Skip to main content

Highly sensitive plasmonic paper substrate fabricated via amphiphilic polymer self-assembly in microdroplet for detection of emerging pharmaceutical pollutants

Abstract

We report an innovative and facile approach to fabricating an ultrasensitive plasmonic paper substrate for surface-enhanced Raman spectroscopy (SERS). The approach exploits the self-assembling capability of poly(styrene-b-2-vinyl pyridine) block copolymers to form a thin film at the air-liquid interface within the single microdroplet scale for the first time and the subsequent in situ growth of silver nanoparticles (AgNPs). The concentration of the block copolymer was found to play an essential role in stabilizing the droplets during the mass transfer phase and formation of silver nanoparticles, thus influencing the SERS signals. SEM analysis of the morphology of the plasmonic paper substrates revealed the formation of spherical AgNPs evenly distributed across the surface of the formed copolymer film with a size distribution of 47.5 nm. The resultant enhancement factor was calculated to be 1.2 × 107, and the detection limit of rhodamine 6G was as low as 48.9 pM. The nanohybridized plasmonic paper was successfully applied to detect two emerging pollutants—sildenafil and flibanserin—with LODs as low as 1.48 nM and 3.45 nM, respectively. Thus, this study offers new prospects for designing an affordable and readily available, yet highly sensitive, paper-based SERS substrate with the potential for development as a lab-on-a-chip device.

1 Introduction

Block copolymers (bCP) are polymers that consist of at least two different or immiscible polymers connected by covalent bonds. Szwarc was the first to synthesize amphiphilic bCPs using living anionic polymerization back in the 1950s [1]. Since then, various bCP, such as linear, graft, and branched copolymers, have been synthesized using different polymerization methods [2]. The discovery of bCPs’ ability to spontaneously self-assemble in the 1960s [3] revolutionized the field of polymer science and paved the road for their use to generate nanostructures [4, 5], such as micelles [6], nanorods [7], and polymersomes [8]. The ability of bCPs to self-assemble into various nanostructures is mainly governed by their molecular structure, composition, concentration, and response to external factors. Assembly of block copolymers at various interfaces, including air-water [9], oil-water [10], or solid-liquid [11], can generate innovative materials with adjustable shapes, structures, and properties.

In our previously reported work, we developed a novel strategy for the fabrication of composite thin films via the self-assembly of poly(styrene)-b-poly(vinyl pyridine) block copolymer (bCP) incorporating silver ions at the air-liquid interface [12]. The morphology of these films could be easily tuned and varied from nanowires to honeycomb-like structures depending on the concentration of the components. A similarly flexible approach to controlling the formation of silver nanoparticles in the desired fashion could be most interesting for developing a surface-enhanced Raman spectroscopy (SERS) substrate.

The SERS is a powerful vibrational spectroscopy technique that enables the enhancement of the Raman signal by many orders of magnitude, providing fine molecular fingerprints and allowing the detection of trace species [13]. In this regard, the development of highly sensitive SERS substrates, i.e., enabling extensive hot spot generation, has been a cornerstone for achieving extraordinary efficiency to the extent of distinguishing even single molecules [14, 15]. The uniform distribution of metal nanoparticles across the substrate in appropriate proximity to each other is pivotal, as it determines the ability of the SERS-sensitive regions to interact with the analytes [16]. Designing a facile and rapid method for the fabrication of a cost-effective SERS substrate that enables traces of target molecules to be detected continues to remain a major challenge, particularly in resource-limited situations.

Recently, paper substrates have become increasingly popular among researchers in the SERS community owing to the numerous advantages these substrates offer over their traditional counterparts [17, 18]. Paper is readily available, inexpensive, and biodegradable; in addition, its surface can be easily functionalized to vary its affinity toward analytes and solvents [19,20,21]. Apart from this, the porous structure of paper enables its application to microfluidics [22, 23], whereas its mechanical properties present unique opportunities for the collection of real-world samples [24]. Most importantly, the above-listed features can be achieved without sacrificing the sensitivity of the SERS technique, with a satisfactory enhancement factor (EF) ranging from 105 to 108 [18, 25, 26].

Gold and silver nanoparticles have been largely utilized as agents to enhance the Raman scattering in SERS mainly because of their stability and excellent enhancement properties [27, 28]. Metal nanoparticles can be fabricated via self-assembly [29], ink-jet printing [25], the deposition of presynthesized colloidal metal nanoparticles [30], surfactant-free synthesis [31], silver-complex ink [32], etc. Nevertheless, any alternative approach that would entail a more facile yet still efficient procedure devoid of disadvantages such as the cost and operating time remains attractive and worthy of investigation.

Here, we report a cost-effective, easy-to-operate, and environmentally friendly method for the in situ fabrication of a SERS substrate. The method exploits the self-assembly of a bCP in a single microdroplet to form a thin organic film doped with silver nanoparticles (AgNPs) on a paper surface for the first time. While numerous reports discuss the self-assembly of bCP in droplets, these droplets typically consist of emulsion or microfluidic droplets within a stable system [33,34,35]. We managed to achieve the use of exceptionally low volumes of components in the microliter range such that the entire self-assembly process occurs within the air/liquid interface of a single droplet deposited on paper substrate (Fig. 1a). This method does not require the preparation of colloidal metal nanoparticles prior to the fabrication, thereby avoiding complications related to their preparation and storage [36]. Wax-patterned office paper was used as a substrate because of its rough surface, availability, and biodegradability. The denser structure of the office paper compared to Whatman No.1 filter paper allowed the contact angle of the droplet to be increased, thus resulting in uniform fabrication devoid of the notorious “coffee ring” effect [37]. More importantly, the results showed that the bCP has a double function. The first is to promote the self-assembly of silver-loaded films, and the second is the stabilization of the microdroplet. These conditions led to the formation of uniform silver nanospheres with an average size of 47.5 nm. SERS measurements of rhodamine 6G (R6G) demonstrated that the developed substrates have good reproducibility in terms of reaching their limit of detection of 48.9 pM. The practicability of the developed sensor was demonstrated by using it for the detection of the emerging environmental pollutants sildenafil (SD) and flibanserin (FLBN) with calculated limits of detection of 1.48 nM and 3.45 nM, respectively.

Fig. 1
figure 1

Schematic illustration of the preparation of an organic film containing silver nanoparticles at the microliter scale via the self-assembly of poly(styrene)-b-poly(vinyl pyridine) copolymer molecules at the air-liquid interface. (a) FE-SEM images of nanohybridized paper-substrate fabricated with 0.1 mg mL− 1 of bCP and 0.05 M AgNO3 at three different magnifications (b, c, and d). Inset: histogram presenting the size distribution of the silver nanoparticles

2 Methods/Experimental

2.1 Materials

Poly(styrene-b-2-vinyl pyridine) (Mn: 110,000-b-12,500) was purchased from Polymer Source, Inc. (Canada). Silver nitrate (AgNO3, ≥ 99.0%) was acquired from Sigma Aldrich. Sodium borohydride (NaBH4, ≥ 99.0%) was purchased from Fluka Analytical. Rhodamine 6G (99%) was purchased from Acros Organics. Sildenafil citrate (HPLC grade, > 98.0%) and flibanserin (HPLC grade, > 97.0%) were purchased from Tokyo Chemical Industry (TCI chemicals, Japan). Dimethylformamide (ACS grade, 99.8+ %) was purchased form Alfa Aeser. Chloroform (99%) was purchased from Samchun Chemical (South Korea). Conventional office paper (A4|75 g m− 2) was purchased from Hansol Copy, South Korea. Ultrapure water purified by the Milli-Q water purification system (Millipore Corp., MA, USA) was used throughout the experiments. All reagents were used without further purification or modification.

2.2 Instrumentation

SERS measurements were performed on an EnSpectr R532-50 Enspectr Portable Raman Luminescent Spectrometer probe (Enhanced Spectrometry Inc, 560 South Winchester Blvd, #500 San Jose, USA) equipped with a 532-nm wavelength laser and a Jasco NRS-3300 Micro/Macro Raman spectrophotometer with a 532-nm laser. The morphology of the paper surface was studied by examining FE-SEM images captured with a CZ/MIRA I LMH microscope (Czech Republic). UV-Vis absorbance measurements were conducted on a V-670 spectrophotometer (Jasco Inc.). Contact angle of droplets were estimated using software ImageJ and “Contact_Angle” plugin. X-ray diffraction (XRD) was carried out on a Bruker D8 Discover diffractometer with Cu Kα radiation (λ = 0.1542 nm). Samples were fixed to a glass substrate for the analysis.

2.3 Fabrication of plasmonic nanohybridized paper substrate

Wax-patterned multi-well paper sheets with designed hydrophilic wells with diameters of 4 mm were printed on a Xerox ColorQube 8570DN PS (Xerox Corporation Wilsonville, 97,070, OR, USA). The solid wax ink was deposited twice to increase the water-resistance of the wax barriers on the 200 × 200 mm A4 (210 × 297 mm) printing paper sheets (Hansol Paper, Jung-gu, Seoul, Korea). The paper sheets were wrapped in aluminum foil and placed on a hot plate (Daihan Scientific Co., Ltd.) for 30 s at 100 °C to melt the wax to allow for the thorough distribution thereof over the cellulose fibers of the paper, after which the diameter of the wells was reduced from the initial 4 mm to 1.5 mm.

In a typical experiment, 4 µL of AgNO3 of the desired concentration was drop-casted into the wells with a micropipette to form a droplet that occupies the entire surface of each well. Then, 2 µL of bCP dissolved in dimethylformamide (DMF) and chloroform (CHCl3) (0.1 mg mL− 1; VDMF/VCHCl3 = 3/2) was slowly injected into each droplet using a microsyringe. The paper substrate was then placed in the oven for 10 min at 30 °C to allow the solvents to gently evaporate. Evaporation of the solvent led to the formation of a thin copolymer film, the existence of which was divulged by the observed reflection of room light from the surface of the wells. Next, the paper substrate was immersed in an aqueous solution of NaBH4 (0.01 M) for 30 min. The color of the wells changed to dark yellow as a result of the reduction of the Ag+ to form AgNPs. Finally, the paper substrate was dried in the vacuum oven for 10 min at 65 °C and used immediately or within 48 h, during which time it was kept at 4 °C in glass vials filled with nitrogen and wrapped with aluminum foil.

2.4 SERS measurements

SERS measurements were conducted on a R532-50 Enspectr Portable Raman Luminescent Spectrometer probe using a wavelength of 532 nm, a Raman shift range of 109 − 4050 cm− 1, and an integration time of 30 s. In a typical experiment, 5 µL of the analyte (R6G, SD, and FLBN) in different concentrations were drop-casted into the wells of a paper substrate flushed with nitrogen and allowed to dry in the vacuum oven for 15 min. The Raman spectra of the analyte samples were collected across the entire well area. The experiments were repeated three times with different batches.

The SERS mapping technique was exploited to reveal the occurrence of molecular binding events between the SERS-sensitive area of the substrate and R6G. SERS signals were collected over a detection area of 30 × 30 μm with a laser spot size of 3 μm.

3 Results and discussion

3.1 Characterization of plasmonic nanohybridized paper substrate

The ability of block copolymers to form molecular assemblies in solution at the air-liquid interface is utilized to fabricate a thin organic film doped with silver nanoparticles (AgNPs) on the surface of the paper (Fig. 1a). Considering the porosity and surface roughness of the paper substrate, the morphology and distribution of the self-assembled AgNPs@film were studied by acquiring FE-SEM images (Fig. 1b-d). The FE-SEM micrographs revealed the formation of AgNPs with a size distribution of 47.5 nm and their close proximity to each other on the paper superficies.

EDS mapping (Figure S1a and S1b) further confirmed the uniform distribution of the spherical AgNPs across the surface of the paper substrate in close proximity to each other. In addition, the more negligible porosity of the office paper (49%) resulted in the retention of particles close to the surface [38]. As anticipated, FE-SEM analysis of the nanohybridized plasmonic paper fabricated with lower concentrations of AgNO3 revealed lesser coverage of the paper surface with AgNPs, indicating that the concentration of precursors in a typical experiment is optimal for the pattern size (Figs. 1 and S1c-e). Interestingly, the lowest tested concentration of AgNO3 (0.01 M) resulted in the formation of non-uniformly aggregated nanoparticles on the surface of cellulose, whereas higher concentrations of AgNO3 resulted in a more even distribution and greater coverage of the surface of the cellulose fibers. Subsequently, we examined the effect of the bCP concentration on the formation of hybridized plasmonic paper, as shown in Fig. 2a and the supplementary video. Typically, the injection of DMF/CHCl3 containing bCP into the aqueous solution of AgNO3 resulted in mass transfer and self-assembly of the film with different structures imposed by the concentrations of the components. However, at the microdroplet scale, the variation of the bCP concentration directly influences the stability of the droplet. Specifically, the injection of a lower concentration of bCP decreased the stability of the droplets, whereas the injection of DMF/CHCl3 without bCP resulted in the immediate destruction of the droplet. FE-SEM images and droplet stability studies confirmed that the most suitable concentration of bCP and AgNO3 was 0.1 mg mL− 1 and 0.05 M, respectively. After treatment with NaBH4, the films were subjected to XRD analysis to ensure that the reduction of silver ions to AgNPs was successful, as shown in Figure S2. The cellulose lattice planes (110), (200), and (004) are responsible for the diffraction peaks at 2θ = 16.2°, 22.6°, and 35.9°, respectively. Weak diffraction peaks at 2θ = 29.7°, 36.1°, 39.6°, 43.3°, 47.8°, and 48.7° correspond to the (104), (110), (013), (020), (018), and (116) planes of CaCO3, which is found in standard office paper [25]. The peaks at 2θ = 44.4°, 64.5°, and 77.6° are ascribed to the (111), (200), (220), and (311) planes of the AgNPs, respectively.

Fig. 2
figure 2

Effect of bCP concentration on microdroplet stability (a), components concentration-dependent variation in the SERS intensities (1 × 10− 6 M of R6G) (b), UV-Vis absorbance spectra of bare office paper and nanohybridized paper substrate (c), and SERS spectra of R6G on nanohybridized plasmonic paper substrate (d)

To further investigate the most suitable ratio of components for the fabrication of nanohybridized plasmonic paper, we investigated the Raman enhancement for each substrate using R6G as the target analyte. As shown in Fig. 2b, the SERS intensity increases with increasing concentration of the precursors until it reaches a threshold beyond which the SERS signals disappear. For bCP, that limit was at the concentration of 0.1 mg mL− 1, at which the SERS intensity exceeded that of 0.05 mg mL− 1 more than threefold, whereas SERS signals were not obtained at 0.2 mg mL− 1. In this way, the optimal concentration of AgNO3, the precursor agent of the AgNPs, was determined to be 0.05 M. This can be attributed to the aggregation of AgNPs when a high concentration of AgNO3 is utilized, thus preventing the formation of hot spots. In contrast, the excess bCP might cover the surfaces of the AgNPs too thoroughly, thus restricting the attachment of R6G molecules to the AgNPs and preventing them from reaching the localized surface plasmon resonance areas. The particles appeared to be spherical and sporadically with small, truncated corners, which is likely owing to the involvement of bCP in the formation of the particles and the porous structure of the paper. This feature, in this regard, might be beneficial for SERS: the edged structure of metal nanoparticles leads to the more efficient formation of localized surface plasmons.

Classification of the developed substrate as reliable to enhance Raman signals requires the plasmon resonance spectrum to be separate from the charge transfer and absorption spectra. This is because plasmonic materials support electromagnetic SERS enhancement, whereas excitation via the chemical enhancement mechanism of charge transfer alone does not result in the conventional EF of SERS [28]. The absorption peak of the paper substrate on which the AgNPs is deposited is centered at about 400 nm, which is in accordance with data reported elsewhere for AgNPs (Fig. 2c). However, a broad absorption extending to higher wavelengths is contributed by the edged structure of the nanoparticles and is more closely associated with dipole plasmon resonance. Despite the possibility that the wavelength of the lasers used in the Raman experiments might cause fluorescence emission of the office paper samples, this fluorescence was quenched by the deposition of the AgNPs@bCP thin film. Thus far, this phenomenon has been explained in terms of Förster resonance energy transfer because of the presence of the deposited metal nanoparticles [38]. The deposited nanoparticles would also shield any fluorescence originating from the paper material.

The fact that electromagnetic enhancement plays a more significant role in SERS than chemical enhancement has been proven experimentally to date [28]. Successful SERS detection mainly depends on the interaction between the adsorbed molecules and plasmonic nanostructures. The presence of the pyridinic nitrogen atom in each bCP unit creates favorable conditions for the interaction of the substrate with analytes bearing any type of group capable of undergoing hydrogen bonding [39]. This feature was expected to contribute to achieving more prominent EFs inevitably.

3.2 Stability and contact angle of microdroplet

The stability of the microdroplet during the injection of bCP in DMF/CHCl3 was strongly dependent on the concentration of bCP. The effect of bCP on the stability of the microdroplet contact angle was further elucidated by comparing aqueous AgNO3 droplets injected with pure DMF/CHCl3 solution with those injected with bCP (Fig. 3). As a reference, an aqueous AgNO3 droplet injected with the same quantity of DI water was used. All three droplets were deposited on the wax-printed paper, which served as a hydrophobic surface. The reference droplets had an average contact angle of θ = 107.72°, whereas the droplet injected with DMF/CHCl3 had an average contact angle of θ = 74.97°. The observed decrease in the contact angle was induced by adding DMF/CHCl3 with a lower surface tension. The dissolution of bCP in DMF/CHCl3 increases the surface tension of the resulting blend, thus resulting in a droplet with a higher contact angle (Figs. 3 and S3). Moreover, the migration of bCP to the liquid/air interface allows the formation of a film that further stabilizes the microdroplet.

Fig. 3
figure 3

The contact angle of aqueous AgNO3 droplet (left) injected with DI water, (middle) injected with 0.1 mg mL− 1 bCP in DMF/CHCl3, and (right) injected with DMF/CHCl3. Inset: photographs of droplets corresponding to each case deposited on the wax-patterned paper

3.3 SERS performances of the nanohybridized plasmonic paper substrate and spot-to-spot reproducibility

The morphological analysis confirmed that the significant Raman signal enhancement and good reproducibility are attributable to the favorable formation of the metal nanoparticles, which efficiently promote the generation of “hot spots” and propagation of surface plasmon resonances. As shown in the FE-SEM micrographs, along with the small interparticle distance, the presence of some irregular protrusions on the surfaces of the nanoparticles might contribute to the enhancement of the electromagnetic field. The SERS characteristics of the plasmonic paper substrate were investigated by using R6G (Fig. 2d). Overall, the characteristic Raman peak at 611 cm− 1 for R6G was selected to determine the EF, calculated to be 1.2 × 107. However, to obtain a reliable sensor, the “hot spots” should be uniformly distributed to ensure that a stable Raman signal is detected from different locations.

Quantitative SERS detection can be achieved when the analyte concentration changes in a predictable manner [40]. This necessitated an examination of the distribution of the “hot spots” via SERS mapping with a laser spot size of  1 μm. This involved performing SERS measurements over a 10 × 10 μm area on the plasmonic nanohybridized paper substrate (Fig. 4a-d). The SERS mapping technique revealed notable signals at R6G concentrations of 1 × 10− 6 M and below. Furthermore, the probability of detecting noticeable SERS signals with 1 × 10− 6 M of R6G was determined to be equal across the substrate, indicating that the developed system is devoid of the notorious “coffee ring” effect associated with SERS substrates consisting of conventional paper (Fig. 4a). The random selection of ten spots to determine the uniformity of the SERS signal intensity led to the observation that the variance in data was not significant, with the RSD calculated to be 2.49% (Fig. 4b). As anticipated, the probability of observing stable SERS signals decreased when the concentration of R6G was lowered to 1 × 10− 9 M (Fig. 4c). At this concentration, much weaker SERS signals and significant fluctuation of the signal intensity were observed. As shown in Fig. 4d, the RSD of the SERS intensities of the 611 cm− 1 band of R6G (1 × 10− 9 M) among the ten random AgNP spots was 10.95%. To the best of our knowledge, the developed plasmonic paper has excellent spot-to-spot reproducibility for detecting R6G at a concentration of 1 × 10− 6 M.

Fig. 4
figure 4

SERS mapping result of R6G at 611 cm− 1 extracted from 1 × 10− 6 M aqueous solutions (a,b) and from 1 × 10− 9 M aqueous solutions (c,d). RSD for aqueous solutions of rhodamine 6G (R6G; 1 × 10− 6 M and 1 × 10− 9 M) were 2.49% and 10.95%, respectively. SERS spectra of R6G obtained by varying the concentration of R6G (front to back: 10− 9, 10− 8, 10− 7, 10− 6, 10− 5 M) on the nanohybridized paper substrate (e). Relationship between the intensity of the band at 610 cm− 1 and the log concentration of R6G. Inset: linear equation, LOD, and LOQ (f)

3.4 Sensitivity of nanohybridized plasmonic paper

To determine the sensitivity, SERS spectra of nanohybridized paper treated with 5 µL R6G droplets with concentrations ranging from 10− 9 to 10− 6 M were recorded on the portable Raman spectrometer, as shown in Fig. 4e. During the experiments, the major peaks were neither observed to shift nor did their Raman intensity change. The limit of detection (LOD) and limit of quantification (LOQ) of R6G using the nanohybridized paper sensor were estimated by conducting a linear fitting of the SERS intensities versus the log R6G concentration, as shown in Fig. 4f. The calibration curve has excellent linearity for the studied concentrations of R6G ranging from 10− 9 to 10− 6 M with a correlation coefficient of 0.99. The LOD and LOQ of R6G using the developed plasmonic paper sensor were estimated to be 48.9 pM and 0.16 nM, respectively.

The analytical performance of recently reported paper-based plasmonic substrates used for R6G detection was collected from the literature and compared to the results of the current study (Table 1). The sensitivity of our paper substrate is comparable to or exceeded that of previous research, but its spot-to-spot reproducibility is substantially superior.

Table 1 Comparison of SERS parameters with other paper-based plasmonic substrates

3.5 Detection of SD and FLBN using nanohybridized plasmonic paper

As proof of concept, two emerging pollutants, SD and FLBN, were chosen as target analytes. SD is a phosphodiesterase-5 inhibitor (PDE-5i), which was originally developed to treat ischemic heart disease, pulmonary hypertension, and altitude sickness, but has since gained the reputation for treating erectile dysfunction (ED) [41, 42]. Illegal use as an ingredient [43] and the accessibility of generics [44], which are sold without a proper prescription, resulted in leakage into the water environment. Recently, Kim’s group reported that the concentrations of SD in local sewage treatment plants before and after purification were alarmingly high, especially on weekends [45]. Moreover, the results showed that the efficiency of biological nutrient removal (BNR) systems—the modified Ludzack-Ettinger (MLE; anoxic/oxic) and the A2/O (anaerobic/anoxic/oxic) processes—for the elimination of SD was alarmingly low. This reality prompted us to investigate the ability of the nanohybridized plasmonic paper we developed to quantitatively detect SD. We dropped 5 µL of solutions of SD with concentrations ranging from 10− 4 M to 10− 8 M in DI water onto the prepared substrate and recorded the Raman spectra, as shown in Fig. 5a. The characteristic Raman peaks of SD appeared with a slight shift in the SERS peaks compared to the Raman peaks of SD powder (Figure S7a and Table S1). As shown in Fig. 5b, the calibration curve exhibits excellent linearity for the tested SD concentrations ranging from 10− 4 to 10− 8 M, with a correlation value of 0.99. The estimated LOD and LOQ of SD utilizing the developed sensor were 1.48 nM and 2.76 nM, respectively.

Fig. 5
figure 5

SERS spectra of SD (a,b) and FLBN (c,d) obtained by varying the concentration (front to back: 10− 8, 10− 7, 10− 6, 10− 5, 10− 4 M) on the nanohybridized paper substrate

On the other hand, FLBN is the only drug approved by the US Food and Drug Administration (USFDA) for treating hypoactive sexual desire (HASD), found in 7% of premenopausal women [46, 47]. The performance of the developed SERS substrate with respect to detecting FLBN was investigated by twice dropping and drying 5 µL of FLBN solutions with concentrations ranging from 10− 4 M to 10− 8 M, as shown in Fig. 5c. The characteristic Raman peaks of FLBN are detected with the SERS spectra slightly shifted compared to the Raman spectra of FLBN powder (Figure S7b and Table S2). The calibration curve of FLBN concentrations ranging from 10− 4 to 10− 8 M had good linearity with a correlation value of 0.97, as shown in Fig. 5d. The LOD and LOQ of FLBN using the nanohybridized plasmonic paper were estimated to be 3.45 nM and 7.45 nM, respectively.

In the evolving landscape of analytical techniques for pharmaceutical compounds, the developed sensing device based on SERS represents a significant advancement, particularly in the detection of FLBN and SD. The sensitivity of the developed SERS substrate demonstrated comparable, if not superior, performance relative to the established gold standard of chromatography-tandem mass spectrometry for analyzing FLBN, as shown in Table 2. Moreover, this method requires more complex setups and requires well-trained staff. On the other hand, other methods, such as spectrofluorimetry have also been used for sensing FLBN but showed a lower sensitivity, as shown in Table 2.

Table 2 Comparison of reported sensing methods for FLBN and SD

However, it’s important to acknowledge that our developed SERS substrate does not exhibit the same level of selectivity as some alternative SERS substrates modified for specific targets, notably those leveraging molecular imprinting or immunoassays. These techniques benefit from a high degree of specificity towards target molecules, a feature currently less pronounced in the developed SERS substrate. Despite this, the primary objective of this research was to pioneer novel, cost-effective, and simplified methodologies for SERS substrate fabrication based on the self-assembly of bCP in microdroplet.

4 Conclusions

In summary, the self-assembly of block-copolymers (bCPs) at the air-liquid interface to form an organic film is an attractive feature. We took advantage thereof to develop an ultrasensitive detection method based on a SERS substrate in the form of a paper surface deposited with the film doped with silver nanospheres. An unprecedentedly low volume of components was utilized such that the entire fabrication process occurred on the scale of a single microdroplet stabilized by the bCP. FE-SEM micrographs revealed that, with the assistance of self-assembled copolymer film, AgNPs were formed with an average size of 47.5 nm and small, truncated corners. These morphological features of the AgNPs endowed the paper surface with plasmonic properties and leveraged the efficient generation of “hot spots.” The sensitivity of the designed nanohybridized plasmonic paper substrate was investigated by recording the SERS profiles of aqueous solutions of R6G. The generated electromagnetic field enhancement yielded an EF of 1.2 × 107 and LOD of 48.9 pM for R6G. The analysis of two emerging pollutants, SD and FLBN, which was performed as proof of concept, showed that the nanohybridized plasmonic paper delivered excellent analytical performance. The LODs of SD and FLBN were estimated to be 1.48 nM and 3.45 nM, respectively. The novel approach has the advantage of allowing additional steps, such as synthesizing colloidal nanoparticles, to be bypassed during the preparation of the ultrasensitive and reproducible SERS substrate.

Data availability

The datasets used and/or analysed during the current study are available from the corresponding author on reasonable request.

Abbreviations

AgNO3 :

silver nitrate

AgNP:

silver nanoparticles

bCP:

poly(styrene-b-2-vinyl pyridine) block copolymers

CHCl3 :

chloroform

DMF:

dimethylformamide

ED:

erectile dysfunction

EDS:

Energy Dispersive Spectroscopy

EF:

enhancement factor

FLBN:

flibanserin

HASD:

hypoactive sexual desire

LOD:

limit of detection

LOQ:

limit of quantification

NaBH4 :

sodium borohydride

PDE:

5i-phosphodiesterase-5 inhibitor

R6G:

rhodamine 6G

SD:

sildenafil

SERS:

surface-enhanced Raman spectroscopy

USFDA:

US Food and Drug Administration

XRD:

X-ray diffraction

References

  1. M. Szwarc, ‘Living’ Polym. Nat. 178, 1168–1169 (1956). https://doi.org/10.1038/1781168a0

    Article  CAS  Google Scholar 

  2. S. Zhang et al., Building block copolymer particles via self-assembly within a droplet. Droplet 2, e81 (2023). https://doi.org/10.1002/dro2.81

  3. I.E. Climie, E.F.T. White, The aggregation of random and block copolymers containing acrylonitrile in mixed solvents. J. Polym. Sci. 47, 149–156 (1960). https://doi.org/10.1002/pol.1960.1204714913

    Article  CAS  Google Scholar 

  4. M. Karayianni, S. Pispas, Block copolymer solution self-assembly: recent advances, emerging trends, and applications. J. Polym. Sci. 59, 1874–1898 (2021). https://doi.org/10.1002/pol.20210430

    Article  CAS  Google Scholar 

  5. J.H. Kim et al., Smart Nanostructured materials based on self-assembly of Block copolymers. Adv. Funct. Mater. 30, 1902049 (2020). https://doi.org/10.1002/adfm.201902049

    Article  CAS  Google Scholar 

  6. S. Quader et al., Supramolecularly enabled pH- triggered drug action at tumor microenvironment potentiates nanomedicine efficacy against glioblastoma. Biomaterials. 267, 120463 (2021). https://doi.org/10.1016/j.biomaterials.2020.120463

    Article  CAS  PubMed  Google Scholar 

  7. S. Kim et al., Self-assembled pagoda-like nanostructure-induced vertically stacked split-ring resonators for polarization-sensitive dichroic responses. Nano Convergence. 9, 40 (2022). https://doi.org/10.1186/s40580-022-00331-9

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. L.J.C. Albuquerque et al., pH-responsive polymersome-mediated delivery of doxorubicin into tumor sites enhances the therapeutic efficacy and reduces cardiotoxic effects. J. Controlled Release. 332, 529–538 (2021). https://doi.org/10.1016/j.jconrel.2021.03.013

    Article  CAS  Google Scholar 

  9. I.I. Perepichka, A. Badia, C.G. Bazuin, Nanostrand Formation of Block Copolymers at the Air/Water Interface. ACS Nano. 4, 6825–6835 (2010). https://doi.org/10.1021/nn101318e

    Article  CAS  PubMed  Google Scholar 

  10. Q. Yuan, T.P. Russell, D. Wang, Self-assembly behavior of PS-b-P2VP Block copolymers and Carbon Quantum Dots at Water/Oil Interfaces. Macromolecules. 53, 10981–10987 (2020). https://doi.org/10.1021/acs.macromol.0c02422

    Article  CAS  Google Scholar 

  11. A.M. Bodratti, B. Sarkar, P. Alexandridis, Adsorption of poly(ethylene oxide)-containing amphiphilic polymers on solid-liquid interfaces: fundamentals and applications. Adv. Colloid Interface Sci. 244, 132–163 (2017). https://doi.org/10.1016/j.cis.2016.09.003

    Article  CAS  PubMed  Google Scholar 

  12. X. Zhao et al., A new strategy to fabricate composite thin films with tunable micro- and nanostructures via self-assembly of block copolymers. Chem. Commun. 51, 16687–16690 (2015). https://doi.org/10.1039/C5CC05548B

    Article  CAS  Google Scholar 

  13. M.A. Hussein, W.A. El-Said, B.M. Abu-Zied, J.-W. Choi, Nanosheet composed of gold nanoparticle/graphene/epoxy resin based on ultrasonic fabrication for flexible dopamine biosensor using surface-enhanced Raman spectroscopy. Nano Convergence. 7, 15 (2020). https://doi.org/10.1186/s40580-020-00225-8

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. E.C.L. Ru, P.G. Etchegoin, Surface-Enhanced Raman Spectroscopy. Annu. Rev. Phys. Chem. 63, 65–87 (2012). https://doi.org/10.1146/annurev-physchem-032511-143757

    Article  CAS  PubMed  Google Scholar 

  15. M. Lu, Y. Joung, C.S. Jeon et al., Dual-mode SERS-based lateral flow assay strips for simultaneous diagnosis of SARS-CoV-2 and influenza a virus. Nano Convergence. 9, 39 (2022). https://doi.org/10.1186/s40580-022-00330-w

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. S. Eslami, S. Palomba, Integrated enhanced Raman scattering: a review. Nano Convergence. 8, 41 (2021). https://doi.org/10.1186/s40580-021-00290-7

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. S. Kim et al., Early and direct detection of bacterial signaling molecules through one-pot au electrodeposition onto paper-based 3D SERS substrates. Sens. Actuators B 358, 131504 (2022). https://doi.org/10.1016/j.snb.2022.131504

    Article  CAS  Google Scholar 

  18. M. Sharipov et al., Novel molecularly imprinted nanogel modified microfluidic paper-based SERS substrate for simultaneous detection of bisphenol A and bisphenol S traces in plastics. J. Hazard. Mater. 461, 132561 (2024). https://doi.org/10.1016/j.jhazmat.2023.132561

    Article  CAS  PubMed  Google Scholar 

  19. S.M. Tawfik et al., Dual emission nonionic molecular imprinting conjugated polythiophenes-based paper devices and their nanofibers for point-of-care biomarkers detection. Biosens. Bioelectron. 160, 112211 (2020). https://doi.org/10.1016/j.bios.2020.112211

    Article  CAS  PubMed  Google Scholar 

  20. Y. Liu, Y. Li, Y. Hang et al., Rapid assays of SARS-CoV-2 virus and noble biosensors by nanomaterials. Nano Convergence. 11, 2 (2024). https://doi.org/10.1186/s40580-023-00408-z

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. M. Lee et al., Subnanomolar Sensitivity of Filter Paper-based SERS Sensor for Pesticide Detection by Hydrophobicity Change of Paper Surface. ACS Sens. 3, 151–159 (2018). https://doi.org/10.1021/acssensors.7b00782

    Article  CAS  PubMed  Google Scholar 

  22. K. Danchana et al., Determination of glutamate using paper-based microfluidic devices with colorimetric detection for food samples. Microchem. J. 179, 107513 (2022). https://doi.org/10.1016/j.microc.2022.107513

    Article  CAS  Google Scholar 

  23. S. Azizov et al., Solvent-resistant microfluidic paper-based analytical device/spray mass spectrometry for quantitative analysis of C18-ceramide biomarker. J. Mass Spectrom. 56, e4611 (2021). https://doi.org/10.1002/jms.4611

    Article  CAS  PubMed  Google Scholar 

  24. C.H. Lee, L. Tian, S. Singamaneni, S.E.R.S. Paper-Based, Swab for Rapid Trace Detection on Real-World Surfaces, ACS Appl. Mater. Interfaces. 2, 3429–3435 (2010). https://doi.org/10.1021/am1009875

    Article  CAS  PubMed  Google Scholar 

  25. T.T. Thuy et al., Inkjet-based microreactor for the synthesis of silver nanoparticles on plasmonic paper decorated with chitosan nano-wrinkles for efficient on-site surface-enhanced Raman Scattering (SERS). Nano Select. 1, 499–509 (2020). https://doi.org/10.1002/nano.202000081

    Article  Google Scholar 

  26. A. Abbas et al., Multifunctional Analytical platform on a paper Strip: separation, Preconcentration, and Subattomolar Detection. Anal. Chem. 85, 3977–3983 (2013). https://doi.org/10.1021/ac303567g

    Article  CAS  PubMed  Google Scholar 

  27. A. Haryanto, C.W. Lee, Shell isolated nanoparticle enhanced Raman spectroscopy for mechanistic investigation of electrochemical reactions. Nano Convergence. 9, 9 (2022). https://doi.org/10.1186/s40580-022-00301-1

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. B. Sharma et al., Materials, applications, and the future. Mater. Today. 15, 16–25 (2012). https://doi.org/10.1016/S1369-7021(12)70017-2

    Article  CAS  Google Scholar 

  29. Z. Zhang, Y. Lee, M.F. Haque et al., Plasmonic sensors based on graphene and graphene hybrid materials. Nano Convergence. 9, 28 (2022). https://doi.org/10.1186/s40580-022-00319-5

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Y. Xu et al., SERS as a probe of Surface Chemistry enabled by Surface-Accessible Plasmonic nanomaterials. Acc. Chem. Res. 56, 2072–2083 (2023). https://doi.org/10.1021/acs.accounts.3c00207

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. G. Fang et al., Interfacial self-assembly of surfactant-free au nanoparticles as a clean surface-enhanced Raman scattering substrate for quantitative detection of As5 + in combination with convolutional neural networks. J. Phys. Chem. Lett. 14, 7290–7298 (2023). https://doi.org/10.1021/acs.jpclett.3c01969

    Article  CAS  PubMed  Google Scholar 

  32. M. Sharipov, Y. Lee, J. Han, Y.-I. Lee, Patterning microporous paper with highly conductive silver nanoparticles via PVP-modified silver–organic complex ink for development of electric valves. Mater. Adv. 2, 3579–3588 (2021). https://doi.org/10.1039/D0MA00960A

    Article  CAS  Google Scholar 

  33. A. Steinhaus et al., Self-assembly of Diblock Molecular Polymer brushes in the spherical confinement of Nanoemulsion droplets. Macromol. Rapid Commun. 39, 1800177 (2018). https://doi.org/10.1002/marc.201800177

    Article  CAS  Google Scholar 

  34. J. Liu et al., Cucurbit[n]uril-Based microcapsules Self-assembled within Microfluidic droplets: a Versatile Approach for Supramolecular architectures and materials. Acc. Chem. Res. 50, 208–217 (2017). https://doi.org/10.1021/acs.accounts.6b00429

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. V. Jain, V.B. Patel, B. Singh, D. Varade, Microfluidic device based molecular self-assembly structures. J. Mol. Liq. 362, 119760 (2022). https://doi.org/10.1016/j.molliq.2022.119760

    Article  CAS  Google Scholar 

  36. H. Duan, D. Wang, Y. Li, Green chemistry for nanoparticle synthesis. Chem. Soc. Rev. 44, 5778–5792 (2015). https://doi.org/10.1039/C4CS00363B

    Article  CAS  PubMed  Google Scholar 

  37. Y.-F. Li, Y.-J. Sheng, H.-K. Tsao, Evaporation stains: suppressing the Coffee-Ring Effect by Contact Angle Hysteresis. Langmuir. 29, 7802–7811 (2013). https://doi.org/10.1021/la400948e

    Article  CAS  PubMed  Google Scholar 

  38. M.J. Oliveira et al., Office paper decorated with silver nanostars - an alternative cost effective platform for trace analyte detection by SERS. Sci. Rep. 7, 2480 (2017). https://doi.org/10.1038/s41598-017-02484-8

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. S. Malynych, I. Luzinov, G. Chumanov, Poly(vinyl pyridine) as a Universal Surface modifier for immobilization of nanoparticles. J. Phys. Chem. B 106, 1280–1285 (2002). https://doi.org/10.1021/jp013236d

    Article  CAS  Google Scholar 

  40. S. Yang, X. Dai, B.B. Stogin, T.-S. Wong, Ultrasensitive surface-enhanced Raman scattering detection in common fluids. Proceedings of the National Academy of Sciences 113, 268–273 (2016). https://doi.org/10.1073/pnas.1518980113

  41. B.-P. Jiann, Evolution of phosphodiesterase type 5 inhibitors in treatment of erectile dysfunction in Taiwan. Urol. Sci. 27, 66–70 (2016). https://doi.org/10.1016/j.urols.2016.04.003

    Article  Google Scholar 

  42. M. Boolell, S. Gepi-Attee, J.C. Gingell, M.J. Allen, Sildenafil, a novel effective oral therapy for male erectile dysfunction. Br. J. Urol. 78, 257–261 (1996). https://doi.org/10.1046/j.1464-410X.1996.10220.x

    Article  CAS  PubMed  Google Scholar 

  43. J.H. Jeong et al., LC-ESI-MS/MS analysis of phosphodiesterase-5 inhibitors and their analogues in foods and dietary supplements in Korea. Food Addit. Contaminants: Part. B 9, 1–8 (2016). https://doi.org/10.1080/19393210.2014.968220

    Article  CAS  Google Scholar 

  44. A. Causanilles, E. Emke, P. de Voogt, Determination of phosphodiesterase type V inhibitors in wastewater by direct injection followed by liquid chromatography coupled to tandem mass spectrometry. Sci. Total Environ. 565, 140–147 (2016). https://doi.org/10.1016/j.scitotenv.2016.04.158

    Article  CAS  PubMed  Google Scholar 

  45. Y. Hong et al., Contribution of sewage to occurrence of phosphodiesterase-5 inhibitors in natural water. Sci. Rep. 11, 9470 (2021). https://doi.org/10.1038/s41598-021-89028-3

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. M.H. Mahnashi et al., A novel design and facile synthesis of nature inspired poly (dopamine-Cr3+) nanocubes decorated reduced graphene oxide for electrochemical sensing of flibanserin. Microchem. J. 164, 106020 (2021). https://doi.org/10.1016/j.microc.2021.106020

    Article  CAS  Google Scholar 

  47. R.C. Rosen et al., Sexual Desire problems in women seeking Healthcare: a Novel Study Design for ascertaining prevalence of hypoactive sexual Desire Disorder in Clinic-based samples of U.S. women. J. Women’s Health. 21, 505–515 (2012). https://doi.org/10.1089/jwh.2011.3002

    Article  Google Scholar 

  48. F. Yang et al., Printer-assisted array flexible surface-enhanced Raman spectroscopy chip preparation for rapid and label-free detection of bacteria. J. Raman Spectrosc. 51, 932–940 (2020). https://doi.org/10.1002/jrs.5857

    Article  CAS  Google Scholar 

  49. E. Rodríguez-Sevilla, G.V. Vázquez, E. Morales-Narváez, Simple, flexible, and Ultrastable Surface enhanced Raman scattering substrate based on Plasmonic Nanopaper decorated with Graphene Oxide. Adv. Opt. Mater. 6, 1800548 (2018). https://doi.org/10.1002/adom.201800548

    Article  CAS  Google Scholar 

  50. Z. Huang et al., Painting and heating: a nonconventional, scalable route to sensitive biomolecular analysis with plasmon-enhanced spectroscopy. J. Raman Spectrosc. 48, 1365–1374 (2017). https://doi.org/10.1002/jrs.5226

    Article  CAS  Google Scholar 

  51. K. Sridhar, B.S. Inbaraj, B.-H. Chen, An improved surface enhanced Raman spectroscopic method using a paper-based grape skin-gold nanoparticles/graphene oxide substrate for detection of rhodamine 6G in water and food. Chemosphere 301, 134702

  52. M. Poplawska, A. Blazewicz, P. Zolek, Z. Fijalek, Determination of flibanserin and tadalafil in supplements for women sexual desire enhancement using high-performance liquid chromatography with tandem mass spectrometer, diode array detector and charged aerosol detector. Journal of Pharmaceutical and Biomedical Analysis 94, 45–53 (2014). https://doi.org/10.1016/j.jpba.2014.01.021

  53. R.M. Ahmed, I.A. Abdallah, Determination of flibanserin in the presence of confirmed degradation products by a third derivative emission spectrofluorometric method: application to pharmaceutical formulation. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 225, 117491 (2020). https://doi.org/10.1016/j.saa.2019.117491

    Article  CAS  Google Scholar 

  54. R.A. Farghali, R.A. Ahmed, A Novel Electrochemical Sensor for determination of Sildenafil Citrate (Viagra) in pure form and in Biological and Pharmaceutical formulations. Int. J. Electrochem. Sci. 7, 13008–13019 (2012). https://doi.org/10.1016/S1452-3981(23)16605-3

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (NRF-2023R1A2C1004414).

Author information

Authors and Affiliations

Authors

Contributions

Mirkomil Sharipov: Conceptualization, Methodology, Investigation, Formal analysis, Writing - original draft, Writing - review & editing, Data curation. Sarvar A. Kakhkhorov: Conceptualization, Synthesis, Methodology, Investigation, Writing - original draft. Salah M. Tawfik: Conceptualization, Writing - review & editing. Shavkatjon Azizov: Writing -Measurements, Data curation. Hong-Guo Liu: Methodology, Writing - review & editing. Joong Ho Shin: Investigation, Writing - review & editing. Yong-Ill. Lee: Conceptualization, Resources, Formal analysis, Data curation, Writing - review & editing, Supervision, Project administration, Funding acquisition.

Corresponding author

Correspondence to Yong-Ill Lee.

Ethics declarations

Competing interest

The authors declare no conflict of interest.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary Material 1

Supplementary Material 2

Supplementary Material 3

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sharipov, M., Kakhkhorov, S.A., Tawfik, S.M. et al. Highly sensitive plasmonic paper substrate fabricated via amphiphilic polymer self-assembly in microdroplet for detection of emerging pharmaceutical pollutants. Nano Convergence 11, 13 (2024). https://doi.org/10.1186/s40580-024-00420-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40580-024-00420-x

Keywords