D.D. Perrin, B. Dempsey, Buffers for pH and metal ion control 864–864 J. Assoc. [59] M. Krieg, K.-J. Gunßer, E. Steinhagen-Thiessen, H. Becker, Vergleichende
Off. Analyt. Chem. 58 (4) (1975), https://doi.org/10.1093/jaoac/58.4.864. quantitative Analytik klinisch-chemischer Kenngrößen im 24-Stunden-urin
[50] R.T. Ambrose, D.F. Ketchum, J.W. Smith, Creatinine determined by ‘‘high- und morgenurin. 24 (11) (1986) 863–870. https://doi.org/10.1515/cclm.1986.
performance” liquid chromatography, Clin. Chem. 29 (2) (1983) 256–259. 24.11.863.
[51] R. Vaz, J. Bettini, J.G.F. Júnior, E.D.S. Lima, W.G. Botero, J.C.C. Santos, M.A. [60] K.N. Mikhelson, Voltammetric sensing of dopamine in urine samples with
Schiavon, High luminescent carbon dots as an eco-friendly fluorescence sensor electrochemically activated commercially available screen-printed carbon
for Cr (VI) determination in water and soil samples, J. Photochem. Photobiol., A electrodes. International Journal of, Biosens. Bioelectron. 4 (4) (2018),
346 (2017) 502–511, https://doi.org/10.1016/j.jphotochem.2017.06.047. https://doi.org/10.15406/ijbsbe.2018.04.00120.
[52] AOAC International, AOAC Official Methods of Analysis (2016), Appendix F: [61] R.E. Wang, L. Tian, Y.H. Chang, A homogeneous fluorescent sensor for human
Guidelines for Standard Method Performance Requirements, 1-18. http:// serum albumin, J. Pharm. Biomed. Anal. 63 (2012) 165–169, https://doi.org/
www.eoma. aoac.org/app_f.pdf (accessed 29 December 2021). 10.1016/j.jpba.2011.12.035.
[53] D.R. Thévenot, K. Toth, R.A. Durst, G.S. Wilson, Electrochemical biosensors: [62] W.R. de Araújo, M.O. Salles, T.R.L.C. Paixão, Development of an enzymeless
recommended definitions and classification, Biosens. Bioelectron. 16 (1–2) electroanalytical method for the indirect detection of creatinine in urine
(2001) 121–131, https://doi.org/10.1016/s0956-5663(01)00115-4. samples, Sens. Actuators, B 173 (2012) 847–851, https://doi.org/10.1016/j.snb.
[54] C. Zhao, Y. Jiao, F. Hu, Y. Yang, Green synthesis of carbon dots from pork and 2012.07.114.
application as nanosensors for uric acid detection, Spectrochim. Acta Part A: [63] T. Verbić, Z. Dorkó, G. Horvai, Selectivity in analytical chemistry, Revue
Mol. Biomol. 190 (2018) 360–367, https://doi.org/10.1016/j.saa.2017.09.037. Roumaine de Chimie 58 (2013) 569–575.
[55] R.P. da Silva, A.W.O. Lima, S.H.P. Serrano, Simultaneous voltammetric [64] W. Gui, X. Chen, Q. Ma, A novel detection method of human serum albumin
detection of ascorbic acid, dopamine and uric acid using a pyrolytic graphite based on CuInZnS quantum dots-Co2+ sensing system, Anal. Bioanal. Chem.
electrode modified into dopamine solution, Anal. Chim. Acta 612 (1) (2008) 409 (15) (2017) 3871–3876, https://doi.org/10.1007/s00216-017-0332-8.
89–98, https://doi.org/10.1016/j.aca.2008.02.017. [65] M.S. Mathew, K. Joseph, Green synthesis of gluten-stabilized fluorescent gold
[56] I. Zelikovic, Hereditary Tubulopathies, in: W. Oh, M. Baum (Eds.), Nephrology quantum clusters: application as turn-on sensing of human blood creatinine,
and fluid/electrolyte Physiology, third Ed., Elsevier, Philadelphia, 2019, pp ACS Sustain. Chem. Eng. 5 (6) (2017) 4837–4845, https://doi.org/10.1021/
315–344. http://doi.org/10.1016/B978-0-323-53367-6.00019-4 (Chapter 19). acssus chemeng.7b00273.
[57] C.J. Saatkamp, M.L. de Almeida, J.A. Bispo, A.L. Pinheiro, A.B. Fernandes, L. [66] S. Mohammadi, G. Khayatian, Highly selective and sensitive photometric
Silveira Jr., Quantifying creatinine and urea in human urine through Raman creatinine assay using silver nanoparticles, Microchim. Acta 182 (7) (2015)
spectroscopy aiming at diagnosis of kidney disease, J. Biomed. Opt. 21 (3) 1379–1386, https://doi.org/10.1007/s00604-015-1460-5.
(2016) 37001, https://doi.org/10.1117/1.Jbo.21.3.037001. [67] Z. Huang, H. Wang, W. Yang, Gold Nanoparticle-based facile detection of
[58] S. Chakraborty, K. Ghosh, R. Hazra, R.N. Biswas, S. Ghosh, A. Bhattacharya, S. human serum albumin and its application as an inhibit logic gate, ACS Appl.
Biswas, Serum and urinary electrolyte levels in cerebro-vascular accident Mater. Interfaces 7 (17) (2015) 8990–8998, https://doi.org/10.1021/
patients: a cross sectional study, Am. J. Int. Med. 1 (2013) 36, https://doi.org/ acsami.5b01552.
10.11648/j.ajim.20130104.13.

12

You might also like