- Research
- Open access
- Published:
Nanocomposite-based PCR reactors to enhance thermal rate and fluorescence intensity in hand-held qPCR device
Journal of Nanobiotechnology volume 23, Article number: 240 (2025)
Abstract
A photonic quantitative polymerase chain reaction (qPCR) has usually implemented a polydimethylsiloxane (PDMS) based disposable inexpensive PCR reactor, worked as the photothermal cycler, to show potential as a point-of-care test (PoCT) for detection nucleic acids. However, the PoCT type photonic qPCR has to overcome the prolonged time for the fabrication of PDMS-based PCR reactors and enable a rapid thermal cycler to shorten diagnosis time with a strong fluorescence intensity. Here, we developed a room-temperature curable titanium dioxide (TiO2) nanoparticle dispersed PDMS (TiO2-PDMS) nanocomposite to reduce the fabrication time of the PCR reactor which enhanced the speed of photothermal cycles and fluorescence signal intensity of photonic qPCR. The TiO2-PDMS nanocomposite was formulated for rapid cross-linking at the room-temperature by introducing an optimized amount of Pt catalyst, resulting in the fabrication of a nanocomposite-based PCR reactor within 8 min at room-temperature. The nanocomposite-based PCR reactor enhanced the heating rate to 18.33 Cº/s and cooling rate to −3.11Cº/s because of the phonon scattering effect of TiO2 in the reactor and successfully amplified λ-DNA (amplicon size of 100 bp) within 10 min. Finally, we improved the qPCR efficiency by 2 cycle threshold (Ct) value compared with pristine PDMS reactor and quantified up to 10 copies/µL nucleic acids by fluorescence intensity enhancement resulting from light reflections property of TiO2. By using TiO2-PDMS nanocomposite-based PCR reactors, the fast and efficient nucleic acid assay was enabled without loss of sensitivity, and it can be practically used in the field of PoCT.
Graphical Abstract

Introduction
Viral infectious diseases need to be accurately and quickly recognized for treating them specifically. To diagnose those diseases, various methods exist, including polymerase chain reaction (PCR), isothermal nucleic acid amplification techniques (iNAAT) such as loop-mediated isothermal amplification (LAMP) and recombinase polymerase amplification (RPA), as well as antigen–antibody based lateral flow assay (LFA). The LAMP offers high sensitivity and enables more affordable and rapid testing but suffers from a notable drawback: elevated false-positive rates due to substantial non-specific amplification. The RPA, while beneficial for its rapidity, simplicity, and selectivity, requires subsequent purification and protein design steps. In addition, this method is prone to non-specific amplification. The LFA boasts high speed, specificity, and simplicity but lacks sensitivity as it relies on detecting antibody responses to infections and viral antigens. Despite the simplicity and speed of these iNAAT and LFA, the quantitative PCR (qPCR) technique remains the golden standard for molecular diagnosis to detect nucleic acids with the highest sensitivity, specificity, and accuracy [1,2,3,4]. Thus, it has been highly demanded to be further developed as a point-of-care testing (PoCT) [5,6,7,8,9,10].
To meet the required level of PoCT in terms of cost, size, speed, and sensitivity, the convection-based [4, 11, 12], silicon-based [13,14,15], and photonic-based qPCR devices [1, 16,17,18,19] have been explored. Among developed PoCT-type qPCR devices, the photonic qPCR device has implemented a simple and efficient photothermal converter instead of a Peltier thermocycler. The photothermal converter usually employs surface plasmonic resonance (SPR) phenomena of nanostructured metal or carbon materials by shining the light with a resonant wavelength to induce plasmonic hot electrons [20,21,22]. Since the induced hot electrons are dissipated as heat, the thermal cycles of the photothermal converter can be controlled by simply turning on and off the lights with the resonance wavelength. Usually two-step thermal cycle (denaturation and annealing/extension) is utilized for the rapid photonic qPCR [23] where the thermal ramping rate has been dramatically enhanced by using the SPR while enhancing the excited fluorescence intensity through the light reflection of nanoparticles in the disposable reactor [1, 24, 25]. For example, a plasmonic nano substrate, made through reactive ion etching and thermal evaporation, was used as a plasmonic thermal cycler [10]. Another study used gold nanopillar arrays, fabricated using photolithography and electron beam evaporation, then integrated with a PDMS mold to form the PCR chamber [26]. While those designs offer good thermal performance, they face challenges in mass production because of a high cost and emission of environmentally harmful byproducts. Therefore, to fabricate plasmonic PCR reactors with low-cost and environmentally benign method, sustainable and high-throughput manufacturing methods should be considered to fabricate the disposable plasmonic PCR reactor with efficient photothermal conversion and enhanced fluorescence intensity.
To fabricate the disposable PCR reactor, polydimethylsiloxane (PDMS)-based PCR reactor was usually employed because of its biocompatibility and non-autofluorescence properties [27]. However, these PDMS-based reactors have low values of thermal conductivities (k), owing to the random arrangement of the macromolecular chains in PDMS [28, 29], thereby creating a non-uniform distribution of incident heat in the disposable reactor. Thus, adding thermally conductive additives [28, 30, 31] will enhance the cooling rate by improving heat exchange between the hotspots near the heat source and the surroundings along with the optical characteristics, and thus could be used to fabricate the disposable reactors with the photothermal converting function. In addition, if thermally conductive additives can enhance the fluorescence intensity by scattering and reflection of fluorescence light, the disposable reactor would be in tandem to provide efficient photothermal conversion and enhance the fluorescence intensity. Thus, a biocompatible nanoparticle, having good thermal conductivity and enhancing fluorescence intensity, should be selected and homogeneously dispersed in the PDMS precursors without disturbing the mechanical properties of the disposable PCR reactor.
In this work, our previous photonic qPCR device, developed using a carbon-graphene mixed plasmonic film [19], was further improved for PoCT use by enhancing both the photothermal cycling and fluorescence signal intensity. For this purpose, TiO2-loaded PDMS (TiO2-PDMS) nanocomposite was designed and formulated to fabricate disposable plasmonic PCR reactors to run the molecular diagnostic assay with lambda DNA (λ-DNA) as a reference sample for proof-of-concept of the PoCT-type qPCR device. Since TiO2 nanoparticles have photocatalytic activity, the nanoparticles should be encapsulated in the PDMS well for detecting target genes with the enhanced speed of amplification and fluorescence intensity in the photonic qPCR, instead of dispersing in the PCR reaction solution as reported previously [32, 33]. Hence, this work is reported as a first-of-its-kind study about the formulation of a room-temperature curable TiO2-PDMS to fabricate PCR reactors to improve both the thermal cycling and fluorescence detection efficiency of the PoCT-type photonic qPCR device. To deliver the novelty of our works, Fig. 1 schematically illustrates the outline of this work: the development of the TiO2-PDMS-based disposable PCR reactor with the carbon-graphene mixture-based photothermal conversion function (Fig. 1a). This reactor can shorten assay time by the phonon scattering of TiO2 nanoparticles based on the Boltzmann transport equation [34], and enhancing the fluorescence signal by the incident light reflection of TiO2 nanoparticles [35, 36] in the PDMS reactor (Fig. 1b, c).
Schematic representation of a photonic PCR along with the data acquisition system used for the target gene assay. The near-infrared LED (940 nm of wavelength) is used for running the photothermal cycler by simply “ON” and “OFF” the LED while the fluorescence detection camera will be used to monitor emitted the fluorescence signal, b enhancing the cooling rate in the thermal cycles due to improved thermal conductivity via the phonon scattering, and c mechanism for incident light reflection based fluorescent intensity enhancement and consequent reduction in Ct values in the qPCR data analysis
Materials and methods
TiO2-PDMS nanocomposite
Poly(dimethylsiloxane)-vinyl terminated (referred to as A), poly(dimethylsiloxane-co-methylhydroxane)-trimethylsilyl terminated (referred to as B), Karstedt’s catalyst (referred to as C) and TiO2 nanoparticles (anatase grade-99.7 %, with size ~25 nm) (referred to as D) were purchased from Sigma Aldrich (Korea) to formulate the TiO2-PDMS nanocomposite to imprint the PCR reactors. Polyethylene terephthalate (PET) with a width of 25 cm and thickness of 100 µm was acquired from SKC (AH71D, Korea). All the chemicals were used without additional purification.
PCR reagents
The λ-DNA (from Bacteriophage lambda cl857 Sam 7) was purchased from Roche Applied Science. Diethyl pyrocarbonate (DEPC) and polyethylene glycol (PEG) were purchased from Sigma Aldrich. TB Green Fast qPCR Mix (2X) containing mutant-type Taq HS polymerase, dNTP mixture, Mg2+, Tli RNaseH, and TB Green mixed at 1X were obtained from Takara for rapid PCR. Forward primer (5’-GTT CTG AGG TCA TTA CTG GAT C-3’) and reverse primer (5’-TAC GCT CCT GTC CGG CAA A-3’) were purchased from Cosmo GENETECH (Korea). Mineral oil from Sigma Aldrich was used as thermal insulation for the PCR mixture. E-Gel™ EX Agarose Gels (2% and 4%) were purchased from Thermo Fisher Scientific for gel electrophoresis. A PCR solution of 1 µL was used for the assay, which included a λ-DNA target amplicon size of 100 bp, 1X Taq polymerase, 0.4 µM of each primer, and 5 wt% of PEG. In addition, a benchtop qPCR (StepOne™ Real-Time PCR system) from Applied Biosystems was used as a reference device to prove the concept of enhancing fluorescence intensity by using TiO2-PDMS PCR reactors.
Formulation of TiO2-PDMS precursor
For fabricating PCR reactors using an imprinting method, A and B were first mixed using a mechanical stirrer with an optimized weight ratio of 5:1 at room-temperature (~ 24 °C) for 2 h. Thereafter, D was added to the mixture in different ratios (0.1, 0.25, 0.5 and 0.75 wt%) and sonicated for 6 h at room-temperature to minimize aggregation. The solution was subsequently sealed to avoid any moisture and cooled till −4 °C. To optimize the curing time, different concentrations of platinum (Pt) (120, 340, and 400 ppm) contained in C, were added to the formulated A, B, and D (TiO2-PDMS precursors), degassed for 10 min in the vacuum desiccator, and then cured, respectively, at 24, 70, 90, and 110 °C (Figure S1). After optimizing the catalyst concentration (400 ppm), TiO2-PDMS was successfully cured at room-temperature in 8 min based on the hydrosilylation process (Fig. 2a) [37, 38].
a Synthesis of TiO2-PDMS nanocomposite through hydrosilylation, b illustration of the imprinting process to fabricate PCR reactors (the fabrication time does not include the TiO2-PDMS ink formulation time), c cross-sectional analysis of 1 mm diameter reactor of pristine PDMS and 0.25 wt% TiO2-PDMS through an optical microscope (1600 × 1200, 400 µm scale), d FE-SEM images (500 nm scale), e EDS spectrum in the SEM image which includes 0.25 wt% of TiO2, indicating the dispersed nanoparticles. f Raman spectra and g UV–Vis absorption spectra with pristine PDMS, 0.1, and 0.25 wt% TiO2-PDMS nanocomposites, respectively
Imprinting process
The TiO2-PDMS nanocomposite-based PCR reactors were imprinted as shown in Fig. 2b. First, the formulated TiO2-PDMS precursor was poured onto the printed carbon-graphene on PET substrate as the photothermal conversion layer and then, the Parylene C coated flexible polyurethane mold [39] was used as a positive mold to imprint the TiO2-PDMS nanocomposite-based PCR reactors, cured at room-temperature for 8 min. The resulting TiO2-PDMS nanocomposite-based PCR reactors were easily detached from the original mold and had a diameter of 1 mm [19].
Portable photonic PCR device set-up
Although details of the photonic PCR device were reported previously [19], general procedures are summarized here. The printed carbon-graphene mixed plasmonic film was used to induce the surface plasmonic resonance for thermal cycling by exposing the light of near-infrared (NIR) from the light-emitting diode (LED). The LED with a wavelength of 940 nm (LZ4-00R708-0000) and optical power of 3.7 W was positioned at 5 mm from the bottom of the PCR reactor (Figure S2a). A PID-based closed-loop control system was employed for temperature regulation, using an 8-bit pulse width modulation (PWM) of frequency 495 Hz to control the NIR LED intensity. The system was connected to a medical application-based power supply (TPP30A-J). The temperature change was monitored using a miniature K-type thermocouple (5SRTC-TT-(K)-40-36), wherein the data was sampled every 50 ms for the control system (Figure S2b). Once the assay commences, the fluorescence readings were obtained using a commercially available fluorescence microscope camera (AM4117MT-G2FBW), equipped with 465 nm excitation and 510–545 nm emission filters purchased from Dion-Lite (Korea). The readings were decoded via custom software developed using MATLAB (Figure S2c), with an end-user interface to control the device settings and evaluate the qPCR assay parameters.
Results and discussion
Characterization of TiO2-PDMS reactor
The synthesized TiO2-PDMS nanocomposites were characterized to analyze the distribution of the TiO2 nanoparticles and study the optical properties of the fabricated PCR reactors to test their suitability for fluorescence detection. Figure 2a, b show the synthetic scheme of room-temperature curable pristine PDMS and TiO2-PDMS nanocomposites, fabricated to form PCR reactors on the carbon-graphene mixed plasmonic film. For evaluating the surface morphology and composition of the synthesized nanocomposite, samples with pristine PDMS and 0.25 wt% of TiO2 in PDMS were prepared for characterization by optical microscope (Olympus, BX53M), field emission scanning electron microscope, and energy dispersive spectroscopy (FE-SEM/EDS) (Hitachi, S-4800). Initially, the cross-sections of frustums of cone-shaped reactors were imprinted on the plasmonic film, having a PDMS thickness of 30 µm beneath the PCR reaction compartment, and were observed using the optical microscope (Fig. 2c). The samples were characterized by FE-SEM to see the dispersion of nanoparticles in the PDMS, and EDS analysis confirmed the presence of the nanoparticles (Fig. 2d, e). To study the possible effect of surface roughness on the enhancement of heat transfer [40], atomic force microscopy (AFM) (PSIA, XE100) was performed for pristine PDMS, 0.1, and 0.25 wt% of TiO2-PDMS nanocomposites. Each 10 × 10 µm2 area of the samples was analyzed with a scan rate of 0.5 Hz, in non-contact mode. Since the roughness of surfaces affects the cooling rate [41, 42], the root mean square (RMS) surface roughness was observed to increase with the weight percentage of TiO2 in PDMS (Figure S3). The higher the surface roughness, the more the nanoscale cavities that can act as nascent sites for nucleation to enhance the heat transfer rate [43].
To prove the cumulative effects of TiO2 nanoparticles in PDMS upon photoexcitation, Raman (WITec, Alpha300 R) and UV–Vis (Jasco, V-770) absorbance characterizations were conducted on the nanocomposite films. The confocal Raman microscope was used for the non-destructive imaging analysis of the samples, wherein the white light LED acted as the source for Köhler illumination, and a laser wavelength of 532 nm was employed. After dropping 100 µL of the precursor solution, respectively, for pristine PDMS, 0.1, and 0.25 wt% of TiO2-PDMS nanocomposites on the PET substrate, each dropped solution was used to fabricate the film with a thickness of 7 µm by a spin coater at 1000 rpm for 1 min, and then dried at room-temperature. Raman spectra were observed by enhancing the material’s chemical fingerprint through the distinct physical stretching and vibrational modes due to the interaction between the PDMS matrix and TiO2 nanoparticles [44]. The spectra for pristine PDMS, 0.1 and 0.25 wt% of TiO2-PDMS nanocomposite were acquired using an excitation wavelength of 595 nm at room-temperature using a Raman spectrometer. The TiO2-PDMS showed a noticeable enhancement of Raman peak intensities, specifically Si–O stretching at 488 cm−1 since the spectrum intensity of PDMS can be enhanced by the induced scattering of nanoparticles (inset in Fig. 2f) [45]. In addition, the UV–Vis absorbance of the TiO2-PDMS reactor upon the addition of nanoparticles was determined to see whether the wavelength range of emitted fluorescence from the PCR probes overlapped with the absorbance of the TiO2-PDMS reactor. The pristine PDMS does not absorb most of the light from 300 to 1000 nm of wavelength range, whereas the light absorption increased as the nanoparticles concentration in PDMS increased. Based on the attained absorbance and transmittance characteristics of pristine PDMS, 0.1 and 0.25 wt% TiO2-PDMS nanocomposite samples (Fig. 2g and S4), there were no serious overlapping issues between TiO2-PDMS reactor and the PCR probes. Furthermore, we can predict a correlation between light absorption and heat transfer to the surrounding medium by the nanoparticles. The nanoparticles' higher incident light absorbance, as shown in Fig. 2g, indicates enhanced heat transfer. Based on the Raman and UV–Vis studies, the emitted fluorescence intensity can be postulated to be enhanced by the reflection effect of TiO2 nanoparticles with the excitation and emission ranges of the PCR probes.
Thermal properties of TiO2-PDMS reactor
To analyze the changes in the heat transfer throughout the PCR reactor upon the addition of thermally conductive TiO2 nanoparticles in PDMS, computational models were used to optimize the thermal performance based on their thermal conductivity under various composition ratios. To calculate the thermal conductivity for different compositions of the TiO2-PDMS nanocomposites, the thermal diffusivity and heat capacity of the reactors were measured. For this analysis, pristine PDMS, 0.1 and 0.25 wt% of TiO2-PDMS nanocomposite blocks with a thickness of 1.5 mm were fabricated. The laser flash analysis (LFA) (NETZSCH, LFA467) was used as the thermal diffusivity measurement technique (Fig. 3a). Five random positions were chosen on the sample for the measurements, and the laser was irradiated in a direction normal to the plane of the sample, with 230 V and 0.3 ms of the pulse width. A differential scanning calorimeter (DSC) (NETZSCH, DSC214) was used to calculate the specific heat capacity of the samples, using a N2 purge gas flow rate of 20 mL/min at 24 °C. For three different PDMS systems, the heating and cooling rates were calculated using the 940 nm wavelength light which were optimized in previous studies [19], and the energy model with \(k\) = 0.27, 0.312 and 0.322 W/mK, respectively (Figures S5), as obtained from the thermal diffusivity analysis. Further, experiments were carried out using different fractions of nanocomposites to study the actual phenomena and observe the target gene amplification. A two-step thermal cycle was applied to analyze the amplification efficiency of pristine PDMS, 0.1 and 0.25 wt% of TiO2-PDMS nanocomposite reactors for the photonic qPCR. The PCR solution of 1 µL was assumed to be inside the reactor covered with 4 µL of mineral oil to prevent evaporation during the thermal cycles. As the LED is switched on, the temperature of the PCR mixture is reached to denaturation temperature (~ 93 °C), and the software detects and sends a signal to switch off the LED power automatically. The ON–OFF sequence continues for 40 thermal cycles. Initial denaturation occurs at 90–93 °C for 5 s, followed by denaturation at 91–93 °C, and annealing at 57–58 °C based on target amplicon size and primer conditions. The heating and cooling data obtained from the photonic PCR were compared with the numerical results, and the findings indicate that thermal cycling data corresponds well with the simulation results, with an error of ± 0.86 °C/s and ± 0.21 °C/s, respectively (Fig. 3b, c). The cooling rates increased to 0.25 wt% of the TiO2-PDMS nanocomposite reactor. However, after loading 0.25 wt% of TiO2, the heating and cooling rates reduced gradually (Figure S6) postulated due to the aggregation and non-uniform dispersion of the nanoparticles in the PDMS matrix. The total time taken for the target gene amplification with the fabricated 0.25 wt% of TiO2-PDMS nanocomposite reactor was approximately 100 s faster than the pristine PDMS reactor (Fig. 3d). Since the thermal conductivity increased as the weight percentage of the TiO2 nanoparticles increased, 40 thermal cycles were completed in 620 s leading to exponential amplification of the targeted sequences.
Comparison of a thermal conductivity and thermal diffusivities, numerical and experimental data (6 sets of samples at 3 different concentrations) for b heating and c cooling rates as a function of TiO2 loading ratio, and d thermal cycling data from the photonic PCR, for pristine PDMS and 0.25 wt% of TiO2-PDMS nanocomposite reactors; inset shows the comparison of each thermal cycling times for showing enhancing cooling ramp from each cycle
DNA amplification in TiO2-PDMS PCR reactor
As fluorescent dyes (SYBR Green) bind to the minor groove of double-stranded DNA (dsDNA) in the annealing step, they emit fluorescence when they are bound to dsDNA only. Thus, the fluorescence intensity increases proportionally as the number of amplicons increases [46, 47]. Therefore, it is essential to improve selective fluorescence signals in sequence-specific DNA detection and quantification [3, 48]. To enhance the fluorescence signal in the photonic qPCR, fluorescence intensities were compared for each sample: pristine PDMS, 0.1, 0.25, 0.5, and 0.75 wt% of TiO2-PDMS reactors in the photonic PCR. While performing qPCR, the fluorescence intensity values at every 5th cycle were compared through the fluorescence camera equipped with the filter. As the TiO2 nanoparticles surrounding the excited fluorophore possess a higher refractive index than pristine PDMS, the fluorescence intensity was increased (Figure S7). Also, the fluorescence brightness gradually enhances upon the addition of TiO2 nanoparticles in the pristine PDMS (Figure S7). This emitted fluorescence scintillates to TiO2 nanoparticles present at the matrix interface, which can reflect and further enhance the fluorescence [36]. Based on the experimental results (Fig. 3 and S6), the TiO2 composition was optimized at 0.25 wt% for fabricating the PCR reactors based on thermal cycling efficiency and distinct threshold cycle (Ct) value. The enhancement of fluorescence intensity via the TiO2-PDMS reactors was further proved by simply transferring 1 µL of the amplified PCR solution (1 copy/25 µL) using the benchtop qPCR (34, 36, 40 and 45 thermal cycles, respectively) into both the pristine PDMS and TiO2-PDMS reactors (Figure S8). The fluorescence intensity from the amplified samples of the 34 cycles was brighter in the TiO2-PDMS reactor while very dim fluorescence was observed from the pristine PDMS reactor.
The fluorescence enhancement using the TiO2-PDMS nanocomposite reactors in the photonic PCR was summarized in Fig. 4. The PCR reactor with and without TiO2 was simulated based on the refractive index using an electric field to confirm fluorescence signal enhancement, and the increase in reflectance when TiO2 is present in PDMS (Fig. 4a, c). The reflectance with TiO2 in PDMS doubled compared to without TiO2 (Fig. 4b) [49, 50]. Electric field intensity distribution and reflectance, as depicted in Fig. 4b, c, were computed utilizing the finite-difference time-domain (FDTD) method with ANSYS LUMERICAL (Version-2018a, ANSYS, Inc). The refractive indices for PDMS and the PCR solution are 1.4 and 1.33, respectively [51]. A linearly polarized plane wave source, with electric field oscillations along the z-direction and a spectral range from 450 to 600 nm, was employed for top-side illumination of the structure at normal incidence. Two power monitors were positioned, one behind the light source and the other in the y-direction of the structure plane, capturing reflectance signals and electric field distributions, respectively, at the same spectral interval and frequency points with respect to the source. All boundaries of the simulation region are set to perfectly matched layers.
Photonic PCR results by using pristine PDMS and 0.25 wt% of TiO2-PDMS PCR reactors. a A numerical model of electric field distribution in a PDMS reactor with the presence of the fluorescent dye and TiO2. The particle size is typically 200 nm, b variation in reflectance between PDMS and TiO2-PDMS nanocomposite reactors obtained from full-wave electromagnetic simulations, c intensity of electric field distribution as obtained from simulation in pristine PDMS and TiO2-PDMS nanocomposite reactor with the presence of fluorescence dye. d, e Fluorescence amplification graphs of the 10th, 20th, 30th, and 40th cycles with the 103 copies/µL of λ-DNA and NTC, f comparison of relative fluorescence intensities between the 5th and 40th cycle from 103 copies/µL of λ-DNA amplification (18 experiments were conducted) in pristine PDMS and TiO2-PDMS nanocomposite PCR reactors, respectively. g, h Normalized quantification graph of various target concentrations of pristine PDMS reactors and TiO2-PDMS reactors, respectively. i Standard curves of the Ct values versus the log concentration of λ-DNA. All the standard deviations are shown as error bars
To elucidate the underlying principles of fluorescent enhancement, the following equation (Eq. 2) is used, where fluorescence enhancement is denoted as \({\eta }_{F}\), the product of gains in excitation intensity enhancement as \({\eta }_{exc}\), quantum yield \({\eta }_{\phi }\), and collection efficiency \({\eta }_{coll}\). The fluorescence quantum yield quantifies the ratio of emitted photons to absorbed photons. Furthermore, the quantum yield gain can be expressed as the ratio between the gain in the radiative rate\(\eta_{\Gamma rad}\) and the total decay rate \(\eta_{\Gamma tot}\) [49].
To extract the enhancement of fluorescent signal in PDMS with and without TiO2 nanoparticles, the quantum yield gain should be around 1 \(\left( {\eta_{\Gamma rad} /\eta_{\Gamma tot} \sim 1} \right)\), due to the large size of the inverted frustum of the cone-shaped well [50]. Finally, the fluorescent enhancement equation can become \({\eta }_{F}={\eta }_{exc} {\eta }_{coll }\approx {(\frac{\left|E\right|}{\left|{E}_{0}\right|})}^{4}\), where \(E\) is the electric field amplitude in TiO2-PDMS and \({E}_{0}\) is the electric field amplitude in PDMS. The numerical simulation of reflectance spectra corresponding to the reactor model proves that the TiO2-PDMS shows higher relative reflectance and with the quantum yield gain as 1, the excitation light can be confined in PCR solution more effectively (Fig. 4b). The photothermal substrate used in the device can contribute to electromagnetic field enhancements, such as sacttering or local field effects, especially when part of a composite material.
To verify specific amplification, 40 thermal cycles of PCR involving λ-DNA target sample and negative template control (NTC) were performed and after completing the PCR reaction, the samples were collected and performed gel electrophoresis to confirm the absence of non-specific amplicons (Figure S9). In the case of 0.25 wt% of TiO2-PDMS nanocomposite PCR reactors, it was observed that the average fluorescence signal at the 10th and 40th cycle was up to 2 times brighter than the pristine PDMS reactors (Fig. 4d). On the other hand, in the NTC sample, no fluorescence signal change was observed from the 10th to the 40th thermal cycles (Fig. 4e). This suggests that the fluorescence enhancement effect is solely attributed by TiO2, without any other influencing factors. The 5th and 40th cycle fluorescence intensity is shown in Fig. 4f using the template concentration of 103 copies/µL in pristine PDMS and 0.25 wt% of TiO2-PDMS nanocomposite PCR reactors, respectively. During annealing, the fluorescent dye emits a stronger fluorescence, postulated to be reflected by TiO2 nanoparticles in the PDMS. Figure 4g, h shows the normalized intensity of PCR amplification curves for each concentration of template using pristine PDMS and 0.25 wt% of TiO2-PDMS nanocomposite reactors to characterize PCR amplification. The same volume of PCR solution with different concentrations (copies/µL) was successfully amplified and confirmed through gel electrophoresis (Figure S10). Figure 4i represents the Ct value according to the log concentration of λ-DNA with the existing reactor to confirm the reproducibility of experiments. The Ct value of 0.25 wt% of TiO2-PDMS reactors for 5 different λ-DNA concentrations (10 to 105 copies/µL) was about 2 times faster as compared to the pristine PDMS reactors. The attained enhancement effect in fluorescence signal can be explained based on the specular/diffuse reflectance of light [52, 53]. The polymers, in general, and PDMS have low values of refractive index, resulting in low reflectance and low fluorescence collection efficiency. TiO2 nanoparticles, on the other hand, have a high refractive index, making them ideal additives for applications involving enhanced fluorescence signal detection. Furthermore, the optical characteristics of TiO2 nanoparticles in the PDMS material were enhanced by efficiently reflected visible light, and improved light collection within the reactor. However, at a composition of above 0.5 wt% of TiO2 loading, the qPCR graph became close to linear, making it difficult to distinguish Ct. This suggests that more than 0.5 wt% of TiO2-PDMS nanocomposite reactor is not suitable for real-time fluorescence detection due to the low limit of quantitation (LoQ).
Conclusion
In this study, we developed a TiO₂-PDMS nanocomposite-based PCR reactor that significantly improves photonic qPCR for PoCT by enhancing both thermal cycling efficiency and fluorescence signal intensity. Employing Karstedt’s catalyst and PDMS-vinyl terminated, trimethylsilyl terminated poly(dimethylsiloxane-co-methylhydroxane) with TiO2 nanoparticles enables the rapid fabrication of TiO2-PDMS nanocomposite-based PCR reactors on R2R printed carbon-graphene-based photothermal conversion film.
The reactor was initially optimized depending on the thermal conductivity variation as the composition changes via a theoretical model. Also, the effect of the refractive index of TiO2 nanoparticles in PDMS was methodically considered in the simulation model to scrutinize the fluorescence signal enhancement. It was found that the addition of biocompatible and thermally conductive TiO2 nanoparticles into the PDMS-based PCR reactor not only improved the distribution of heat inside the PCR reactor during denaturation and annealing cycles to lessen the time for the amplification of the target gene but also enhanced the fluorescence detection during the assay, resulting in lower Ct values with lower assay time. The direct comparison between the pristine PDMS and TiO2-PDMS nanocomposite reactors proved that the assay time can be lowered by up to 100 s via two-step thermal cycling using the same PCR system [19]. In addition, the enhancement of fluorescence intensity in TiO2-PDMS nanocomposite PCR reactors was proved by comparing two samples’ fluorescence images after running 34 to 45 thermal cycles of 1 copy/25 µL sample using benchtop PCR. The Raman spectra peak intensities of TiO2-PDMS nanocomposite were enhanced by increasing the composition of nanoparticles, confirming the reflection effect of dispersed nanoparticles.
Although the proposed method doesn’t provide a drastic increase in speed, it demonstrates the potential for achieving faster PCR cycles by incorporating additives with good dispersibility in PDMS and higher thermal conductivity. Consequently, this simple fabrication approach for TiO₂-PDMS nanocomposite-based PCR reactors enable the mass production of PCR reactors for photonic qPCR devices, effectively reducing thermal cycling time while enhancing fluorescence signal intensity. This advancement contributes to the development of fast, sensitive, and cost-effective PoCT-like PCR for molecular diagnostics, particularly in the detection of viral infectious diseases.
Availability of data and materials
No datasets were generated or analysed during the current study.
Abbreviations
- qPCR:
-
Quantitative polymerase chain reactio
- PoCT:
-
Point-of-care testing
- SPR:
-
Surface plasmon resonance
- PDMS:
-
Polydimethylsiloxane
- PET:
-
Polyethylene terephthalate
- DEPC:
-
Diethyl pyrocarbonate
- PEG:
-
Polyethylene glycol
- NIR:
-
Near infrared
- LED:
-
Light-emitting diode
- PWM:
-
Pulse width modulatio
- FE-SEM:
-
Field emission scanning electron microscope
- EDS:
-
Energy dispersive spectroscopy
- AFM:
-
Atomic force microscopy
- RMS:
-
Root mean square
- LFA:
-
Laser flash analysis
- DSC:
-
Differential scanning calorimeter
- Ct :
-
Cycle threshold
- FDTD:
-
Finite-difference time-domain
- NTC:
-
Negative template control
- LoQ:
-
Limit of quantitation
References
Blumenfeld NR, Bolene MAE, Jaspan M, Ayers AG, Zarrandikoetxea S, Freudman J, et al. Multiplexed reverse-transcriptase quantitative polymerase chain reaction using plasmonic nanoparticles for point-of-care COVID-19 diagnosis. Nat Nanotechnol. 2022;17:984–92.
Guo M, Xia Y, Chen F, Hao L, Xin Y, Liu B. Development of an efficient dye-based qPCR system still functional for low levels of transgenic DNA in food products. Food Anal Methods. 2023;16:445–58. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s12161-022-02408-x.
Mackay IM, Arden KE, Nitsche A. Real-time PCR in virology. Nucleic Acids Res. 2002;30:1292–305.
Khodakov D, Li J, Zhang JX, Zhang DY. Highly multiplexed rapid DNA detection with single-nucleotide specificity via convective PCR in a portable device. Nat Biomed Eng. 2021;5:702–12. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41551-021-00755-4.
Valera E, Jankelow A, Lim J, Kindratenko V, Ganguli A, White K, et al. COVID-19 point-of-care diagnostics: present and future. ACS Nano. 2021;15:7899–906.
Mohammadyousef P, Paliouras M, Trifiro MA, Kirk AG. Plasmonic and label-free real-time quantitative PCR for point-of-care diagnostics. Analyst. 2021;146:5619–30.
Zhu H, Zhang H, Ni S, Korabečná M, Yobas L, Neuzil P. The vision of point-of-care PCR tests for the COVID-19 pandemic and beyond. TrAC Trends Anal Chem. 2020;130:115984.
Kang T, Lu J, Yu T, Long Y, Liu G. Advances in nucleic acid amplification techniques (NAATs): COVID-19 point-of-care diagnostics as an example. Biosens Bioelectron. 2022;206:114109.
Yin J, Tong J, Li J, Shao G, Xie B, Zhuang J, et al. A portable, high-throughput real-time quantitative PCR device for point-of-care testing. Anal Biochem. 2023;674:115200. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.ab.2023.115200.
Kang BH, Jang KW, Yu ES, Na H, Lee YJ, Ko WY, et al. Ultrafast plasmonic nucleic acid amplification and real-time quantification for decentralized molecular diagnostics. ACS Nano. 2023;17:6507–18.
Hong SH, Il ShuJ, Wang Y, Baysal O. Automated optimization of double heater convective polymerase chain reaction devices based on CFD simulation database and artificial neural network model. Biomed Microdevices. 2021. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s10544-021-00551-6.
Lee SH, Song J, Cho B, Hong SG, Hoxha O, Kang T, et al. Bubble-free rapid microfluidic PCR. Biosens Bioelectron. 2019;126:725–33.
Zhang J, Yang Z, Liu L, Zhang T, Hu L, Hu C, et al. Ultrafast nucleic acid detection equipment with silicon-based microfluidic chip. Biosensors. 2023;13:1–12.
Nunez-Bajo E, Collins ASP, Kasimatis M, Cotur Y, Asfour T, Tanriverdi U, et al. Disposable silicon-based all-in-one micro-qPCR for apid on-site detection of pathogens. Nat Commun. 2020. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41467-020-19911-6.
Ke C, Berney H, Mathewson A, Sheehan MM. Rapid amplification for the detection of Mycobacterium tuberculosis using a non-contact heating method in a silicon microreactor based thermal cycler. Sens Actuators B Chem. 2004;102:308–14.
You M, Cao L, Xu F. Plasmon-driven ultrafast photonic PCR. Trends Biochem Sci. 2020;45:174–5. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.tibs.2019.11.007.
Jalili A, Bagheri M, Shamloo A, Kazemipour Ashkezari AH. A plasmonic gold nanofilm-based microfluidic chip for rapid and inexpensive droplet-based photonic PCR. Sci Rep. 2021;11:1–13. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41598-021-02535-1.
Cheong J, Yu H, Lee CY, Lee J, Choi HJ, Lee JH, et al. Fast detection of SARS-CoV-2 RNA via the integration of plasmonic thermocycling and fluorescence detection in a portable device. Nat Biomed Eng. 2020;4:1159–67. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41551-020-00654-0.
Shrestha K, Kim S, Han J, Florez GM, Truong H, Hoang T, et al. Mobile efficient diagnostics of infectious diseases via on-chip RT-qPCR: MEDIC-PCR. Adv Sci. 2023;2302072:1–12.
Cunha J, Guo TL, Della Valle G, Koya AN, Proietti Zaccaria R, Alabastri A. Controlling light, heat, and vibrations in plasmonics and phononics. Adv Opt Mater. 2020. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/adom.202001225.
Chen J, Ye Z, Yang F, Yin Y. Plasmonic nanostructures for photothermal conversion. Small Sci. 2021;1:2000055.
Kim M, Lee JH, Nam JM. Plasmonic photothermal nanoparticles for biomedical applications. Adv Sci. 2019. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/advs.201900471.
Zhang Y, He S, Guo W, Hu Y, Huang J, Mulcahy JR, et al. Surface-plasmon-driven hot electron photochemistry. Chem Rev. 2018;118:2927–54.
Lee JH, Cheglakov Z, Yi J, Cronin TM, Gibson KJ, Tian B, et al. Plasmonic photothermal gold bipyramid nanoreactors for ultrafast real-time bioassays. J Am Chem Soc. 2017;139:8054–7.
Li TJ, Chang CM, Chang PY, Chuang YC, Huang CC, Su WC, et al. Handheld energy-efficient magneto-optical real-time quantitative PCR device for target DNA enrichment and quantification. NPG Asia Mater. 2016;8:e277.
Seo SE, Kim KH, Kim SJ, Ko KC, Kim WK, Lee KG, et al. Plasmon-driven gold nanopillar multiarrayed gene amplification methodology for the high-throughput discrimination of pathogens. Adv Sci. 2025;2411849:1–9.
Miranda I, Souza A, Sousa P, Ribeiro J, Castanheira EMS, Lima R, et al. Properties and applications of PDMS for biomedical engineering: a review. J Funct Biomater. 2022;13:2.
Yi P, Awang RA, Rowe WST, Kalantar-Zadeh K, Khoshmanesh K. PDMS nanocomposites for heat transfer enhancement in microfluidic platforms. Lab Chip. 2014;14:3419–26.
Huang C, Qian X, Yang R. Thermal conductivity of polymers and polymer nanocomposites. Mater Sci Eng R Rep. 2018;132:1–22. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.mser.2018.06.002.
He X, Huang Y, Wan C, Zheng X, Kormakov S, Gao X, et al. Enhancing thermal conductivity of polydimethylsiloxane composites through spatially confined network of hybrid fillers. Compos Sci Technol. 2019;172:163–71. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.compscitech.2019.01.009.
Lee DK, Yoo J, Kim H, Kang BH, Park SH. Electrical and thermal properties of carbon nanotube polymer composites with various aspect ratios. Materials. 2022;15:1356.
Abdul Khaliq R, Sonawane PJ, Sasi BK, Sahu BS, Pradeep T, Das SK, et al. Enhancement in the efficiency of polymerase chain reaction by TiO2 nanoparticles: crucial role of enhanced thermal conductivity. Nanotechnology. 2010;21:255704.
Amadeh A, Ghazimirsaeed E, Shamloo A, Dizani M. Improving the performance of a photonic PCR system using TiO2 nanoparticles. J Ind Eng Chem. 2021;94:195–204. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.jiec.2020.10.036.
Li R, Wang JX, Lee E, Luo T. Physics-informed deep learning for solving phonon Boltzmann transport equation with large temperature non-equilibrium. NPJ Comput Mater. 2022;8:1–10.
Prasher RS. Mie scattering theory for phonon transport in particulate media. J Heat Transfer. 2004;126:793–804.
Eyer K, Root K, Robinson T, Dittrich PS. A simple low-cost method to enhance luminescence and fluorescence signals in PDMS-based microfluidic devices. RSC Adv. 2015;5:12511–6.
Pina-Hernandez C, Kim JS, Guo LJ, Fu PF. High-throughput and etch-selective nanoimprinting and stamping based on fast-thermal-curing poly(dimethylsiloxane)s. Adv Mater. 2007;19:1222–7.
Hiltunen J, Liedert C, Hiltunen M, Huttunen OH, Hiitola-Keinänen J, Aikio S, et al. Roll-to-roll fabrication of integrated PDMS-paper microfluidics for nucleic acid amplification. Lab Chip. 2018;18:1552–9.
Chen Y, Pei W, Tang R, Chen S, Chen H. Conformal coating of parylene for surface anti-adhesion in polydimethylsiloxane (PDMS) double casting technique. Sens Actuators A Phys. 2013;189:143–50. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.sna.2012.09.024.
Tiara AM, Chakraborty S, Sarkar I, Pal SK, Chakraborty S. Effect of alumina nanofluid jet on the enhancement of heat transfer from a steel plate. Heat Mass Transfer/Waerme- und Stoffuebertragung. 2017;53:2187–97.
Hu W, Jia Y, Chen X, Le Q, Chen L, Chen S. Determination of secondary cooling zone heat transfer coefficient with different alloy types and roughness in DC casting by inverse heat conduction method. Crystals. 2022;12:1571.
Cabrera JN, Ruiz MM, Fascio M, D’Accorso N, Minchev R, Dubois P, et al. Increased surface roughness in polydimethylsiloxane films by physical and chemical methods. Polymers. 2017;9:331.
Može M, Hadžić A, Zupančič M, Golobič I. Boiling heat transfer enhancement on titanium through nucleation-promoting morphology and tailored wettability. Int J Heat Mass Transf. 2022;195:123161.
Mariyappan K, Tandon A, Park S, Kokkiligadda S, Lee J, Jo S, et al. Nanomaterial-embedded DNA films on 2D frames. ACS Appl Bio Mater. 2022;5:2812–8.
Li X, Liu H, Gu C, Zhang J, Jiang T. PDMS/TiO2/Ag hybrid substrate with intrinsic signal and clean surface for recyclable and quantitative SERS sensing. Sens Actuators B Chem. 2022;351:130886. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.snb.2021.130886.
Yu Z, Xu Q, Xiao C, Li H, Wu W, Du W, et al. SYBR Green real-time qPCR method: Diagnose drowning more rapidly and accurately. Forensic Sci Int. 2021;321:110720. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.forsciint.2021.110720.
Lekanne Deprez RH, Fijnvandraat AC, Ruijter JM, Moorman AFM. Sensitivity and accuracy of quantitative real-time polymerase chain reaction using SYBR green I depends on cDNA synthesis conditions. Anal Biochem. 2002;307:63–9.
Mackay IM. Real-time PCR in the microbiology laboratory. Clin Microbiol Infect. 2004;10:190–212. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/j.1198-743X.2004.00722.x.
Barulin A, Claude JB, Patra S, Bonod N, Wenger J. Deep ultraviolet plasmonic enhancement of single protein autofluorescence in zero-mode waveguides. Nano Lett. 2019;19:7434–42.
Barulin A, Roy P, Claude JB, Wenger J. Ultraviolet optical horn antennas for label-free detection of single proteins. Nat Commun. 2022;13:1–9.
Liao P, Jiang M, Chen Z, Zhang F, Sun Y, Nie J, et al. Three-dimensional digital PCR through light-sheet imaging of optically cleared emulsion. Proc Natl Acad Sci U S A. 2020;117:25628–33.
Takatani T, Mukaigawa Y, Matsushita Y, Yagi Y. Decomposition of reflection and scattering by multiple-weighted measurements. IPSJ Trans Comput Vis Appl. 2018. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s41074-018-0049-4.
Tao P, Li Y, Rungta A, Viswanath A, Gao J, Benicewicz BC, et al. TiO2 nanocomposites with high refractive index and transparency. J Mater Chem. 2011;21:18623–9.
Acknowledgements
We thank Mr. Samanth Kokkiligadda of SAINT, SKKU, for his helpful support in the Raman analysis. We also thank Mr. Trung Hoang of IQB, SKKU, for his helpful assistance in the PCR reactor fabrication method.
Funding
This work is supported by the Korea Medical Device Development Fund grant funded by the Korean government (The Ministry of Science and ICT, the Ministry of Trade, Industry and Energy, the Ministry of Health & Welfare, the Ministry of Food and Drug Safety) (Project Number: RS-2020-KD000004) and the National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT) (No. 2020R1A5A1019649).
Author information
Authors and Affiliations
Contributions
‡J.H., T.A.M., and S.K. contributed equally to this work. J.H., T.A.M., and S.K. carried out the data curation, formal analysis, and original draft writing. G.M.F., K.S., D.D.N., and I.K. participated in the data curation. *J.L. and G.C. conceived and supervised the project. All the authors discussed the results and commented on the manuscript.
Corresponding authors
Ethics declarations
Ethics approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Competing interests
The authors declare no competing interests.
Additional information
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary Information
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, 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 you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. 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-nc-nd/4.0/.
About this article
Cite this article
Han, J., A M, T., Kim, S. et al. Nanocomposite-based PCR reactors to enhance thermal rate and fluorescence intensity in hand-held qPCR device. J Nanobiotechnol 23, 240 (2025). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12951-025-03287-0
Received:
Accepted:
Published:
DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12951-025-03287-0