Next Article in Journal
Influence of Extrusion Rate on Microstructure and Mechanical Properties of Magnesium Alloy AM60 and an AM60-Based Metal Matrix Nanocomposite
Next Article in Special Issue
Plasma-Modified PI Substrate for Highly Reliable Laser-Sintered Copper Films Using Cu2O Nanoparticles
Previous Article in Journal
A Strategy for Preparing Solid Polymer Electrolytes Containing In Situ Synthesized ZnO Nanoparticles with Excellent Electrochemical Performance
Previous Article in Special Issue
Optimization of Piezoresistive Strain Sensors Based on Gold Nanoparticle Deposits on PDMS Substrates for Highly Sensitive Human Pulse Sensing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanism Understanding for Size Regulation of Silver Nanowires Mediated by Halogen Ions

1
School of Materials Science and Engineering, Fuzhou University, Fuzhou 350108, China
2
Zijin School of Geology and Mining, Fuzhou University, Fuzhou 350108, China
3
Fujian Key Laboratory of Green Extraction and High-Value Utilization of Energy Metals, Fuzhou University, Fuzhou 350108, China
4
Zijin Mining Group Co., Ltd., Shanghang 364200, China
5
State Key Laboratory of Comprehensive Utilization of Low Grade Refractory Gold Ores, Shanghang 364200, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(15), 2681; https://doi.org/10.3390/nano12152681
Submission received: 7 July 2022 / Revised: 29 July 2022 / Accepted: 1 August 2022 / Published: 4 August 2022
(This article belongs to the Special Issue Low-Dimensional Nanomaterials and Their Applications)

Abstract

:
The controllable preparation of silver nanowires (AgNWs) with a high aspect ratio is key for enabling their applications on a large scale. Herein, the aspect ratio regulation of AgNWs mediated by halogen ion composition in ethylene glycol system was systematically investigated and the size evolution mechanism is elaborately understood. The co-addition of Br and Cl results in AgNWs with the highest aspect ratio of 1031. The surface physicochemical analysis of AgNWs and the density functional theory calculations indicate that the co-addition of Br and Cl contributes to the much-enhanced preferential growth of the Ag(111) crystal plane. At the same time, when Cl and Br coexist in the solution, the growth of the Ag(100) crystal plane on the AgNWs was restrained compared with that in the single Cl system. Resultantly, the enhanced growth of Ag(111) and the inhibited growth of Ag(100) contribute to the formation of AgNWs with a higher aspect ratio in the Cl–Br mixed solution. The results can provide new insights for understanding the morphology and size evolution during the AgNWs preparation in ethylene glycol system.

Graphical Abstract

1. Introduction

Silver nanowires (AgNWs) possesses excellent electrical, thermal, optical, and mechanical properties as well as extraordinary thermoelectric merit [1,2], thus being widely used in surface-enhanced Raman scattering (SERS) [3,4], transparent conductive film (TCF) [5], chemical/biological sensing [6,7], electromagnetic interference (EMI) shielding materials [8,9], flexible and stretchable electronic devices [10], etc. Specifically, the one-dimensional characteristics of AgNWs can facilitate the charge transfer and decrease the electrical resistance as well as promise high flexibility. The aspect ratio of AgNWs (ratio of length to diameter) is a critical parameter in providing highly conductive networks with good transparency for AgNWs-based films [11]. Additionally, AgNWs with a thinner diameter and higher aspect ratio are beneficial for the performance of the TCF, especially for high-end applications, such as touch panel screens that require low sheet resistance and low haze [12]. For example, Khanarian et al. [13] found that superfine Ag NWs (diameter < 50 nm, length > 5 μm) were required to achieve a high transmittance of >90% with a low sheet resistance ~100 Ω sq−1. Mutiso et al. [14] proposed that an aspect ratio as large as 800 was required for the high performance of AgNW TCFs with sheet resistance ~10 Ω sq−1 for a transparency of ~90% at 550 nm. Thinner AgNWs generate more “open area” for light to pass through, and higher aspect ratio decreased the number of high resistance wire contacts in the film [11,15]. Therefore, size regulation for preparing AgNWs with high aspect ratio is urgent needed [2,16].
The polyol method is one of the most commonly used processes for AgNWs preparation among the widely explored strategies, including template-assisted synthesis [17,18,19], electrochemical technique [20,21], the UV irradiation photo-reduction technique [22,23], hydrothermal synthesis [24,25,26], wet chemical synthesis [27], and polyol synthesis [28,29,30,31,32,33,34,35]. Polyol synthesis is highly desirable due to the following merits: (1) simple operation with high yield. AgNWs can be obtained by a facile reduction of AgNO3 in the ethylene glycol (EG) solution at about 160 °C, with the yield being high up to 97% [36]. (2) The easy regulation for AgNWs growth. The length and diameter of AgNWs can be easily tuned by changing the molecular weight of polyvinyl pyrrolidone (PVP), controlling the concentration of added agents or adjusting the agent species such as Fe3+, Cu2+, Cl, and Br [29,30,33,34,35,36]. Nevertheless, further increasing the aspect ratio of AgNWs by the polyol process is an appealing but challenging task.
The formation of AgNWs with high aspect ratio relies on the preferential growth of the Ag(111) crystal plane than the Ag(100) crystal plane [34,35,37]. The preferential growth of the Ag(111) crystal plane can be realized by adding the PVP agent because the selectively adsorbed PVP on the Ag(100) crystal plane can retard the growth of Ag(100) plane [34,35,37,38]. However, the adhered PVP molecules on the AgNWs act as an insulating layer, weakening the application ability of AgNWs [31,39]. A redeeming strategy to regulate the aspect ratio of AgNWs is the addition of halogen ions, namely Cl and Br, which can also tune the relative growth rate of Ag(111) and Ag(100) [2,36]. Compared with the PVP agent, the halogen ions can be easily removed by rinsing with ethanol, therefore posing negligible negative effects on the properties of AgNWs [2]. Deeply understanding the influence mechanism of halogen ions on AgNWs growth is of prime importance for elaborately tuning the morphology of AgNWs, yet still lacks consensus [36,40,41,42].
Herein, the synergy mechanism of Br and Cl on the formation of AgNWs with a high aspect ratio was investigated with the aid of surface chemistry analysis and adsorption energy calculation by density functional theory (DFT). As shown in Figure 1, the length increase of the AgNWs relies on the growth of Ag(111) crystal plane while diameter increase originates from the growth of Ag(100) plane. For single Cl addition, the preferential adsorption of Cl on Ag(100) crystal plane occurs. Resultantly, more Ag+ is accumulated on the Ag(100) plane due to the very low solubility of silver halides, which results in the faster growth of the diameter (d) than the length (L), leading to a relatively low aspect ratio (L/d). For single Br addition, adsorption of Br on both Ag(111) and Ag(100) is weak, which means that only a small amount of Ag+ are accumulated on both planes, resulting in very short and thin AgNWs products. For the co-addition of Br and Cl, the halides are more strongly adsorbed on the Ag(111) crystal plane than that of Ag(100), which facilitates the faster growth of length than of diameter, contributing to the AgNWs with the highest aspect ratio. The results can provide new insights for understanding the AgNWs growth mechanism in the halogen-containing systems of the polyol method.

2. Materials and Methods

2.1. Materials

Silver nitrate (AgNO3, ≥99.8%), acetone (≥99.5%), NaCl (99.5%), NaI (99.5%), and absolute ethyl alcohol (≥99.5%) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Polyvinylpyrrolidone (PVP) (Mw ≈ 40,000 g/mol) was purchased from Sigma-Aldrich (St. Louis, MO, USA). Ferric chloride hexahydrate (FeCl3·6H2O, ≥99.5%) was obtained from Kemiou Chemical Reagent Co., Ltd. (Tianjin, China). Ferric tribromide (FeBr3, 98%) was obtained from Macklin Biochemical Co., Ltd. (Shanghai, China). Ethylene glycol (EG, ≥99.5%) was acquired from Xilong Scientific Co., Ltd. (Shantou, China). Ultra-pure water obtained by a Direct-Q 3 UV water purification system (Millipore, Burlington, MA, USA) and used throughout. All chemicals were used as received without further purification.

2.2. Synthesis of Silver Nanowires

In a typical synthesis procedure, 40 mL EG and 2.489 g PVP (Mw ≈ 40,000 g/mol) were added into a 100 mL three-necked flask placed in an oil bath at room temperature. The flask was slowly heated to 160 °C within 40 min with vigorous magnetic stirring. The flask was sealed with frosted glass caps. When the temperature was stable at 160 °C for 1 h, 200 μL of 0.05 M control agent, i.e., FeCl3, FeBr3, or their mixture in EG, was added into the flask. After ten minutes, 1 mL EG solution containing 0.1274 g of AgNO3 ( C AgNO 3 = 0.75   M ) solutions was added to the flask at the rate of 25 μL/s by injection pump. Then, 9 mL of EG solution containing 2.4201 g of AgNO3 ( C AgNO 3 = 1.583   M ) solution was added into the flask immediately. Then the flask was sealed until the solution became glistening, indicating the formation of AgNWs. Upon the completion of the reaction, the reaction solution was transferred to an ice bath and allowed to cool. The products were collected by acetone sedimentation combined with centrifugation. The final products were dispersed in ethanol and stored at room temperature for further characterization. The influence of I was also explored in the mixture system of Cl and I using a similar procedure.

2.3. Density Function Theory Calculation

Density functional theory (DFT) calculations were performed by using the Vienna Abinitio Simulation Package (VASP 5.4.4, Hafner Group at the University of Vienna, Vienna, Austria). The kinetic energy cutoff of electron wave functions was set as 450 eV. The convergence of energy standard for electronic self-consistent processes was set to be 1 × 10−6 eV, and the convergence criteria for force in geometric structure optimization was set to be 0.015 eV Å−1. The adsorption models of different halogens on Ag(111) and Ag(100) crystal planes were constructed, respectively. After structural optimization, the adsorption energy of halogen in the structural model was calculated as follows:
E a d s = E s l a b a d s E a d s E s l a b
where E a d s represents the adsorption energy of halogens on different crystal planes of silver and E s l a b , E a d s , and E s l a b a d s represent the energy of the initial, halogen, and the adsorbed structure, respectively.

2.4. Characterization

The morphologies and microstructures of as-prepared samples were investigated by field emission scanning electron microscopy (FE-SEM, Verios G4, Thermo Fisher Scientific, Waltham, MA, USA) and transmission electron microscopy (TEM, Tecnai G2 F20, FEI, Hillsboro, OR, USA). The crystal structure of products was characterized by X-ray diffraction (XRD, D8 Advance, Bruker, Karlsruhe, Germany). The ultraviolet–visible (UV–vis) extinction spectra of the AgNWs suspensions in ethanol were recorded by an optical spectrometer (UV2600i, Shimadzu, Kyoto, Japan). The surface chemistries were characterized by X-ray photoelectron spectra (XPS) on an X-ray photoelectron spectrometer (ESCALAB 250 Xi, Thermo Fisher Scientific, Waltham, MA, USA) with results being calibrated by C1s at 284.8 eV. The zeta potential was measured by Zeta potential analyzer (Zetasizer-Nano-ZS90, Malvern, UK). The specific surface area was tested using the nitrogen adsorption and desorption isotherms on a specific surface area analysis instrument (ASAP 2460, Micromeritics, Norcross, GA, USA), with the surface area being calculated by the Brunauer–Emmett–Teller (BET) method. The diameter and length of 200 wires were calculated from the SEM images for each sample, with the aid of the Nano Measurer 1.2 software (Fudan University, Shanghai, China). In particular, in order to measure the actual length of the wire, the final products dispersed in ethanol were subtly diluted for preparing SEM samples, so that most of the boundaries between nanowires could be clearly distinguished. Moreover, some bended nanowires are divided into several segments and each segment was separately calculated to decrease the measurement errors.

3. Results and Discussion

3.1. Growth Characterization of AgNWs in Halogen-Containing Systems

The co-addition of Br and Cl results in the higher aspect ratio of AgNWs than both single Br and single Cl systems. The total halogen ions content was kept the same for all experiments by controlling the added amount of FeCl3 and FeBr3. Typical XRD patterns and TEM images are shown in Figures S1 and S2, proving the successful formation of AgNWs. As shown in Figure 2a, the average length of AgNWs was only 45.1 μm for single Cl system (labeled as 0% in Figure 2), which is much lower than 72.3 μm of 33.3% Br, 65.9 μm of 50% Br, and 60.8 μm of 66.7% Br. The results imply that the co-addition of Cl and Br can facilitate the growth of the Ag(111) crystal planes, resulting in AgNWs with larger length.
The diameter shows a similar trend as shown in Figure 2b,c. Specifically, the average diameter of AgNWs at 66.7% Br system is much lower than the single Cl system (59.0 nm vs. 71.9 nm), pointing to restrained growth of the Ag(100) crystal plane when 66.7% Cl was replaced by Br. It should be mentioned that both the length and diameter present the smallest values for single Br system in all tested samples, showing that the growth of both the Ag(111) and Ag(100) are the slowest. Resultantly, the AgNWs with the highest value of 1031 was obtained in the 66.7% Br system.
The above results are further manifested by the morphology observations and UV–vis extinction spectra as well as the specific surface area comparisons. As shown in Figure 3a–c, the low-magnification SEM images present a much shorter length of AgNWs for 100% Br system (Figure 3c) than that of both 0% Br system (Figure 3a) and 66.7% Br system (Figure 3b), further proving that the growth of Ag(111) crystal plane is restrained in the single Br system [29]. Correspondingly, the inhibited growth of the Ag(100) crystal plane in both 66.7% Br system and 100% Br system were again validated by the high-magnification SEM images (Figure 3d–f) [29], as proved by the much thinner nanowires for these two cases when compared with that of 0% Br system (Figure 3d). The small difference of distribution density for AgNWs in different SEM images is mainly caused by the discrepancy of the dispersed concentration of AgNWs during the SEM sample preparation process, rather than the difference of product yield in various conditions. The yield of AgNWs is located in the range of 30~60%, depending on the added halogen ions, being consistent with the reported results [2,29,36]. The above characteristics for the diameter of the AgNWs were also revealed by the UV–vis extinction spectra (Figure 3g). In the UV–vis extinction spectra of AgNWs, lower wavelength of the absorption peak means thinner nanowire [2,43], which clearly indicate the diameter obeys the order of 0% Br > 66.7% Br > 100% Br (Figure 3g), again validating the inhibition effect of Br for the growth of the Ag(100) crystal plane. In addition to the UV–visible extinction spectra, the comparison of the specific surface areas of AgNWs also embodies the above morphology features. The specific surface area (S) of AgNWs can be determined by its length and diameter ( S = K L + K d , where K is a constant, L is the length of AgNWs, and d is the diameter of AgNWs). When the average length and diameter were substituted for calculation, the specific surface area of AgNWs could be ranked as S(100% Br) > S(0% Br) > S(66.7% Br), which is consistent with the detected BET values in Figure 3h, again manifesting the reliability of the above-mentioned morphology characteristics.
The influence of I on the growth of AgNWs is also explored in the mixture system of Cl and I. In the I ions dominated conditions (50–100% I), Ag nanoparticles are the only products without the generation of AgNWs. The generation of Ag nanorods (only ~0.74 μm) with an extremely small aspect ratio of 15.1 is observed only when the I is as low as 33.3%. The results indicate that the existence of I can restrain the formation of AgNWs, being consistent with the reported results [44,45].
In a word, the growth characteristics of AgNWs in different halogen-containing systems were understood. Compared with single Cl system, replacing 66.7% Cl with Br can promote the growth of Ag(111) crystal plane while the inhibition of the growth of the Ag(100) crystal plane at the same time, contributing to substantially increased aspect ratio of the AgNWs from 627 to 1031.

3.2. Analysis of Surface Physicochemical Characteristics of AgNWs

Physicochemical analysis reveals that Cl and Br show huge differences of adsorption behavior on the surface of the AgNWs. As shown in Figure 4a, in FeBr3 solution, the zeta potential of added AgNWs is 50.7 mV, which is much higher than 38.6 mV in the FeCl3–FeBr3 mixed solution and 30.9 mV in FeCl3 solution. The decrease of zeta potential is due to the adsorption of halogen anions (Cl or Br) on the surfaces of AgNWs, which means that the least amount of halogen is adsorbed in FeBr3 solution as revealed by the highest zeta potential value (Figure 4a). The adsorbed halogen anions can propel the accumulation of Ag+ on the surface of AgNWs because of the extreme low solubility of silver halides [2,35,36,43,46]. The low amount of Ag+ on the AgNWs surfaces in single Br system lead to slow growth rate, resulting in very thin and short nanowires (see Figure 3a–f) as well as weak crystallinity, as revealed by the lowest peak intensity in the XRD pattern of AgNWs obtained in the 100% Br system (Figure 4b). The similar peak intensity for AgNWs obtained in 0% Br system and 66.7% Br system (Figure 4b) are consistent with the low zeta potential gap between FeCl3 solution and FeCl3–FeBr3 mixture solution (Figure 4a), also manifesting the influence of halogen ions adsorption on the growth of AgNWs. The change of zeta potential of AgNWs from the negative value (−22.8 mV) in blank solution to positive values (30.9~50.7 mV) in iron halides-containing solutions is due to the adsorption of Fe3+ cations.
The surface chemistry analysis by XPS technique also reveals the different adsorption behavior between various preparation systems. As shown in Figure 4c, compared with the AgNWs in 100% Br system (blue line), positive shifts of 0.2 eV of Ag 3d peaks were observed for AgNWs in the 66.7% Br system and 0% Br system. The results indicate that the AgNWs in the later two cases are inclined to lose more electrons, meaning stronger interactions with the surface halogen ions (Figure 4d,e) [47,48].
In this system, the added Fe3+ can be reduced to Fe2+ by ethylene glycol, which can efficiently remove the adsorbed oxygen atoms, facilitating the formation of multiply twinned seeds. The generated multiple–twinned seeds are critical for forming AgNWs [41]. Similarly, other cations, such as Cu2+ and Mn2+, can also induce a similar effect, also being used for promoting the formation of AgNWs [35,49,50]. The cations of Na+ and K+ cannot induce such an effect, therefore needing a deoxygenation procedure by inert gas during the production of AgNWs [29,49,51].
The physicochemical characteristics on the surface of AgNWs play an important role in determining the growth behavior of AgNWs. Replacing Br with Cl can promote the adhesion of more halogen ions to the surfaces of AgNWs, contributing to faster growth and higher crystallization degree. Additionally, the interaction of AgNWs with the surface halogen can be enhanced by tuning the relative ratios of added Br and Cl.

3.3. DFT Calculation for Understanding AgNWs-Halogen Interactions

The interactions between AgNWs and halogen ions were further revealed by DFT calculations for further understanding the preferential growth of specific crystal planes in different halogen-containing systems. As shown in Figure 5a, the adsorption energy of halogen ions on Ag(111) crystal plane is the strongest (−1.16 eV) for the Br–Cl mixture system (labeled as Br–Cl), which implies the largest number of Ag+ is accumulated [2,35,36,43,46]. Correspondingly, the Ag(111) crystal plane of AgNWs in Br–Cl system possesses the fastest growth rate among the three halogen-containing systems, being consistent with the morphology analysis in Figure 3a–c and length trend in Figure 2c. The adsorption energy of halogen ions on the Ag(100) crystal planes (Figure 5a) obey the order of single Cl system (labeled as Cl–Cl, −1.68 eV) > mixed Br–Cl system (labeled as Br–Cl, −1.56 eV) > single Br system (labeled as Br–Br, −1.05 eV), which means the diameter size of the AgNWs follows the similar trend, being in accordance with the diameter characteristics in morphology results (Figure 3d–f) and UV–vis extinction spectra (Figure 3g).
The above results are further revealed by the partial density of states (PDOS) and Bader charge distributions. As shown in Figure 5b, the density state of silver atoms on Ag(100) crystal plane in single Cl system (labeled as Cl–Cl) is the closest to the fermi energy among the three calculated models, which means that the silver atoms on Cl–Cl model is the most active for adsorbing the halogen ions. Similarly, the density state of silver atoms on the Ag(111) crystal plane in the Br–Cl mixture system (labeled as Br–Cl) is the closest to the fermi energy among the calculated three models (Figure 5c), which indicates that AgNWs in the Br–Cl mixed solution possesses the highest growth rate in the length direction, being in accordance with the results in Figure 2c. The Bader charge distributions also reveal the above results (Tables S1 and S2, Figure 5d,e). Among the three calculated models, the electron number being transferred from the Ag(100) crystal planes to the halogen atoms is the largest for the single Cl system (1.02 e, Figure 5d), which means the single Cl system promises the strongest adsorption of Ag(100) to the halogens, resulting in the fastest growth of diameter. Correspondingly, the electron number being transferred from the Ag(111) crystal planes to the halogen atoms is the largest for the Br–Cl mixed system (0.95 e, Figure 5e), which indicates the Br–Cl mixed system results in the strongest interactions between the Ag(111) and halogen atoms, contributing to the highest growth rate of the length.
The accumulation of electrons around the halogen atoms are also revealed by the local charge density difference. As shown in Figure 6a for the Ag(100) crystal planes, more electrons are accumulated around the halogens in the Cl–Cl model (yellow isosurfaces) than that in both the Br–Br model and Br–Cl model, which implies the single Cl system results in the strongest interactions between halogen atoms and the Ag(100) crystal plane, being consistent with the discussions in Figure 5. Correspondingly, the mixed Br–Cl system contributes to the most amount of electron accumulations around the halogen atoms on the Ag(111) crystal plane surface, as shown in Figure 6b, being in accordance with the adsorption energy and Bader charge distributions in Figure 5. The above results again embody that the single Cl system promotes the fastest growth of the Ag(100) crystal plane, while the Br–Cl mixed system contributes to the highest growth rate of the Ag(111) crystal plane.
Conclusively, compared with the single Cl system, the co-addition of Br and Cl can enhance the adsorption of halogens on the Ag(111) crystal planes while weaken the adsorption of halogens on the Ag(100) crystal planes. The adsorbed halogen ions contribute to the concomitant accumulation of Ag+ due to the extreme low solubility of silver halides. Resultantly, the aspect ratio of AgNWs prepared in the Br–Cl mixed solution is much higher than that in both single Br solution and single Cl solution. Correspondingly, the adsorption energy of halogen atoms on both the Ag(100) and Ag(111) planes are the lowest in the single Br system, resulting in AgNWs with the thinnest diameter and shortest length. The above analysis was further embodied by the PDOS and Bader charge distributions as well as the isosurfaces of the local charge density difference.

4. Conclusions

The size regulation and mechanism analysis of AgNWs mediated by halogen ion composition were systematically conducted. The co-addition of Br and Cl can greatly increase the aspect ratio of the AgNWs. When Br accounts for 66.7% in total halogens, the aspect ratio of AgNWs reaches to the highest value of 1031. The physicochemical analysis, including zeta potential testing and XRD analysis as well as XPS results, all indicate that the morphology and size evolution of AgNWs growth are highly dependent on the interactions of halogen ions with the AgNWs. The DFT calculation results demonstrate that the co-addition of Br and Cl contributes to the highest adsorption energy of −1.16 eV for halogen atoms on the Ag(111) crystal planes, facilitating the strongest interactions between halogens and Ag(111) crystal planes. On the contrary, compared with the single Cl solution, the co-addition of Br and Cl decreases the adsorption energy of halogen atoms on the Ag(100) crystal planes from −1.68 eV to −1.56 eV, with the interactions between halogen atoms with the Ag(100) crystal planes being largely weakened at the same time. Resultantly, the enhanced preferential growth of Ag(111) crystal planes as well as the restrained growth of the Ag(100) crystal planes contribute to a much increased aspect ratio. The results herein can provide new insights for understanding the morphology and size evolution during the AgNWs preparation in the ethylene glycol system.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12152681/s1, Figure S1: XRD pattern of AgNWs synthesized with different halogen composition. All the peaks could be indexed to fcc silver. The blue line, red line, and black line represent the synthesis conditions of 100% Br, 66.7% Br, and 0% Br, respectively; Figure S2: TEM images of AgNWs obtained with 66.7% Br and 33.3% Cl. (a,b) Low magnification image of AgNWs; (c) high magnification image of AgNWs; (d) a zoom-in TEM image taken from the marked square area in image (c), demonstrating the lattice fringe with spacing of 0.237 nm; Table S1: The number of valence electrons for different halogens on Ag(100) crystal face in AgNWs; Table S2: The number of valence electrons for different halogens on Ag(111) crystal face in AgNWs.

Author Contributions

Conceptualization, N.X.; methodology, N.X. and Y.C.; software, W.W.; validation, S.Z.; formal analysis, D.T.; investigation, N.X. and Y.C.; resources, S.Z.; data curation, N.X.; writing—original draft preparation, N.X.; writing—review and editing, X.C.; visualization, H.C.; supervision, S.Z.; project administration, W.W.; funding acquisition, S.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (51874101).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

We gratefully acknowledge the Test Center of Fuzhou University for its support in testing instruments.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Patil, S.; Kate, P.R.; Deshpande, J.B.; Kulkarni, A.A. Quantitative understanding of nucleation and growth kinetics of silver nanowires. Chem. Eng. J. 2021, 414, 128711. [Google Scholar] [CrossRef]
  2. Wang, X.M.; Chen, L.; Sowade, E.; Rodriguez, R.D.; Sheremet, E.; Yu, C.M.; Baumann, R.R.; Chen, J.J. Ultra-uniform and very thin Ag nanowires synthesized via the synergy of Cl, Br and Fe3+ for transparent conductive films. Nanomaterials 2020, 10, 237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Zhang, X.; Liu, B.; Hu, C.; Chen, S.; Liu, X.; Liu, J.; Chen, F.; Chen, J.; Xie, F. A facile method in removal of PVP ligands from silver nanowires for high performance and reusable SERS substrate. Spectrochim. Acta A 2020, 228, 117733. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, C.; Xu, J.; Ke, S.; Li, Y.; Zhu, X.; Gu, X.; Kan, C.; Shi, D. Silver nanowires deposited on cellulose nanofibers/graphene oxide hybrid membranes as sandwich-structured films for optoelectronic and SERS applications. ACS Appl. Nano Mater. 2020, 3, 10844–10854. [Google Scholar] [CrossRef]
  5. Sun, J.; Zhou, W.; Yang, H.; Zhen, X.; Ma, L.; Williams, D.; Sun, X.; Lang, M.F. Highly transparent and flexible circuits through patterning silver nanowires into microfluidic channels. Chem. Commun. 2018, 54, 4923–4926. [Google Scholar] [CrossRef]
  6. Subburaj, S.; Arumugam, B.; Chen, S.M.; Chen, T.W.; Seetharam, A.; Ramaraj, S.K. Polyol synthesis of Ag Nanowires as an electrochemical sensor for the quantification of Carcinogenic Hydrazine. Int. J. Electrochem. Sci. 2021, 16, 210773. [Google Scholar] [CrossRef]
  7. Salvo-Comino, C.; Martin-Pedrosa, F.; Garcia-Cabezon, C.; Rodriguez-Mendez, M.L. Silver nanowires as electron transfer mediators in electrochemical catechol biosensors. Sensors 2021, 21, 899. [Google Scholar] [CrossRef]
  8. Zhu, X.; Xu, J.; Qin, F.; Yan, Z.; Guo, A.; Kan, C. Highly efficient and stable transparent electromagnetic interference shielding films based on silver nanowires. Nanoscale 2020, 12, 14589–14597. [Google Scholar] [CrossRef]
  9. Zhu, X.; Guo, A.; Yan, Z.; Qin, F.; Xu, J.; Ji, Y.; Kan, C. PET/Ag NW/PMMA transparent electromagnetic interference shielding films with high stability and flexibility. Nanoscale 2021, 13, 8067–8076. [Google Scholar] [CrossRef]
  10. Zhang, Q.; Jiang, D.; Xu, C.; Ge, Y.; Liu, X.; Wei, Q.; Huang, L.; Ren, X.; Wang, C.; Wang, Y. Wearable electrochemical biosensor based on molecularly imprinted Ag nanowires for noninvasive monitoring lactate in human sweat. Sens. Actuat. B Chem. 2020, 320, 128325. [Google Scholar] [CrossRef]
  11. Zhu, X.; Guo, A.; Xu, J.; Kan, C. The synthesis of silver nanowires with tunable diameters using halide ions for flexible transparent conductive films. CrystEngComm 2020, 22, 8421–8429. [Google Scholar] [CrossRef]
  12. Ran, Y.; He, W.; Wang, K.; Ji, S.; Ye, C. A one-step route to Ag nanowires with a diameter below 40 nm and an aspect ratio above 1000. Chem. Commun. 2014, 50, 14877–14880. [Google Scholar] [CrossRef]
  13. Khanarian, G.; Joo, J.; Liu, X.-Q.; Eastman, P.; Werner, D.; O’Connell, K.; Trefonas, P. The optical and electrical properties of silver nanowire mesh films. J. Appl. Phys. 2013, 114, 024302. [Google Scholar] [CrossRef]
  14. Mutiso, R.M.; Sherrott, M.C.; Rathmell, A.R.; Wiley, B.J.; Winey, K.I. Integrating simulations and experiments to predict sheet resistance and optical transmittance in nanowire films for transparent conductors. ACS Nano 2013, 7, 7654–7663. [Google Scholar] [CrossRef]
  15. Sorel, S.; Lyons, P.E.; De, S.; Dickerson, J.C.; Coleman, J.N. The dependence of the optoelectrical properties of silver nanowire networks on nanowire length and diameter. Nanotechnology 2012, 23, 185201. [Google Scholar] [CrossRef]
  16. Li, Y.; Yuan, X.; Yang, H.; Chao, Y.; Guo, S.; Wang, C. One-step synthesis of silver nanowires with ultra-long length and thin diameter to make flexible transparent conductive films. Materials 2019, 12, 401. [Google Scholar] [CrossRef] [Green Version]
  17. Cui, L.; Du, Z.; Zou, W.; Li, H.; Zhang, C. The in situ growth of silver nanowires on multi-walled carbon nanotubes and their application in transparent conductive thin films. RSC Adv. 2014, 4, 27591–27596. [Google Scholar] [CrossRef]
  18. Sun, X.Y.; Xu, F.Q.; Li, Z.M.; Zhang, W.H. Cyclic voltammetry for the fabrication of high dense silver nanowire arrays with the assistance of AAO template. Mater. Chem. Phys. 2005, 90, 69–72. [Google Scholar] [CrossRef]
  19. Zhang, C.; Li, C.; Si, X.; He, Z.; Qi, J.; Feng, J.; Cao, J. Single-crystalline silver nanowire arrays directly synthesized onto substrates by template-assisted chemical wetting. Materialia 2020, 9, 100529. [Google Scholar] [CrossRef]
  20. Zheng, X.J.; Jiang, Z.Y.; Xie, Z.X.; Zhang, S.H.; Mao, B.W.; Zheng, L.S. Growth of silver nanowires by an unconventional electrodeposition without template. Electrochem. Commun. 2007, 9, 629–632. [Google Scholar] [CrossRef]
  21. Wang, J.; Pan, S. Electrodeposition of vertically standing Ag nanoplates and nanowires on transparent conductive electrode using porous anodic aluminum oxide template. Nanotechnology 2017, 28, 425601. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, L.; He, C.; Li, J.; Guo, J.; Yang, D.; Wei, J. Green synthesis of silver nanowires via ultraviolet irradiation catalyzed by phosphomolybdic acid and their antibacterial properties. New J. Chem. 2013, 37, 2179–2185. [Google Scholar] [CrossRef]
  23. Lin, H.Z.; Ohta, T.; Paul, A.; Hutchison, J.A.; Demid, K.; Lebedev, O.; Van Tendeloo, G.; Hofkens, J.; Uji-i, H. Light-assisted nucleation of silver nanowires during polyol synthesis. J. Photochem. Photobiol. A 2011, 221, 220–223. [Google Scholar] [CrossRef]
  24. Wang, Z.; Chen, X.; Liu, J.; Zhang, M.; Qian, Y. Glucose reduction route synthesis of uniform silver nanowires in large-scale. Chem. Lett. 2004, 33, 1160–1161. [Google Scholar] [CrossRef]
  25. Bari, B.; Lee, J.; Jang, T.; Won, P.; Ko, S.H.; Alamgir, K.; Arshad, M.; Guo, L.J. Simple hydrothermal synthesis of very-long and thin silver nanowires and their application in high quality transparent electrodes. J. Mater. Chem. A 2016, 4, 11365–11371. [Google Scholar] [CrossRef]
  26. Sharma, D.; Rakshana, D.A.; Balakrishnan, R.M.; JagadeeshBabu, P. One step synthesis of silver nanowires using fructose as a reducing agent and its antibacterial and antioxidant analysis. Mater. Res. Express 2019, 6, 075050. [Google Scholar] [CrossRef]
  27. Liu, X.L.; Han, S.; Zhang, S.B.; Zhou, S.S.; Jiao, N.; Zhao, H.Y.; Li, J. One-step growth method of silver nanowires in aqueous environment. Mater. Res. Express 2020, 7, 095001. [Google Scholar] [CrossRef]
  28. Sun, Y.; Gates, B.; Mayers, B.; Xia, Y. Crystalline silver nanowires by soft solution processing. Nano Lett. 2002, 2, 165–168. [Google Scholar] [CrossRef]
  29. Li, B.; Ye, S.; Stewart, I.E.; Alvarez, S.; Wiley, B.J. Synthesis and purification of silver nanowires to make conducting films with a transmittance of 99%. Nano Lett. 2015, 15, 6722–6726. [Google Scholar] [CrossRef] [Green Version]
  30. Zhang, Y.; Guo, J.; Xu, D.; Sun, Y.; Yan, F. One-pot synthesis and purification of ultralong silver nanowires for flexible transparent conductive electrodes. ACS Appl. Mater. Interfaces 2017, 9, 25465–25473. [Google Scholar] [CrossRef]
  31. Han, M.; Ge, Y.; Liu, J.; Cao, Z.; Li, M.; Duan, X.; Hu, J. Mixed polyvinyl pyrrolidone hydrogel-mediated synthesis of high-quality Ag nanowires for high-performance transparent conductors. J. Mater. Chem. A 2020, 8, 21062–21069. [Google Scholar] [CrossRef]
  32. Jharimune, S.; Pfukwa, R.; Chen, Z.; Anderson, J.; Klumperman, B.; Rioux, R.M. Chemical identity of poly (N-vinylpyrrolidone) end groups impact shape evolution during the synthesis of Ag nanostructures. J. Am. Chem. Soc. 2020, 143, 184–195. [Google Scholar] [CrossRef]
  33. Parente, M.; van Helvert, M.; Hamans, R.F.; Verbroekken, R.; Sinha, R.; Bieberle-Hutter, A.; Baldi, A. Simple and fast high-yield synthesis of silver nanowires. Nano Lett. 2020, 20, 5759–5764. [Google Scholar] [CrossRef]
  34. Guo, Y.; Hu, Y.; Luo, X.; Lin, S.; Hu, J.; Liu, Y. Investigation into the role of poly(vinylpyrrolidone) in the growth of high aspect ratio silver nanowires. Inorg. Chem. Commun. 2021, 128, 108558. [Google Scholar] [CrossRef]
  35. Fahad, S.; Yu, H.J.; Li, W.; Liu, J.H.; Li, S.B.A.; Fu, J.C.; Ul Amin, B.; Khan, R.U.; Mehmood, S.; Haq, F.; et al. Synthesis of AgNWs using copper bromide as stabilizing agent and oxygen scavenger and their application in conductive thin flms. Mater. Chem. Phys. 2021, 267, 124643. [Google Scholar] [CrossRef]
  36. Rui, Y.; Zhao, W.; Zhu, D.; Wang, H.; Song, G.; Swihart, M.T.; Wan, N.; Gu, D.; Tang, X.; Yang, Y. Understanding the effects of NaCl, NaBr and their mixtures on silver nanowire nucleation and growth in terms of the distribution of electron traps in silver halide crystals. Nanomaterials 2018, 8, 161. [Google Scholar] [CrossRef] [Green Version]
  37. Sun, Y.; Mayers, B.; Herricks, T.; Xia, Y. Polyol synthesis of uniform silver nanowires: A plausible growth mechanism and the supporting evidence. Nano Lett. 2003, 3, 955–960. [Google Scholar] [CrossRef]
  38. Lau, K.S.; Chin, S.X.; Tan, S.T.; Lim, F.S.; Chang, W.S.; Yap, C.C.; Jumali, M.H.H.; Zakaria, S.; Chook, S.W.; Chia, C.H. Silver nanowires as flexible transparent electrode: Role of PVP chain length. J. Alloys Compd. 2019, 803, 165–171. [Google Scholar]
  39. Ge, Y.; Duan, X.; Zhang, M.; Mei, L.; Hu, J.; Hu, W.; Duan, X. Direct room temperature welding and chemical protection of silver nanowire thin films for high performance transparent conductors. J. Am. Chem. Soc. 2018, 140, 193–199. [Google Scholar] [CrossRef]
  40. Lei, B.W.; Wang, J.; Du, Y.G.; Zhang, K.L. Controlling the size of silver nanowires through one-pot polyol method with trace halide and its effect on kinetic process. Mater. Res. Express 2017, 4, 075052. [Google Scholar] [CrossRef]
  41. Wiley, B.; Sun, Y.; Xia, Y. Polyol synthesis of silver nanostructures: Control of product morphology with Fe (II) or Fe (III) species. Langmuir 2005, 21, 8077–8080. [Google Scholar] [CrossRef] [PubMed]
  42. Tang, X.; Tsuji, M.; Jiang, P.; Nishio, M.; Jang, S.M.; Yoon, S.H. Rapid and high-yield synthesis of silver nanowires using air-assisted polyol method with chloride ions. Colloids Surface A 2009, 338, 33–39. [Google Scholar] [CrossRef]
  43. Zhan, K.; Su, R.; Bai, S.; Yu, Z.; Cheng, N.; Wang, C.; Xu, S.; Liu, W.; Guo, S.; Zhao, X.Z. One-pot stirring-free synthesis of silver nanowires with tunable lengths and diameters via a Fe3+ & Cl co-mediated polyol method and their application as transparent conductive films. Nanoscale 2016, 8, 18121–18133. [Google Scholar] [PubMed]
  44. Toybou, D.; Celle, C.; Aude-Garcia, C.; Rabilloud, T.; Simonato, J.-P. A toxicology-informed, safer by design approach for the fabrication of transparent electrodes based on silver nanowires. Environ. Sci. Nano 2019, 6, 684–694. [Google Scholar] [CrossRef]
  45. Bhattacharya, M.; Mandal, A.; Chakraborty, S.D.; Maiti, A.; Maity, A.; Kuznetsov, D.; Mondal, P.; Senapati, D. Direct experimental observation of salt induced aspect ratio tunable PFPT silver-nanowire formation: SERS-based ppt level Hg2+ sensing from ground water. RSC Adv. 2016, 6, 45279–45289. [Google Scholar] [CrossRef]
  46. Chen, D.; Qiao, X.; Qiu, X.; Chen, J.; Jiang, R. Large-scale synthesis of silver nanowires via a solvothermal method. J. Mater. Sci. Mater. Electron. 2010, 22, 6–13. [Google Scholar] [CrossRef]
  47. Weng, W.; Wang, S.; Xiao, W.; Lou, X.W. Direct conversion of rice husks to nanostructured SiC/C for CO2 photoreduction. Adv. Mater. 2020, 32, 2001560. [Google Scholar] [CrossRef]
  48. Zhou, J.; Xiao, H.; Weng, W.; Gu, D.; Xiao, W. Interfacial confinement of Ni-V2O3 in molten salts for enhanced electrocatalytic hydrogen evolution. J. Energy Chem. 2020, 50, 280–285. [Google Scholar] [CrossRef]
  49. Korte, K.E.; Skrabalak, S.E.; Xia, Y. Rapid synthesis of silver nanowires through a CuCl- or CuCl2-mediated polyol process. J. Mater. Chem. 2008, 18, 437–441. [Google Scholar] [CrossRef]
  50. Prabukumar, C.; Bhat, K.U. Beneficial effect of manganese (II) ions on the morphology of polyol synthesised silver nanowires. Electron. Mater. Lett. 2020, 16, 264–275. [Google Scholar] [CrossRef]
  51. Zhang, P.; Wei, Y.; Ou, M.; Huang, Z.; Lin, S.; Tu, Y.; Hu, J. Behind the role of bromide ions in the synthesis of ultrathin silver nanowires. Mater. Lett. 2018, 213, 23–26. [Google Scholar] [CrossRef]
Figure 1. Proposed mechanism for the size regulation of AgNWs by halogen ions.
Figure 1. Proposed mechanism for the size regulation of AgNWs by halogen ions.
Nanomaterials 12 02681 g001
Figure 2. The histograms of length (a) and diameter (b) of AgNWs with Br ion content varied from 0% to 100% in total halogens. Changes in AgNWs diameter and length with Br ion content are summarized in (c). The aspect ratios of AgNWs with Br ion content are summarized in (d).
Figure 2. The histograms of length (a) and diameter (b) of AgNWs with Br ion content varied from 0% to 100% in total halogens. Changes in AgNWs diameter and length with Br ion content are summarized in (c). The aspect ratios of AgNWs with Br ion content are summarized in (d).
Nanomaterials 12 02681 g002
Figure 3. (af) SEM images of purified samples of AgNWs obtained with different Br content. (g) UV–visible extinction spectra of the final products. (h) The specific surface area of AgNWs.
Figure 3. (af) SEM images of purified samples of AgNWs obtained with different Br content. (g) UV–visible extinction spectra of the final products. (h) The specific surface area of AgNWs.
Nanomaterials 12 02681 g003
Figure 4. (a) Zeta potential of as-synthesized AgNWs in different solutions. Equal amount of halogen ions are added in FeCl3 solution, FeCl3–FeBr3 mixed solution, and FeBr3 solution; (b) XRD pattern of as-synthesized AgNWs. The high-resolution XPS regions of Ag 3d (c), Cl 2p (d), and Br 3d (e) for AgNWs synthesized at different halogen compositions: the blue line, red line, and black line represent the synthesis conditions of 100% Br, 66.7% Br, and 0% Br, respectively.
Figure 4. (a) Zeta potential of as-synthesized AgNWs in different solutions. Equal amount of halogen ions are added in FeCl3 solution, FeCl3–FeBr3 mixed solution, and FeBr3 solution; (b) XRD pattern of as-synthesized AgNWs. The high-resolution XPS regions of Ag 3d (c), Cl 2p (d), and Br 3d (e) for AgNWs synthesized at different halogen compositions: the blue line, red line, and black line represent the synthesis conditions of 100% Br, 66.7% Br, and 0% Br, respectively.
Nanomaterials 12 02681 g004
Figure 5. (a) Comparison of adsorption energies of halogens on Ag(100) and Ag(111) crystal planes. (b,c) Partial density of states (PDOS) of silver atoms on Ag(100) and Ag(111) crystal plane. (d,e) Bader charge distribution diagram of Ag(100) and Ag(111) crystal planes. The insets in (a) illustrates a schematic diagram of a single AgNWs.
Figure 5. (a) Comparison of adsorption energies of halogens on Ag(100) and Ag(111) crystal planes. (b,c) Partial density of states (PDOS) of silver atoms on Ag(100) and Ag(111) crystal plane. (d,e) Bader charge distribution diagram of Ag(100) and Ag(111) crystal planes. The insets in (a) illustrates a schematic diagram of a single AgNWs.
Nanomaterials 12 02681 g005
Figure 6. (a,b) Isosurfaces of local charge density difference for different models (yellow and cyan denotes 5.2 × 10−3 and −5.2 × 10−3 e per Å3 isosurfaces, respectively).
Figure 6. (a,b) Isosurfaces of local charge density difference for different models (yellow and cyan denotes 5.2 × 10−3 and −5.2 × 10−3 e per Å3 isosurfaces, respectively).
Nanomaterials 12 02681 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xiao, N.; Chen, Y.; Weng, W.; Chi, X.; Chen, H.; Tang, D.; Zhong, S. Mechanism Understanding for Size Regulation of Silver Nanowires Mediated by Halogen Ions. Nanomaterials 2022, 12, 2681. https://doi.org/10.3390/nano12152681

AMA Style

Xiao N, Chen Y, Weng W, Chi X, Chen H, Tang D, Zhong S. Mechanism Understanding for Size Regulation of Silver Nanowires Mediated by Halogen Ions. Nanomaterials. 2022; 12(15):2681. https://doi.org/10.3390/nano12152681

Chicago/Turabian Style

Xiao, Ni, Yinan Chen, Wei Weng, Xiaopeng Chi, Hang Chen, Ding Tang, and Shuiping Zhong. 2022. "Mechanism Understanding for Size Regulation of Silver Nanowires Mediated by Halogen Ions" Nanomaterials 12, no. 15: 2681. https://doi.org/10.3390/nano12152681

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop