We report a novel template method for synthesizing silica nanoparticles that are mesoporous as well as biocompatible. The mesoporous silica nanoparticles were synthesized using the Stober process and peptization method. We have used surface-modified deoxyribonucleic acid (DNA) with cetyl trimethyl ammonium bromide as a capping agent. The obtained silica particles were analyzed using x-ray diffraction, UV–vis spectroscopy, the Brunauer–Emmet–Teller method, scanning electron microscopy, and the open aperture Z-scan technique. The characterization results indicated that the DNA polymer’s presence influenced the formation of the silica particles. The silica particles are mesoporous, nanosized, and good adsorbent and also show enhanced non-linear optical properties when compared with existing silica nanoparticles. The solubility of the silica nanoparticles was also verified in dimethyl sulfoxide (DMSO). We have verified that the enhanced mesoporous surface area and reduced porous size of the silica nanoparticles influenced the photoluminescence of Rhodamine 6G dye in DMSO. This can be applied to lower the lasing threshold of the gain medium in lasing applications. Hence, the obtained silica nanoparticles have a variety of applications such that they can be used as adsorbents of nanosized particles, which is applicable for drug delivery purposes, bioimaging, catalytic activities, doping surface for thin film making, biosensing applications, and improvement of material quality for solar devices. The non-linear optical property of the mesoporous silica nanoparticles can be used for optical limiting applications in photonic devices.

The emergence of silica nanoparticles has been brought to attention in material chemistry since the 1970s, soon after the development of nanotechnology. In general, silica is defined as transparent, odorless, vitreous luster, non-conductive to electricity, diamagnetic, and amorphous. These properties of silica limit their roles in certain applications, such as a substrate for thin film making, fiber optics, ceramics, and semiconductors. Since the invention of the Stober process in 1968, it is proven that nano-level, controllable spherical silica particles can be achieved. Subsequently, various synthesis methods were reported to obtain controllable and uniform-sized silica nanoparticles in material science for different industrial and chemical aspects. Moreover, the material research field realized that silica nanoparticles were undergone surface modification under different experimental conditions and acquired adsorptive and non-linear optical properties. The morphologically modified silica particles are specified as mesoporous nanoparticles in which the pore size falls to 1–10 nm. In addition, they offer a porous structure, tunable size, large surface area, biocompatibility, and ordered uniform structure.1–18,20 This awakened the material chemistry in a way that silica can be subject to various usages due to its enhanced morphological adaptations and non-linear optical features. Over the last few decades, various chemical and physical methods were introduced to enhance the properties of silica nanoparticles at different morphological and functional levels. Based on the easiness of production and physio-chemical features, nowadays silica nanoparticles became one of the class materials, which are known for their different roles in the chemical industry, such as nanomedicine, catalysis, plasmonic color thin films, surface-enhanced spectroscopy, deoxyribonucleic acid (DNA) extraction, magnetic separation, and photonic devices and instrumentation development. Recently, porous nanomaterials are widely used for tuning laser systems and cell imaging purposes. Among the porous nanomaterials, silica nanoparticles are treated as a favorite subject in material science, so silica-incorporated research works become a focus in tuning dye laser systems.19 These applications of silica are merely based on its unique mesoporous surface properties. The surface properties of silica nanoparticles can be modified for different purposes. These features of the mesoporous silica nanoparticles help to attach the functional chemical groups on them in comparison with amorphous colloidal silica. The reported research papers revealed that mesoporous silica nanoparticles can influence the structural, microstructural, and biological properties of polymer nanocomposites. Functionalized mesoporous silica nanoparticles are widely recognized for their stable chemical properties, controlled bioactive compound release, adsorbent-adsorbate stability, thermal stability, control of hydrophilicity, and pH resistance. The surface properties of silica nanoparticles can be enhanced by altering various parameters, such as temperature, pressure, the molar ratio of reagents, pH, and chemical additives as well. The presence of chemical additives during the precipitation of silica nanoparticles can control the growth of nanoparticles within the mesoporous limit along with the surface modification of other parameters. Since the features of silica nanoparticles in terms of porosity, size of the pores, pore volume, surface area, and biocompatibility are found to enhance by the presence of chemical additives, different methods and chemical additives have been reported.1–18,20 In 2009, Qiao et al.21 reported the synthesis of mesoporous silica nanoparticles through a controlled hydrolysis process using cetyl trimethyl ammonium chloride (CTAC). In this article, they claimed that they could control the size of the silica nanoparticles to a range of 25 to 200 nm. In 2012, Lodha et al.22 stated that they have synthesized mesoporous silica nanoparticles for drug loading of the purely water-soluble drug. They used cetyl trimethyl ammonium bromide (CTAB) and concentrated hydrochloric acid (HCl) for the synthesis of mesoporous silica nanoparticles. In 2017, Vazquez et al.23 claimed that they could achieve to form silica particles having pore volume sizes of 2.5 to 2.8 nm. They have used CTAB as the pore-generating agent. In 2020, Presentato et al.24 stated that they prepared biodegradable, high-pore volume mesoporous silica nanoparticles using CTAB and concentrated HCl. These reported silica particle synthesis methods were carried out using the Stober process method. Such mesoporous silica nanoparticles are applied to the biomedical field for target drug delivery, bio-sensing, and cellular uptake. Since their role in the bio-medical field is found to increase, the recent works mainly focus on making mesoporous silica to be functionalized and biocompatible using different surfactants, such as cetyl trimethyl ammonium chloride (CTAC).1–32 

This work introduces a biotemplate to prepare mesoporous silica nanoparticles by incorporating the Stober process. Among the available reported synthesis methods of silica nanoparticles, we have approached an effective bio-template to obtain tailored morphology of silica particles. The bio-template was synthesized by forming a complex of deoxyribonucleic acid and the cationic surfactant cetyltrimethyl ammonium bromide (CTAB). The crucial roles surfactants play in nanoparticle formation are described by adsorbing the nanoparticles to their surface and lowering the surface energy of the nanoparticles soon after they are formed. This is the way in which the aggregation of particles has been prevented and the formation of size and surface controlled nanoparticles. So far, the stabilizing mechanism of DNA in nanoparticle formation has been recognized, as its usage is being increased in material science as a bio-template and conjugate polymer as well. Studies show that the nanoparticles that are formed in the presence of a DNA template exhibit excellent physiochemical properties.33–61  Figure 1 illustrates a schematic representation of the polynucleotide structure of DNA. Since the DNA is considered as a fascinating biophotonic material, it is verified in many research papers that it can improve the functional properties of photonic device materials. Moreover, the seductive factor that makes DNA to be a vital material in material science is the double helical-lengthy polymer structure of the molecule. Due to the unique double helical structure, DNA possesses high transparency, thermal stability, non-linear optical activity, and amplified emission, its applications are widely found in electron blocking, hole transport, hosting of laser dyes, modification capability, electrical interconnectors, thin films, lasing, and sensing purposes. The strands of the double helical structure have been made of alternating sugar (deoxy ribose) and phosphate groups. The sugar group is attached to one of the four bases: adenine (A), cytosine (c), guanine (G), or thymine (T). The stability of the dsDNA molecule depends on GC base pairs and length. It is found that longer molecules are more stable. Long DNA helices with a high GC-base pairs have more strongly interacting strands, while short helices with high AT content have more weakly interacting strands. The structural features that defines a DNA polymer to engage for a reaction are length, base sequence, temperature, and buffer condition.33–60 

FIG. 1.

Portion of polynucleotide chain of deoxyribonucleic acid (DNA).

FIG. 1.

Portion of polynucleotide chain of deoxyribonucleic acid (DNA).

Close modal

As far as concerning the diverse properties of DNA, it is found that they can be incorporated with nanoparticles, which have positive charge. Mostly, DNA is used for functionalizing with metal nanoparticles to form functionalized or conjugated polymer under a buffer condition. This confirmed the negative charge of the DNA that resembles the chemical behavior of ligand. The studies dealt with its conjugating effect as they are taking part in a chemical reaction with other polymers (surfactants, metal nanoparticles, etc.) they got functionalized, so that they can form organic–inorganic complexes and salt with the polymers. In turn, the functionalized DNA has diverse properties that it can tune nanomaterials about their size, shape, surface compatibility with adsorbents, electrical, magnetic, optical, and biological properties.56–61 

The critical disadvantage of the DNA polymer, which prevents it from using directly in the functionalization of silica, is that the solvent used to synthesis silica nanoparticles in the Stober process is an alcohol medium (either ethanol or methanol). As DNA alone cannot readily dissolve in alcoholic solvents, it is required to transform DNA into its complex form with a cationic surfactant. Considering the properties of the functionalized DNA polymer, we selected cetyl trimethyl ammonium bromide (CTAB). It is also used for synthesizing mesoporous silica nanoparticles. It is a known cationic surfactant and belongs to one of the compounds of the cetrimonium cation family. The fascinating feature of CTAB is that it is widely used as the main component of the buffer solution for the extraction of DNA so that it reacts with the polysaccharide chains of the DNA.16,19,61–63

Its chemical formula is hexadecyltrimethyl ammonium bromide ([C16H33)N(CH3)3]Br) and is highly biocompatible. Figure 2 is a schematic representation of the CTAB structure. It is a familiar compound to be used for nanoparticle synthesis. It has 16 carbon as a long tail and an ammonium head group with three methyl groups attached. In some cases, functionalized biomolecules, such as DNA, cannot sufficiently penetrate a dense layer of capping agents or replace them. Therefore, surfactant-assisted methods aim about modifying the surface ligands by adding excess of small molecules with higher affinity toward dielectric nanoparticles. In the case of the DNA molecule, the deoxyribose-phosphate groups contribute to the backbone of the DNA strand. Among these groups, the phosphate group comprises one negatively charged oxygen atom and is responsible for the negative charge of the entire DNA strand. It is the result of the presence of bonds between phosphorous and oxygen. Hence, the cationic surfactant becomes attached to the phosphate group of the DNA and forms the DNA–surfactant complex.20–22,30–32

FIG. 2.

Chemical structure of CTAB.

FIG. 2.

Chemical structure of CTAB.

Close modal

Meeting the features of CTAB and DNA together, we proposed CTAB as the cationic surfactant for synthesizing the CD complex without using any buffer chemical. Silica is a well-known substrate material, and vast synthesis methods of mesoporous silica are available. Majority of the applications of mesoporous silica are broadly spanned in medical purposes, especially in the drug delivery system, binding of biomolecule for extraction of cell organelles, and cell imaging purposes. Therefore, functionally modified silica should be biocompatible and has to possess tailored morphological properties. Keeping this fact, the relevance of this work among other synthesis methods of silica is that this method proposes very simple, cost effective, and functionally modified silica with the DNA polymer without using any buffer solutions. By excluding buffer solutions, we can avoid toxic chemicals during the synthesis process and make the experimental conditions free from the toxic environment. It is found that the CD complex can dissolve in alcohol group. The obtained mesoporous silica nanoparticles have shown an enhanced tailored morphology along with the controllable adhesive property. It is also proven that the CD bio-complex template was an effective template to incorporate with the Stober process to enhance the particle size, pore size, volume, and area of the silica particles. The fluorescence property of Rhodamine 6G (R6G) dye particles was studied by encapsulating them with obtained silica nanoparticles in dimethyl sulfoxide (DMSO). The study shows that the fluorescence of dye particles was enhanced with the presence of CD-functionalized silica nanoparticles. The non-linear optical properties of dielectric materials with metal composites were investigated since two decades ago for fabricating non-linear optical devices. In those studies, the non-linear optical properties of metal nanoparticles doped with silica glass substrate were only considered.64 This pointed that the non-linear optical property of mesoporous silica nanoparticles alone is seemed ignorant in non-linear optical physics.65 Even though, silica glass inclusion with semiconducting materials and metal nanoparticles in order to enhance the non-linear optical properties of the doping materials is still maintained as a research interest. DNA is a well-known non-linear optical material, and the nanoparticles, which are functionalized with DNA, also possess nonlinear optical properties. Hence, along with the morphological analysis, we have also verified the non-linear optical properties of the CD-functionalized silica nanoparticles using the Z-scan technique.

Chemicals used were CTAB, DNA, tetra ethyl orthosilicate (TEOS), ammonium hydroxide (NH4OH), methanol, hydrochloric acid (HCL), acetone and deionized (DI) water. All chemicals except DNA were purchased from Merck Life Science Private Limited. DNA powder was purchased from Sisco Research Private Limited. All the glass wares used for the experiment were cleansed by acetone, followed by DI water, dried in hot air oven at 100 °C, and were again subject to Piranha cleaning followed by DI water cleansing and dried at 100 °C using hot air oven.

We prepared the complex of DNA and the cationic surfactant CTAB with different concentration of DNA solution and kept the concentration of CTAB constant. The concentration of DNA solutions were taken as 0.1, 0.5, 0.15, and 0.1525 of weight-percentage (wt. %), and they were prepared by dissolving DNA in DI water followed by sonication at room temperature for 30 min to achieve homogeneous solution. In another beaker, 0.4 g of CTAB was taken, and 30 ml of DI-water was poured into the beaker and sonicated until the CTAB was completely dissolved at room temperature. It was a colorless viscous solution. The CTAB solution was added slowly into the DNA solution under constant stirring using sonication at room temperature. This process allowed to precipitate a white colloidal form of a CD complex where the surfactant CTAB attached with the DNA molecule. The precipitation was filtered and washed with DI water in many times to remove unreacted CTAB. The CD complex precipitate was kept in a hot air oven at 60 °C for 24 h to obtain CD complex powder.48,49 Figure 3 shows a schematic representation of CD complex formation.

FIG. 3.

Schematic diagram of CD complex formation.

FIG. 3.

Schematic diagram of CD complex formation.

Close modal

The silica nanoparticles were synthesized in the presence of the CD complex under the conditions of the Stober process.1 The samples of 0.01, 05, 0.15, and 0.1525 wt. % of CD complex powder were dissolved in each of 150 ml of methanol under constant stirring. 35 ml of DI water was added into each four CD complex-methanol solutions under constant stirring for 10 min at room temperature. 5 ml of tetraethyl orthosilicate solution was added into these solutions under constant stirring using ultra sonication at room temperature in a 500 ml borosilicate glass beaker. 14 ml of ammonium hydroxide was added as catalyst slowly into the solution under constant stirring. The solution was kept stirring for 30 min at room temperature, and a white colloidal precipitation was formed during stirring. The colloidal form was kept for one hour for complete gelation. The precipitation is filtered by waterman filter paper and washed 5 times with DI water followed by methanol. The filtered precipitation was kept in a hot air oven at 70 °C for 1 h for drying. Then, the obtained silica powder was calcined at 10 °C/min. In addition, we have performed different experimental conditions for observing the effect of CD complex on making silica nanoparticles. Figure 4 shows the different methods involved in the formation of silica nanoparticles. One of them was carried out without the CD complex. In order to distinguish the roles of DNA and CTAB individually, one sample of silica nanoparticles was synthesized using CTAB as the surfactant.19 The Stober process can be performed in both alkaline and acidic medium. Therefore, the reliability of the CD complex were examined in both acidic and alkaline conditions of the Stober process. For acidic condition, we have used concentrated HCl. During the preparation of silica nanoparticles using the Stober process, 5 ml of concentrated HCl was added immediately after the colloidal form of silica gel is formed. This method is called the peptization process, and it is generally used for making mesoporous silica nanoparticles. Figure 5 illustrates a schematic representation of peptization method, which is performed in both the absence and presence of the CD complex.

FIG. 4.

Schematic representation of formation of silica nanoparticles via (a) using CTAB (b) using the Stober process without any surfactant and (c) using the CD complex.

FIG. 4.

Schematic representation of formation of silica nanoparticles via (a) using CTAB (b) using the Stober process without any surfactant and (c) using the CD complex.

Close modal
FIG. 5.

Schematic representation of formation of silica nanoparticles via the (a) peptization method and (b) peptization method using the CD complex as the stabilizing agent.

FIG. 5.

Schematic representation of formation of silica nanoparticles via the (a) peptization method and (b) peptization method using the CD complex as the stabilizing agent.

Close modal
The two differently prepared powder samples of silica were subject to XRD analysis. One of the samples was synthesized in the presence of 0.1525 wt. % of the CD complex (S1), and the remaining sample was prepared in the absence of CD complex template to examine the effects of CD complex in making nanoparticles. The XRD patterns of S1 and S2 are provided in Fig. 6. From the XRD pattern, we determined the peak intensity, position, and full width at half maximum (FWHM) data using the Gaussian fitting method. The diffraction angle’s 2θ range was taken between 20° and 80°. The XRD patterns for both samples show amorphous nature by the single broad peak, which ratifies that the sample is silica by forming the broad peak at 22.61° for S1 and 23.53° for S2, which are in the range of amorphous silica found generally in the XRD pattern for silica particles.68,69 The diffraction peak obtained for S1 is wider than that of S2. The average size of the silica particles of the two samples are estimated using the Debeye–Sherrer formula, and they are 1.65 and 2.63 nm for S1 and S2, respectively,70–73 
(1)
where 0.89 is the shape factor, λ is the x-ray wavelength = 1.540 56 Å for copper Kα x ray, b is the line broadening at FWHM in radians, and θ is the Bragg angle.
FIG. 6.

XRD of silica nanoparticles formed in the presence of CD complex (S1) and silica particles formed in the absence of CD complex (S2).

FIG. 6.

XRD of silica nanoparticles formed in the presence of CD complex (S1) and silica particles formed in the absence of CD complex (S2).

Close modal

Figure 6 shows the difference between the FWHM of S1 and S2 due to the presence of CD complex template. When comparing the XRD patterns of S1 and S2, the diffraction peak of S1 tends to move lower angle and getting narrower than the peak of S2. These features of S1 indicate that the silica particles may have crystalline structure or better orientation.35 From the FWHM difference, it is understood that the size of the particles of S1 is smaller than as that of S2. The XRD results validate that the presence of CD complex template influences the size of the silica particles to reduce at the nano-scale range.

The optical absorption spectrum of the samples were carried out using UV–vis spectrophotoscopy in the scan range of 230–800 nm wavelength. The silica samples subjected for UV–vis analysis were S1, S2, S3, S4, S5, S6, S7, and S8 corresponding to silica prepared by the Stober process, the Stober process using CTAB as the stabilizing agent, 0.01, 0.05, 0.15, and 0.1525 wt. % of DNA concentrations of the CD complex in the Stober process, and the peptization process with 0.1525 wt. % of the CD complex template, respectively. The amount of CTAB used for synthesizing sample S2 is approximately the same with the amount of CTAB used for preparing samples S7 and S8. This experimental condition is maintained for differentiating the background role of CTAB in the CTAB incorporated Stober process, CD complex template involving Stober and peptization processes for forming the nanoparticles. The solvent used for UV–vis analysis was dimethyl sulphoxide (DMSO). Figure 7 represents the UV absorption spectra of silica samples. It illustrates that the absorption peaks for silica prepared in the CD complex are showing blue shift (samples S2–S8). The peak blue shift for S8 is approximately about 10 nm. This indicated that as increasing the concentration of the CD complex template, the particle size goes on decreasing. This result exemplifies that the presence of CD complex reduces the size of the silica nanoparticles.

FIG. 7.

UV–vis absorption of silica nanoparticles prepared via the (S1) Stober process, (S2) Stober process incorporated with CTAB, (S3) peptization method, (S4) Stober process with 0.01 wt. % of the CD complex, (S5) Stober process 0.05 wt. % of the CD complex, (S6) Stober process with 0.15 wt. % of CD complex, (S7) Stober process with 0.1525 wt. % of the CD complex, and (S8) peptization process with 0.1525 wt. % of the CD complex.

FIG. 7.

UV–vis absorption of silica nanoparticles prepared via the (S1) Stober process, (S2) Stober process incorporated with CTAB, (S3) peptization method, (S4) Stober process with 0.01 wt. % of the CD complex, (S5) Stober process 0.05 wt. % of the CD complex, (S6) Stober process with 0.15 wt. % of CD complex, (S7) Stober process with 0.1525 wt. % of the CD complex, and (S8) peptization process with 0.1525 wt. % of the CD complex.

Close modal

The peak shift difference between S3 and S8 indicates that the CD complex can also make impact in acidic condition so that S3 and S8 are prepared under acidic medium.

Surface area is the one of the quantities to explain the particle size, particle morphology, surface texturing, and porosity of any particle. It is considered as the one of the vital factors, which explains the potential of physical properties of particles on taking part in biological and inorganic catalytic activities. To endorse the surface area of the silica nanoparticles derived from the CD complex templated method, BET analysis was conducted. Based on the BET analysis results, N2 adsorption/desorption isotherms were employed to identify surface parameters. We have subject six samples of silica particles for BET analysis. The samples were named as S1, S2, S3, S4, S5, and S6 corresponding to the silica particles of prepared by the Stober process, prepared by using 0.01, 0.05, 0.15, and 0.1525 wt. % in the peptization method and 0.1525 wt. % CD complex template. The BET results discussed the size of the pores, the type of pores whether it is open or closed, and the pore diameter distribution. The N2 adsorption/desorption isotherm curves of the six samples of SNP exhibit type IV hysteresis loop with two branches, adsorption (capillary condensation) and desorption (evaporation). These hysteresis loops are indicating the presence of open pores between the particles and mesoporous surface. Since the Stober process gives mesoporous silica nanoparticles, this general feature is supported from the hysteresis loops of the all the six samples.74 

The adsorption branch of type IV is composite of type I and II. At low relative pressure (p/p0), the uptake of gas molecules is associated with the filling of micropores (forms monolayer adsorption). These type IV loops are often found with aggregated crystals of zeolites, some mesoporous zeolites, and micro-mesoporous carbon. The isotherm curves of samples shown in Fig. 8 illustrated that the adsorption pore volume of the samples went increasing as going from sample S1 to S6 in accordance with the pore diameter, which shows a gradual increase on increasing the concentration of the CD complex template.

FIG. 8.

BET isotherm results of (a) silica nanoparticles synthesized in the Stober process, silica nanoparticles synthesized in (b) 0.01 wt. % of the CD complex template, (c) 0.05 wt. % of the CD complex template, (d) 0.15 wt. % of the CD complex template, (e) peptization method in 0.1525 wt. % of the CD complex template, and (f) 0.1525 wt. % of the CD complex template.

FIG. 8.

BET isotherm results of (a) silica nanoparticles synthesized in the Stober process, silica nanoparticles synthesized in (b) 0.01 wt. % of the CD complex template, (c) 0.05 wt. % of the CD complex template, (d) 0.15 wt. % of the CD complex template, (e) peptization method in 0.1525 wt. % of the CD complex template, and (f) 0.1525 wt. % of the CD complex template.

Close modal

The highest value of BET surface area is for sample S6. It is evident from the isotherm data shown in Table I of the samples that, for large pore size, the BET surface area is comparably found to low, which implicates that for low mesoporous properties, and such a surface cannot adsorb particle to their surface than particles having small pores on their surface. For small pores, the BET surface area and pore volume are large. This feature implies that their adsorption stability is high. This is evident from the isotherm curves of samples S1–S6 shown in Fig. 8. At higher relative pressure (p/p0), the hysteris loop width is quiet wider for samples S5 and S6 as comparing with other samples. The fact that connects the loop width and adsorption property is that the wider the loop, the adsorption stability is also higher. Moreover, the hysteresis loop width is found to increase as going from S1 to S6. The only factor that varies as going from S1 to S6 is the concentration of the CD complex template. Samples S5 and S6 were made in the presence of high concentration of the CD complex template (0.1525 wt. %) among other samples. Another fact obtained from the BET results is that the pore size of all the samples are within the mesoporous range (2 nm < pore diameter< 50 nm). Among the six samples, S6 exhibited a small value of mesoporous size by analyzing the size distribution of the pores. In fact, this pointed that the presence of CD complex induces to make the porous surface of the silica nanoparticles to become more adsorbent surface in which the adsorbents cannot be escaped easily. As we know, DNA is a double stranded lengthy biopolymer and CTAB is also a lengthy polymer. These together form a complex structure that is spanned between the silica particles. Due to the lengthy network of the CD complex, the silica particles cannot be aggregated beyond a particular range, and thus, nanoparticles are formed. The nanoparticles are confined in the CD complex surface, after the removal of the CD complex, which leads voids in the surface of the silica nanoparticles, which, in turn, caused the silica particles become mesoporous. This feature makes them to be a good adsorbent and that can be efficiently used for drug delivery agent, doping agents, and gas sensing applications.

TABLE I.

Experimental data obtained from N2 adsorption–desorption isotherms.

SampleMean pore diameter (nm)Pore volume (cm3/g)BET surface area (m2/g)
S1 7.942 0.019 9.543 
S2 6.532 0.029 21.496 
S3 6.379 0.035 22.686 
S4 6.188 0.070 45.516 
S5 5.063 0.086 54.172 
S6 5.946 0.081 54.335 
SampleMean pore diameter (nm)Pore volume (cm3/g)BET surface area (m2/g)
S1 7.942 0.019 9.543 
S2 6.532 0.029 21.496 
S3 6.379 0.035 22.686 
S4 6.188 0.070 45.516 
S5 5.063 0.086 54.172 
S6 5.946 0.081 54.335 

The SEM images of the silica particles are shown in Fig. 9. The silica particles prepared without the CD complex template is seen in Fig. 9(a). The particles are showing a spherical geometry with smooth surface. Figure 9(b) is representing the silica particles prepared in the presence of CD complex, in which the particles are seen in aggregated spherical and flat surfaces but irregular size. The surfaces of the particles were seemed to be irregular spherical structures, which were seen as projecting from the surface. These projected like spherical nanoparticles can enhance the surface area. Hence, the SEM results show that the presence of CD complex influences the morphology of the silica particles to a favorable feature of mesoporous surface, and this synthesis method could preserve the spherical shape of the silica particles, which is seen as a general feature of the Stober process.

FIG. 9.

SEM images of silica particles formed (a) without the presence of CD complex and (b) with the presence of CD complex.

FIG. 9.

SEM images of silica particles formed (a) without the presence of CD complex and (b) with the presence of CD complex.

Close modal

1. Non-linear absorption studies

DNA, the major component of the biotemplate CD complex, is a well-known non-linear optical material.34–37,42,55,58–61 It is proved that non-linear material can influence non-linear optical properties of another material when functionalizing each other.42,58–61 In order to understand whether any change has been formed in the non-linear optical property of the silica particles, we have conducted open aperture Z scan experiment. The open aperture Z-scan analysis was carried out under 100 µJ laser excitation (532 nm, 9 ns pulse rate) with an on-axis intensity of 2.46 GW cm−2. The transmitted beam was measured without aperture in front of the detector to determine the non-linear absorption of the molecules. The linear transmittance was set at 71%. Figures 10(a), 10(c), 10(e), 10(g), and 10(i) show the open aperture Z-scan signatures of silica particles prepared at different concentrations of the CD complex template in different methods. The silica particles was uniformly dispersed in DMSO for carrying out open aperture Z scan analysis. The interaction of the laser beam with the molecules produces a valley pattern, i.e., the transmittance of the molecules decreases gradually toward the focal point and reaches a minimum with the deep transmittance trough at the focus where the curves are symmetric at Z = 0, which signifies the reverse saturation absorption (RSA) behavior of the molecules with the positive NLA of the incident light. On the nanosecond time scale, the RSA is combined with two photon absorption (TPA) and excited state absorption (ESA), which is collectively as the effective TPA process.61,75–78 The experimental data are matching well with the theoretical model for the ESA assisted TPA process in all six samples of silica nanoparticles.

FIG. 10.

Z Scan curves of silica particles synthesized at (a) 0.05 wt. %. CD complex template and (c) 0.15 wt. %. CD complex template, (e) 0.1525 conc. CD complex template, (g) peptization method, and (i) 0.1525 wt. % CD complex templated peptization method along with OL characteristics of silica particles synthesized at (b) at 0.05 wt. % CD complex template, (d) 0.15 wt. % CD complex template, (f) 0.1525 wt. % CD complex template, (h) peptization method, and (j) 0.1525 wt. % CD complex templated peptization.

FIG. 10.

Z Scan curves of silica particles synthesized at (a) 0.05 wt. %. CD complex template and (c) 0.15 wt. %. CD complex template, (e) 0.1525 conc. CD complex template, (g) peptization method, and (i) 0.1525 wt. % CD complex templated peptization method along with OL characteristics of silica particles synthesized at (b) at 0.05 wt. % CD complex template, (d) 0.15 wt. % CD complex template, (f) 0.1525 wt. % CD complex template, (h) peptization method, and (j) 0.1525 wt. % CD complex templated peptization.

Close modal
In order to determine the non-linear absorption coefficient βeff, the intensity dependent absorption coefficient [α(I)] for ESA assisted TPA open aperture Z scan recordings were fitted theoretically using the equation, given by61,75–78
(2)
where α is the linear absorption coefficient, I is the incident laser intensity, Is is the saturation intensity, and βeff is the nonlinear absorption coefficient associated with the RSA.66–69 
The non-linear absorption in this case is proportional to the square of the simultaneous intensity (I), given by
(3)
where z is the propagation distance within the sample. The first term in Eq. (3) expresses the SA, and the next term indicates effective TPA part.
Normalized transmittance is given by
(4)
where
(5)
and
(6)
Here, Leff is the effective sample length, L is the sample length, α0 is the unsaturated linear absorption coefficient, z is the position of the sample,
(7)
is the Raleigh range, ω0 is the beam waist radius at the focal point, and λ is the wavelength of the laser beam. The imaginary part of the third order non-linear susceptibility (χ3) is given by
(8)
where n0 is the linear refractive index, c is the speed of the light, and ɛ0 is the permittivity of free space. Using Eqs. (2) and (4), the experimental data were fitted to the theoretical model to obtain non-linear parameters.

The numerically fitted OA Z-scan results brought to light the enhanced NLO responses of the mesoporous silica nanoparticles synthesized at different concentrations of the CD complex template in the Stober process method and peptization method. The obtained βeff were tabulated in Table II.

TABLE II.

The non-linear optical data of silica nanoparticles.

SampleNonlinear absorption coefficient, β × 10−10 m/WOnsets of limiting action 1012 W/m2Optical limiting threshold × 1013 W/m2
S1 0.45 1.35 2.43 
S2 0.49 1.05 2.35 
S3 0.58 0.88 2.21 
S4 1.24 0.75 1.24 
S5 0.65 0.75 
SampleNonlinear absorption coefficient, β × 10−10 m/WOnsets of limiting action 1012 W/m2Optical limiting threshold × 1013 W/m2
S1 0.45 1.35 2.43 
S2 0.49 1.05 2.35 
S3 0.58 0.88 2.21 
S4 1.24 0.75 1.24 
S5 0.65 0.75 

2. Optical limiting studies

The optical limiting data of silica samples for different concentration of DNA of CD complex template were extracted from the graphs of normalized transmittance obtained from open aperture Z-scan analysis against input fluence [Figs. 10(b), 10(d), 10(f), 10(h), and 10(j)]. From the graph, the onsets of limiting action (the value of input fluence at which the intensity of output transmittance starts decreasing) are observed at 3.29, 4.74, 5.23, 8.97, 1.46, and 2.16 W/m2 for the corresponding silica particles synthesized at 0.05 wt. % of the CD complex template, 0.15 wt. % of the CD complex template, 0.1525 wt. % of the CD complex template, peptization method, and 0.1525 wt. % of the CD complex in the peptization method, respectively, and the limiting threshold (LT) values (the value of input fluence at which the intensity of output pulse becomes 50% of the initial value) for the corresponding samples are 2.43, 2.35, 2.21, 1.24, and 0.75 of the order of 1013 W/m2 respectively. The values of the optical limiting threshold corresponding to each sample was tabulated in Table II. From the results, it is evident that the particles in the present study exhibit exceptional optical limiting action with low onset and limiting threshold values. The open aperture Z scan report shown that the sample S5 shows great non-linear absorption coefficient and has low optical limiting threshold value. The studies on mesoporous nanoparticles show that the non-linear optical property can be enhanced by reducing pore size and particle crystalline orientation.77 It is noticed that the enhancement of nonlinearity may be due to the influence of the CD complex template during the formation of silica particles. However, the materials in the present study have shown remarkable nonlinear response with substantial increment in βeff of the order 10−10 m W−1 and exceptionally well optical limiting behavior with very high limiting threshold, making them capable materials for optical power limiting devices in photonics.

The influence of porous substrate on the enhancement of photoluminescence of dyes is always a subject of laser applications.28 We have selected Rhodamine 6G (R6G) as its inclusion with various nanomaterials, such as, silica has been reported to improve the photoluminescence property. Moreover, it is a known cationic dye and has a strong optical absorption with high fluorescence yield. As already discussed in the Introduction, silica possesses a negative polarity in aqueous medium; therefore, R6G has an affinity to adsorb on the silica surface.

We have conducted UV–visible absorption spectra of R6G– DMSO solution with different concentrations of CD complex templated silica nanoparticles. We have taken three R6G–DMSO solutions of samples A, B, and C, of which samples B and C were dispersed with 0.1 and 0.3 g of silica nanoparticles, respectively. From Fig. 11, the UV absorption peak was shifted to the red region on increasing the concentration of silica particles. The UV–visible absorption peak of R6G in DMSO (sample A) was found around at 520 nm. The corresponding absorption peaks of B and C are seen around at 535 and 539 nm, respectively. This red shift of the absorption peaks indicates that there is a change in the electronic ground state and excited state of the R6G molecules so that the size of the energy gap between the dye particles changes as the nature of solvent (DMSO) changes. As on adding the silica particles to DMSO, a dispersive force (interaction) between DMSO and silica particles would take place. This dispersive force is influenced by the polarity of both solvent and solute. As we know, silica is a highly polar compound and DMSO is a polar solvent. For a polar solute in polar solvent, dipole–dipole interactions will stabilize the ground state of the solute according to the polarity of the solvent. In this case, the solvent molecules will be oriented around the ground-state dipole, and if the dipole moment is increased by the transition, the excited state will be stabilized more by more polar solvents (those with higher dielectric constants). This reflected in the absorption or emission spectrum of the solute as differences in the position, intensity, and shape of the spectroscopic bands. This effect is called solvatochromism.79–82 When the spectroscopic band occurs in the visible part of the spectrum, solvatochromism is observed as a change of the color. From this observation, it is clear that on the addition of silica particles to the DMSO, its polarity got changed and, subsequently, the orientation of the DMSO molecules. Therefore, when considering the polarity and orientation of DMSO of samples B and C, they have been changed as comparing with sample A. Therefore, the excitation energy of the dye molecules is decreased, thereby a red shift occurred. Moreover, the particles present in the solvent influence the polarity of the solvent, and the solvent–solute bonding may increase. Thereby, Fig. 11 provided an additional information about the stability of solubility of the CD complex templated silica nanoparticles in DMSO.

FIG. 11.

UV–visible absorption results of (a) R6G in DMSO, (b) R6G in 0.1 g of CD complex templated silica nanoparticles dissolved DMSO, and (c) R6G in 0.3 g of CD complex templated silica nanoparticles dissolved DMSO.

FIG. 11.

UV–visible absorption results of (a) R6G in DMSO, (b) R6G in 0.1 g of CD complex templated silica nanoparticles dissolved DMSO, and (c) R6G in 0.3 g of CD complex templated silica nanoparticles dissolved DMSO.

Close modal

Additionally, Fig. 11 also shows that the absorption peak value is seen as reducing on increasing the CD complex templated silica particles. This may be due to that as the inclusion of CD complex templated silica in the DMSO, they were uniformly dispersed in the solvent and formed as a platform to adsorb R6G to their surface. Hence, the number of dye particles in the solvent along the path of the UV–visible beam was reduced; hence, this, in turn, affected the absorption peak value.

Photoluminescence (PL) response of R6G in different polar state of DMSO was conducted using the OHSP-350 model of the spectral analyzer. The experiment was carried out in eight differently prepared silica samples dispersed DMSO to analyze the PL characterization. Sample S1 is corresponding to silica nanoparticles prepared using the Stober process without any surfactant; S2 is for silica particles prepared in the presence of CTAB as the surfactant in the Stober process; S3 belongs to silica particles prepared by the peptization method; and S4, S5, S6, S7, and S8 are corresponding to the silica particles, which were prepared in 0.01 weight percentage of the CD complex, 0.05 weight percentage of the CD complex, 0.15 weight percentage of the CD complex, 0.1525 weight percentage of the CD complex, and 0.1525 weight percentage of the CD complex in the peptization method, respectively. The weight of silica for eight samples was maintained with the same quantity, and it was 0.0124 g. The eight samples of silica were added in 20 ml of DMSO and sonicated at room temperature until the silica particles were dispersed uniformly. Another solution of 0.01 mM of R6G in DMSO was prepared. From Fig. 12, it is clear that the PL response of R6G shows an enhancement as going from silica samples S1 to S8. Figure 12 pointed that the PL response was so high for samples S4–S8, which were synthesized using as the order of increasing concentration of the CD complex template. As the concentration of the CD complex template increases, the silica samples also show improved physical property, which was also discussed in the BET analysis. The CD complex templated silica nanoparticles provided large area to adsorb R6G particles on their surface. As going from S4–S8, the adsorption rate is also increased. The silica particles in DMSO were uniformly dispersed and their stability to hold with the solvent molecule is so high so that their duration to present in the DMSO is also high. These facts, in turn, enhance the absorption of spectral energy against the R6G particles so that the concentration of dye molecules is decreased in the solvent, and thereby, energy could distribute through the medium in a way that dye molecule could get enough time to absorb sufficient energy. In addition, the silica particles encapsulated the dye particles; hence, this may reduce the oxidation of dye particles and make the dye particle to become photostable. The presence of silica nanosubstrate in the solvent also reduced the interaction between dye molecules. Thereby, the concentration quenching can be avoided. These all features may be the net effect in the enhancement of PL response of R6G. This enhancement of photoluminescence of R6G can be used to lower the lasing threshold of gain medium in the tunable dye laser systems and also can be used for synthesizing photo stable dye incorporated silica nanoparticles for cell imaging purposes.83–85 

FIG. 12.

PL responses of 0.01 mM of R6G in DMSO solution of silica nanoparticles prepared by (S1) the Stober process, (S2) the Stober process in which CTAB was used as the surfactant, (S3) the peptization method and silica nanoparticles prepared in the Stober process (S4) using 0.01 wt of the CD complex template, (S5) 0.05 wt. % of the CD complex template, (S6) 0.15 wt. % of the CD complex template, (S7) 0.1525 wt. % of the CD complex template, and (S8) 0.1525 wt. % of the CD complex template in the peptization method.

FIG. 12.

PL responses of 0.01 mM of R6G in DMSO solution of silica nanoparticles prepared by (S1) the Stober process, (S2) the Stober process in which CTAB was used as the surfactant, (S3) the peptization method and silica nanoparticles prepared in the Stober process (S4) using 0.01 wt of the CD complex template, (S5) 0.05 wt. % of the CD complex template, (S6) 0.15 wt. % of the CD complex template, (S7) 0.1525 wt. % of the CD complex template, and (S8) 0.1525 wt. % of the CD complex template in the peptization method.

Close modal

Mesoporous silica nanoparticles were synthesized using the Stober process and peptization method in which surface modified DNA with CTAB was used as the capping agent. The characterization results show that the CD complex provides twin benefits of both DNA and CTAB during the synthesis procedure. The CD complex acted as a stabilizer, which enhanced the morphological and the non-linear optical properties of the obtained silica nanoparticles without altering the spherical geometry of the silica nanoparticles. The XRD and UV results confirmed the formation of silica nanoparticles, and they also show the presence of CD complex functionalized silica. In XRD results, the FWHM was higher for the silica nanoparticles, which were prepared in the presence of CD complex. The UV results also proved that the concentration of the DNA effects the size and porosity of the silica particles. The particles size were reduced by the effect of DNA, and it was confirmed by varying the concentration of DNA. As the concentration of the DNA was increased, a blue shift in the peak is observed. This indicated that the presence of DNA in the synthesis process influenced in reducing the size of the silica particles. The surface properties of the obtained silica particles were studied by BET–BJH analysis. BET results show that on increasing the CD complex concentration, the pore diameter was found to decrease within the mesoporous range. The silica nanoparticles obtained from the CD complex template have remarkable hysteresis loop. This property of them helped to stick the adsorbent tightly to the adsorbate. The SEM characterization revealed the surface modification occurred in the silica nanoparticles formed in the presence of CD complex template. In addition to this, the SEM results shown that the presence of CD complex did not alter the specificity of the Stober process. Non-linear optical characterization was also performed by Z-scan techniques. With that shown, that the novel silica nanoparticles show a significant improvement of the non-linear optical absorption property, and their nonlinear absorption coefficient is found to increase on increasing the concentration of CD complex template, and the optical limiting threshold value of the CD complex templated silica nanoparticles was also reduced on increasing the concentration of DNA.

All the characterization results pointed that the CD complex works effectively as a stabilizing agent encompassed with the conventional Stober process for synthesizing mesoporous nano-sized non-linear optical silica nanoparticles. This method can be performed in both acidic [peptization method (HCL)] and basic media (the Stober process in which ammonia is used as the catalyst). The effect of CD complex templated silica nanoparticles is also found in the enhancement of photoluminescence of R6G. This peculiarity of silica nanoparticles will be a promise for the tunable dye laser system. While using the CD complex as the biotemplate, the toxic reagents and the complicity of experimental procedure can be avoided during the synthesis of mesoporous silica nanoparticles. Hence, the CD complex incorporating the Stober process is a novel, non-toxic, and cost effective method for synthesizing surface modified, tunable structured, biocompatible, and enhanced non-linear optical silica nanoparticles. These features of the silica nanoparticles make them to be used for drug delivery, catalysts, surface modified thin films using the doped surface, lasing, bio-sensing, gas sensing, bio-imaging, and semiconducting-photonic devices.

The authors conveyed their gratefulness to the University of Calicut; NIT Calicut; Bharathidasan University, Thiruchirappalli; and Government College for Women, Thiruvananthapuram, for the different studies of synthesized samples.

The authors have no conflicts to disclose.

All authors contributed to the study conception and design. Material conception, experimental design, data collection, and analysis were performed by Bhagyasree G. S., Reena V. N., Abith M., Sabari Girisun T. C., and Nithyaja B. The first draft of the manuscript was written by Bhagyasree G. S., and all authors commented on previous versions of the manuscript. All authors read and approved the final manuscript.

G. S. Bhagyasree: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Resources (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). V. N. Reena: Conceptualization (equal); Data curation (equal); Investigation (equal); Methodology (equal); Resources (equal); Writing – review & editing (equal). M. Abith: Conceptualization (equal); Methodology (equal); Resources (equal); Software (equal); Writing – review & editing (equal). T. C. Sabari Girisun: Conceptualization (equal); Formal analysis (equal); Methodology (equal); Resources (equal); Software (equal); Writing – review & editing (equal). B. Nithyaja: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
W.
Stöber
,
A.
Fink
, and
E.
Bohn
, “
Controlled growth of monodisperse silica spheres in the micron size range
,”
J. Colloid Interface Sci.
26
,
62
(
1968
).
2.
V.-C.
Niculescu
, “
Mesoporous silica nanoparticles for bio-applications
,”
Front. Mater.
7
,
36
(
2020
).
3.
C.
Bharti
,
U.
Nagaich
,
A. K.
Pal
, and
N.
Gulati
, “
Mesoporous silica nanoparticles in target drug delivery system: A review
,”
Int. J. Pharm. Invest.
5
,
124
(
2015
).
4.
X.
Ma
,
Y.
Zhao
,
K. W.
Ng
, and
Y.
Zhao
, “
Integrated hollow mesoporous silica nanoparticles for target drug/siRNA co‐delivery
,”
Chem. Eur. J.
19
,
15593
(
2013
).
5.
N.
Iturrioz-Rodríguez
,
M. A.
Correa-Duarte
, and
M. L.
Fanarraga
, “
Controlled drug delivery systems for cancer based on mesoporous silica nanoparticles
,”
Int. J. Nanomed.
14
,
3389
(
2019
).
6.
Y.
Alyassin
,
E. G.
Sayed
,
P.
Mehta
,
K.
Ruparelia
,
M. S.
Arshad
,
M.
Rasekh
,
J.
Shepherd
,
I.
Kucuk
,
P. B.
Wilson
,
N.
Singh
, and
M.-W.
Chang
, “
Application of mesoporous silica nanoparticles as drug delivery carriers for chemotherapeutic agents
,”
Drug Discov.
25
,
1513
(
2020
).
7.
B.
Siddiqui
,
A. u.
Rehman
,
I.-u.
Haq
,
A. A.
Al-Dossary
,
A.
Elaissari
, and
N.
Ahmed
, “
Exploiting recent trends for the synthesis and surface functionalization of mesoporous silica nanoparticles towards biomedical applications
,”
Int. J. Pharm.
4
,
100116
(
2022
).
8.
A. S.
Taleghani
,
A. T.
Nakhjiri
,
M. J.
Khakzad
,
S. M.
Rezayat
,
P.
Ebrahimnejad
,
A.
Heydarinasab
,
A.
Akbarzadeh
, and
A.
Marjani
, “
Mesoporous silica nanoparticles as a versatile nanocarrier for cancer treatment: A review
,”
J. Mol. Liq.
15
,
328
(
2021
).
9.
S.
Jafari
,
H.
Derakhshankhah
,
L.
Alaei
,
A.
Fattahi
,
B. S.
Varnamkhasti
, and
A. A.
Saboury
, “
Mesoporous silica nanoparticles for therapeutic/diagnostic applications
,”
Biomed. Pharmacother.
109
,
1100
(
2019
).
10.
R. R.
Castillo
,
D.
Lozano
,
B.
González
,
M.
Manzano
,
I.
Izquierdo-Barba
, and
M.
Vallet-Regí
, “
Advances in mesoporous silica nanoparticles for targeted stimuli-responsive drug delivery: An update
,”
Expert Opin. Drug Delivery
16
,
415
(
2019
).
11.
Z.
Li
,
Y.
Zhang
, and
N.
Feng
, “
Mesoporous silica nanoparticles: Synthesis, classification, drug loading, pharmacokinetics, biocompatibility, and application in drug delivery
,”
Expert Opin. Drug Delivery
16
,
219
(
2019
) (
2019
).
12.
V.
Mamaeva
,
C.
Sahlgren
, and
M.
Lindén
, “
Mesoporous silica nanoparticles in medicine–Recent advances
,”
Adv. Drug Deliv. Rev.
65
,
689
(
2013
).
13.
S. K.
Natarajan
and
S.
Selvaraj
, “
Mesoporous silica nanoparticles: Importance of surface modifications and its role in drug delivery
,”
RSC Adv.
4
,
14328
(
2014
).
14.
F.
Farjadian
,
A.
Roointan
,
S.
Mohammadi-Samani
, and
M.
Hosseini
, “
Mesoporous silica nanoparticles: Synthesis, pharmaceutical applications, biodistribution, and biosafety assessment
,”
J. Chem. Eng.
359
,
684
(
2019
).
15.
S.
Nasreen
,
U.
Rafique
,
S.
Ehrman
, and
M. A.
Ashraf
, “
Synthesis and characterization of mesoporous silica nanoparticles for environmental remediation of metals, PAHs and phenols
,”
Ekoloji
27
,
1625
(
2018
).
16.
M.
Ghaferi
,
M.
Koohi Moftakhari Esfahani
,
A.
Raza
,
S.
Al Harthi
,
H.
Ebrahimi Shahmabadi
, and
S. E.
Alavi
, “
Mesoporous silica nanoparticles: Synthesis methods and their therapeutic use-recent advances
,”
J. Drug Target.
29
,
1
(
2020
).
17.
M.
Khoeini
,
A.
Najafi
,
H.
Rastegar
, and
M.
Amani
, “
Improvement of hollow mesoporous silica nanoparticles synthesis by hard-templating method via CTAB surfactant
,”
Ceram. Int.
45
,
12820
(
2019
).
18.
H.
Zhou
,
J.
Sun
,
B.
Ren
,
F.
Wang
,
X.
Wu
, and
S.
Bai
, “
Effects of alkaline media on the controlled large mesopore size distribution of bimodal porous silicas via sol-gel methods
,”
Powder Technol.
259
,
46
(
2014
).
19.
V. N.
Reena
,
H.
Misha
,
G. S.
Bhagyasree
, and
B.
Nithyaja
, “
Enhanced photoluminescence and color tuning from Rhodamine 6G-doped sol–gel glass matrix via DNA templated CdS nanoparticles
,”
AIP Adv.
12
,
105217
(
2022
).
20.
G. C.
Carvalho
,
G. D.
Marena
,
J. C. F.
Karnopp
,
J.
Jorge
,
R. M.
Sábio
,
M. A. U.
Martines
,
T. M.
Bauab
, and
M.
Chorilli
,
Cetyltrimethylammonium bromide in the synthesis of mesoporous silica nanoparticles: General aspects and in vitro toxicity
,
Adv. Colloid Interface Sci.
307
,
102746
(
2022
).
21.
Z.-A.
Qiao
,
L.
Zhang
,
M.
Guo
,
Y.
Liu
, and
Q.
Huo
, “
Synthesis of mesoporous silica nanoparticles via controlled hydrolysis and condensation of silicon alkoxide
,”
Chem. Mater.
21
,
3823
(
2009
).
22.
A.
Lodha
,
M.
Lodha
,
A.
Patel
,
J.
Chaudhuri
,
J.
Dalal
,
M.
Edwards
, and
D.
Douroumis
, “
Synthesis of mesoporous silica nanoparticles and drug loading of poorly water soluble drug cyclosporin A
,”
J. Pharm. Bioallied Sci.
4
,
92
(
2012
).
23.
N. I.
Vazquez
,
Z.
Gonzalez
,
B.
Ferrari
, and
Y.
Castro
, “
Synthesis of mesoporous silica nanoparticles by sol–gel as nanocontainer for future drug delivery applications
,”
Bol. Soc. Esp. Ceram. Vidr.
56
,
139
(
2017
).
24.
A.
Presentato
,
F.
Armetta
,
A.
Spinella
,
D. F.
Chillura Martino
,
R.
Alduina
, and
M. L.
Saladino
, “
Formulation of mesoporous silica nanoparticles for controlled release of antimicrobials for stone preventive conservation
,”
Front. Chem.
8
,
699
(
2020
).
25.
R. K.
Satvekar
,
M. R.
Phadatare
,
V. A.
Karande
,
R. N.
Patil
,
B. M.
Tiwale
, and
S. H.
Pawar
, “
Influence of silane content on the optical properties of sol gel derived spin coated silica thin films
,”
Int. J. Basic Appl. Sci.
1
,
468
(
2012
).
26.
J.
Adira Jaafar
,
N.
Hidayatul Nazirah Kamarudin
,
H.
Dina Setiabudi
,
S.
Najiha Timmiati
, and
T.
Lee Peng
, “
Mesoporous silica nanoparticles and waste derived-siliceous materials for doxorubicin adsorption and release
,”
Mater. Today: Proc.
19
,
1420
(
2019
).
27.
A.
Mehdi
,
S.
Dourdain
,
J. F.
Bardeau
,
C.
Reyé
,
R. J. P.
Corriu
, and
A.
Gibaud
, “
First direct synthesis of highly ordered bifunctionalized mesoporous silica thin films
,”
J. Nanosci. Nanotechnol.
6
,
377
(
2006
).
28.
X.
Li
,
O.
Niitsoo
, and
A.
Couzis
, “
Electrostatically driven adsorption of silica nanoparticles on functionalized surfaces
,”
J. Colloid Interface Sci.
394
,
26
(
2013
).
29.
S.
García-Revilla
,
J.
Fernández
,
M. A.
Illarramendi
,
B.
García-Ramiro
,
R.
Balda
,
H.
Cui
, and
D.
Levy
, “
Ultrafast random laser emission in a dye-doped silica gel powder
,”
Opt. Express
16
,
12251
(
2008
).
30.
Y.
Shin
,
D.
Lee
,
K.
Lee
,
K. H.
Ahn
, and
B.
Kim
, “
Surface properties of silica nanoparticles modified with polymers for polymer nanocomposite applications
,”
J. Ind. Eng. Chem.
14
,
515
(
2008
).
31.
K. C. W.
Wu
and
Y.
Yamauchién
, “
Controlling physical features of mesoporous silica nanoparticles (MSNs) for emerging applications
,”
J. Mater. Chem.
22
,
1251
(
2012
).
32.
S. A.
Jadhav
, “
Incredible pace of research on mesoporous silica nanoparticles
,”
Inorg. Chem. Front
1
,
735
(
2014
).
33.
Ł.
Zych
,
A. M.
Osyczka
,
A.
Łacz
,
A.
Różycka
,
W.
Niemiec
,
A.
Rapacz-Kmita
,
E.
Dzierzkowska
, and
E.
Stodolak-Zych
, “
How surface properties of silica nanoparticles influence structural, microstructural and biological properties of polymer nanocomposites
,”
Materials
14
,
843
(
2021
).
34.
B.
Shi
,
Y. K.
Shin
,
A. A.
Hassanali
, and
S. J.
Singer
, “
DNA binding to the silica surface
,”
J. Phys. Chem. B
119
,
11030
(
2015
).
35.
S. R.
Dugasani
,
B.
Gnapareddy
,
M. R.
Kesama
,
S.
Jeon
,
J. H.
Jeong
, and
S. H.
Park
, “
Optoelectronic properties of DNA thin films implanted with titania nanoparticle-coated multiwalled carbon nanotubes
,”
AIP Adv.
(
1
)
9
,
015011
(
2019
).
36.
M.
Samoc
,
A.
Samoc
, and
J. G.
Grote
, “
Complex nonlinear refractive index of DNA
,”
Chem. Phys. Lett.
431
,
132
(
2006
).
37.
M. R.
Kesama
,
S. R.
Dugasani
,
Y. J.
Cha
,
J.
Son
,
B.
Gnapareddy
,
S.
Yoo
,
D. K.
Yoon
, and
S. H.
Park
, “
Optoelectrical and mechanical properties of multiwall carbon nanotube-integrated DNA thin films
,”
Nanotechnology
30
,
245704
(
2019
).
38.
V.
Arasu
,
S.
Reddy Dugasani
,
J.
Son
,
B.
Gnapareddy
,
S.
Jeon
,
J.-H.
Jeong
, and
S.
Ha Park
, “
Thickness, morphology, and optoelectronic characteristics of pristine and surfactant-modified DNA thin films
,”
J. Phys. D: Appl. Phys.
50
,
415602
(
2017
).
39.
K.
Liu
,
L.
Zheng
,
C.
Ma
,
R.
Göstl
, and
A.
Herrmann
, “
DNA-surfactant complexes: Self-assembly properties and applications
,”
Chem. Soc. Rev.
46
,
5147
(
2017
).
40.
A.
Cuervo
,
P. D.
Dans
,
J. L.
Carrascosa
,
M.
Orozco
,
G.
Gomila
, and
L.
Fumagalli
, “
Direct measurement of the dielectric polarization properties of DNA
,”
Proc. Natl. Acad. Sci. U. S. A.
111
,
E3624
(
2014
).
41.
D. M.
Joyce
, “
The development of DNA–based bio- polymer hybrid thin films for capacitor applications
,” Ph.D thesis,
The School of Engineering of the University of Dayton
,
2013
.
42.
B.
Singh
,
N. S.
Sariciftci
,
J. G.
Grote
, and
F. K.
Hopkins
, “
Bio-organic-semiconductor-field-effect-transistor based on deoxyribonucleic acid gate dielectric
,”
J. Appl. Phys.
100
,
024514
(
2006
).
43.
S.
Husaini
,
R. G.
Bedford
,
E. M.
Heckman
, and
A. C.
Lesko
, “
Nonlinear optical properties of a graphene-based DNA composite
,” in
2014 Conference on Lasers and Electro-Optics (CLEO)
(
Optical Society of America, IEEE
,
2014
), p.
1
.
44.
A.
Tane
,
F.
Kajzar
,
R.
Zgarian
,
I.
Rau
,
D.
Grabarek
,
P.
Karpinski
, and
A.
Miniewicz
, “
Refractive index and surface relief grating formation in DNA based dye-doped films
,”
Macromol. Res.
21
,
331
(
2013
).
45.
M.
Yoshikawa
,
M.
Maruhashi
,
Y.
Iwamoto
, and
N.
Ogata
, “
Optical resolution of racemic amino acids through DNA-Poly(4-vinylbenzyl)trimethylammonium polyion complex membranes
,”
Polym. J.
39
,
193
(
2007
).
46.
L.
Wang
,
J.
Yoshida
,
N.
Ogata
,
S.
Sasaki
, and
T.
Kajiyama
, “
Self-assembled supramolecular films derived from marine deoxyribonucleic acid (DNA)–cationic surfactant complexes: Large-scale preparation and optical and thermal properties
,”
Chem. Mater.
13
,
1273
(
2001
).
47.
O.
Ikkala
and
G.
ten Brinke
, “
Functional materials based on self-assembly of polymeric supramolecules
,”
Science
295
,
2407
(
2002
).
48.
E.
Alizadeh
and
L.
Sanche
, “
Role of humidity and oxygen level on damage to DNA induced by soft X-rays and low-energy electrons
,”
J. Phys. Chem. C
117
,
22445
(
2013
).
49.
H.
You
,
H.
Spaeth
,
V. N. L.
Linhard
, and
A. J.
Steckl
, “
Role of surfactants in the interaction of dye molecules in natural DNA polymers
,”
Langmuir
25
,
11698
(
2009
).
50.
A. J.
Steckl
,
H.
Spaeth
,
H.
You
,
E.
Gomez
, and
J.
Grote
, “
DNA as an optical material
,”
Opt. Photonics News
2
,
34
(
2011
).
51.
C.
Cheng
and
S.-Y.
Ran
, “
Interaction between DNA and trimethyl-ammonium bromides with different alkyl chain lengths
,”
Sci. World J.
11
,
16
(
2014
).
52.
T. B.
Singh
,
N. S.
Sariciftci
, and
J. G.
Grote
, “
Bio-organic optoelectronic devices using DNA
,”
Adv. Polym. Sci.
223
,
73
(
2010
).
53.
C.
Yang
,
D.
Moses
, and
A. J.
Heeger
, “
Base-pair stacking in oriented films of DNA–surfactant complex
,”
Adv. Mater.
15
,
1364
(
2003
).
54.
J. S.
Hwang
,
S. W.
Hwang
, and
D.
Ahn
, “
Formation of electrical interconnects by self-trapping of deoxyribonucleic acid molecules
,”
Jpn. J. Appl. Phys.
43
,
3803
(
2004
).
55.
Y. W.
Kwon
,
C. H.
Lee
,
D. H.
Choi
, and
J.-I.
Jin
, “
Materials science of DNA
,”
J. Mater. Chem.
19
,
1353
(
2008
).
56.
A.
Samoc
,
M.
Samoc
,
J. G.
Grote
,
A.
Miniewicz
, and
B.
Luther-Davies
, “
Optical properties of deoxyribonucleic acid (DNA) polymer host
,” in
Optical Materials in Defence Systems Technology III
(
2006
).
57.
N.
Slesiona
,
S.
Thamm
,
H. L. K. S.
Stolle
,
V.
Weißenborn
,
P.
Müller
,
A.
Csáki
, and
W.
Fritzsche
, “
DNA-Biofunctionalization of CTAC-capped gold nanocubes
,”
Nanomaterials
10
,
1119
(
2020
).
58.
Z.
Yu
,
W.
Li
,
J. A.
Hagen
,
Y.
Zhou
,
D.
Klotzkin
,
J. G.
Grote
, and
A. J.
Steckl
, “
Photoluminescence and lasing from deoxyribonucleic acid (DNA) thin films doped with sulforhodamine
,”
Appl. Opt.
46
,
1507
(
2007
).
59.
A.
Paul
and
B.
Bhattacharya
, “
DNA functionalized carbon nanotubes for nonbiological applications
,”
Mater. Manuf.
25
,
891
(
2010
).
60.
N.
Balan
,
M.
Hari
, and
V. P. N.
Nampoori
, “
Selective mode excitation in dye-doped DNA polyvinyl alcohol thin film
,”
Appl. Opt.
48
,
3521
(
2009
).
61.
B.
Nithyaja
,
H.
Misha
, and
V. P. N.
Nampoori
, “
Fluorescence enhancement of silver nanoparticles using DNA as a stabilizing agent
,”
Proc. SPIE
8173
,
383
(
2011
).
62.
B.
Nithyaja
,
H.
Misha
, and
V.
P. N. Nampoori
, “
Synthesis of silver nanoparticles in DNA template and its influence on nonlinear optical properties
,”
Nanosci. Nanotechnol.
2
,
99
(
2012
).
63.
N. A. F.
Aboul-Maaty
and
H. A. S.
Oraby
, “
Extraction of high-quality genomic DNA from different plant orders applying a modified CTAB-based method
,”
Bull. Natl. Res. Cent.
43
,
1
(
2019
).
64.
J. D.
Clarke
, “
Cetyltrimethyl ammonium bromide (CTAB) DNA miniprep for plant DNA isolation
,”
Cold Spring Harb. Protoc.
pdb-prot
,
5177
(
2009
).
65.
A. N.
Golubev
,
S. I.
Nikitin
,
M. A.
Smirnov
, and
A. L.
Stepanov
, “
Spectral dependence of nonlinear optical absorption of silica glass with copper nanoparticles
,”
J. Phys.: Conf. Ser.
324
,
012038
(
2011
).
66.
A.
Liberman
,
N.
Mendez
,
W. C.
Trogler
, and
A. C.
Kummel
, “
Synthesis and surface functionalization of silica nanoparticles for nanomedicine
,”
Surf. Sci. Rep.
69
,
132
(
2014
).
67.
V.
Jaya
and
B.
Arpita
, “
Analysis on synthesis of silica nanoparticles and its effect on growth of T. Harzianum & Rhizoctonia Species
,”
Biomed. J. Sci. Technol. Res.
10
,
7890
(
2018
).
68.
Y.-K.
Lee
,
Y.
Rock Yoon
, and
H.-K.
Rhee
, “
Preparation of colloidal silica using peptization method
,”
Colloids Surf., A
173
,
109
(
2000
).
69.
H.
Darabi
,
M.
Adelifard
, and
Y.
Rajabi
, “
Characterization of nonlinear optical refractive index for graphene oxide–Silicon oxide nanohybrid composite
,”
J. Nonlinear Opt. Phys. Mater.
28
,
1950005
(
2019
).
70.
I. M.
Joni
,
L.
Nulhakim
,
M.
Vanitha
, and
C.
Panatarani
, “
Characteristics of crystalline silica (SiO2) particles prepared by simple solution method using sodium silicate (Na2SiO3) precursor
,”
J. Phys.: Conf. Ser.
1080
,
012006
(
2018
).
71.
A. O.
Bokuniaeva
and
A. S.
Vorokh
, “
Estimation of particle size using the Debye equation and the Scherrer formula for polyphasic TiO2 powder
,”
J. Phys.: Conf. Ser.
1410
,
012057
(
2019
).
72.
A.
Monshi
,
M. R.
Foroughi
, and
M. R.
Monshi
, “
Modified Scherrer equation to estimate more accurately nano-crystallite size using XRD
,”
World J. Nano. Sci. Eng.
2
,
154
(
2012
).
73.
M. S.
Hossain
,
M.
Mahmud
,
M. B.
Mobarak
,
S.
Sultana
,
M. A. A.
Shaikh
,
S.
Ahmed
, and
S.
Ahmed
, “
New analytical models for precise calculation of crystallite size: Application to synthetic hydroxyapatite and natural eggshell crystalline materials
,”
Chem. Pap.
76
,
7245
(
2022
).
74.
M.
Rabiei
,
A.
Palevicius
,
A.
Monshi
,
S.
Nasiri
,
A.
Vilkauskas
, and
G.
Janusas
, “
Comparing methods for calculating nano crystal size of natural hydroxyapatite using X-ray diffraction
,”
Nanomaterials
10
,
1627
(
2020
) All the initials of the authors are verified using google scholar site and found no error..
75.
M.
Thommes
,
K.
Kaneko
,
A. V.
Neimark
,
J. P.
Olivier
,
F.
Rodriguez-Reinoso
,
J.
Rouquerol
, and
K. S.
Sing
, “
Physisorption of gases with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report)
,”
Pure Appl. Chem.
87
,
1051
(
2015
).
76.
S.
Shettigar
,
K.
Chandrasekharan
,
G.
Umesh
,
B. K.
Sarojini
, and
B.
Narayana
, “
Studies on nonlinear optical parameters of bis-chalcone derivatives doped polymer
,”
Polymer
47
,
3565
(
2006
).
77.
S.
Shettigar
,
G.
Umesh
,
K.
Chandrasekharan
, and
B.
Kalluraya
, “
Third order nonlinear optical properties and two photon absorption in newly synthesized phenyl sydnone doped polymer
,”
Synth. Met.
157
,
142
(
2007
).
78.
L. A.
Golovan
and
V. Y.
Timoshenko
, “
Nonlinear-optical properties of porous silicon nanostructures
,”
J. Nanoelectron. Optoelectron.
8
(
3
),
223
(
2013
).
79.
D.
Udayakumar
,
A.
John Kiran
,
A. V.
Adhikari
,
K.
Chandrasekharan
,
G.
Umesh
, and
H. D.
Shashikala
, “
Third-order nonlinear optical studies of newly synthesized polyoxadiazoles containing 3, 4-dialkoxythiophenes
,”
Chem. Phys.
331
,
125
(
2006
).
80.
M. C.
Fuertes
,
S.
Colodrero
,
G.
Lozano
,
A. R.
González-Elipe
,
D.
Grosso
,
C.
Boissière
, and
H.
Míguez
, “
Sorption properties of mesoporous multilayer thin films
,”
J. Phys. Chem. C
112
,
3157
(
2008
).
81.
A. M.
Losa
,
A.
Biancardi
, and
B.
Mennucci
, “
What is solvatochromism
,”
J. Phys. Chem. B
114
,
17128
(
2010
).
82.
S.
Nigam
and
S.
Rutan
, “
Principles and applications of solvatochromism
,”
Appl. Spectrosc.
55
,
362A
(
2001
).
83.
E. S.
Dodsworth
,
M.
Hasegawa
,
M.
Bridge
, and
W.
Linert
, Solvatochromism, C. C. C. II, 351,
2003
.
84.
A.
Ahmad
,
N. D.
Zakaria
, and
K. A.
Razak
, “
Photostability effect of silica nanoparticles encapsulated fluorescence dye
,”
AIP Conf. Proc.
1901
,
020010
(
2017
).
85.
I.
Miletto
,
A.
Gilardino
,
P.
Zamburlin
,
S.
Dalmazzo
,
D.
Lovisolo
,
G.
Caputo
, and
G.
Martra
, “
Highly bright and photostable cyanine dye-doped silica nanoparticles for optical imaging: Photophysical characterization and cell tests
,”
Dyes Pigm.
84
,
121
(
2010
).