Journal Pre-Proof: Materials Research Bulletin

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

Journal Pre-proof

Flexible multilayered transparent electrodes with less than 50 nm


thickness using nitrogen-doped silver layers for flexible heaters

Gayoung Kim , Jung Wook Lim , Jaehee Lee , Su Jae Heo

PII: S0025-5408(21)00500-6
DOI: https://doi.org/10.1016/j.materresbull.2021.111703
Reference: MRB 111703

To appear in: Materials Research Bulletin

Received date: 29 July 2021


Revised date: 12 November 2021
Accepted date: 21 December 2021

Please cite this article as: Gayoung Kim , Jung Wook Lim , Jaehee Lee , Su Jae Heo ,
Flexible multilayered transparent electrodes with less than 50 nm thickness using
nitrogen-doped silver layers for flexible heaters, Materials Research Bulletin (2021), doi:
https://doi.org/10.1016/j.materresbull.2021.111703

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2021 Published by Elsevier Ltd.


Flexible multilayered transparent electrodes with less than 50 nm thickness using

nitrogen-doped silver layers for flexible heaters

Gayoung Kim, Jung Wook Lim*, Jaehee Lee, and Su Jae Heo

Emerging Nano-Materials Research Section, Electronics and Telecommunications Research

Institute (ETRI), Daejeon 34129, Republic of Korea

Advanced Device Technology, University of Science and Technology (UST), Daejeon 34113,

Republic of Korea

Corresponding author: Jung Wook Lim

E-mail: limjw@etri.re.kr

Phone: +82-10-3446-3340

Fax: +82-42-860-6836

Highlights
 Al-doped ZnO electrodes are fabricated by an optimized nitrogen insertion method
 The multilayered films exhibit high optical transmittance (96.0%)
 The films maintain their sheet resistance even after 5,000 bending cycles
 The films perform better than conventional indium-tin-oxide-based films
 The films are promising for high-efficiency thin film heaters and optoelectronics

Graphical abstract

1
2
Abstract

In this study, Al-doped ZnO/Ag/Al-doped ZnO multilayered electrodes are fabricated on a

colorless polyimide substrate through an optimized nitrogen insertion technique, and

subsequently applied to transparent flexible thin film heaters. The resulting electrodes have

remarkable performance and durability. A completely continuous and smooth Ag thin film

with extremely small thickness is fabricated by optimizing the N2 gas flow rates at room

temperature. The multilayered films exhibit a high optical transmittance of 96.0% at a

wavelength of 550 nm and low sheet resistance of 8.5 Ω sq.−1. The bending test results show

that there is no significant change in the sheet resistance even after 5,000 bending cycles,

highlighting their mechanical flexibility. For its application in transparent flexible heaters, the

performance of the multilayered films exceeds that of conventional ITO-based films, as

confirmed by their higher saturation temperatures and flexibility, indicating their outstanding

thermal properties and high conductivity with small thickness.

KEYWORDS: A. thin films, A. oxides, B. mechanical properties, D. electrical properties, B.

optical properties.

3
1. Introduction

The flexible electronics industry has seen continuous expansion over the recent years and has

emerged as a next-generation industry due to the development in the fields of semiconductors

and display technologies.[1-7] Transparent electrode is a requisite for developing flexible

electronic devices. Among them, indium tin oxide (ITO) films have emerged as the most

widely used transparent electrode material.[8-10] However, ITO is limited by several

constraints, such as high cost due to the scarcity of indium, crack formation under mechanical

bending, and high-temperature deposition process.[11-13] As common flexible substrates,

such as polyethylene terephthalate, polyethylene naphthalate, and polyethersulfone, should

not be heated to temperatures above 200 °C, the high substrate temperature required for ITO

leads to difficulties during the deposition process.[14-16] Thus, it is important to improve

the flexibility of the transparent electrode materials at room temperature for the

advancement of flexible electronic devices.[17-19]

To replace ITO in flexible devices, various candidates have been investigated, such as metal-

based meshes,[20-22] metal nanowires,[23-25] carbon-based nanotubes,[26-28] graphene,[29-

31] and organic material-based conducting polymers.[32-34] Metal nanowire film-based

electrodes have been extensively investigated because their transmittance and sheet

resistance are comparable to that of ITO. However, the trade-off between the optical

and electrical properties, the roughness, and oxidation of the silver nanowires have

hindered their development.[35, 36] Therefore, transparent electrodes based on oxide–

metal–oxide (OMO) structures, which comprise a thin film metal inserted between transparent

oxide thin films, have been proposed as an alternative to ITO and other transparent

electrodes.[37-40] These electrodes can be roll-to-roll processed at room temperature for high

throughput mass production, and exhibit high stability and conductivity with a small thickness

4
(<100 nm).[41-43] Although OMO transparent electrodes have the potential to replace ITO,

they still suffer from the classic trade-off between optical transmittance and electrical

conductivity.[44] Therefore, we proposed asymmetric multilayered transparent electrodes to

enhance the optical transmittance and electrical conductivity by controlling the optical

interference, leading to the quasi-elimination of this trade-off.[45-49] To secure flexibility, it

is necessary to reduce the thickness of oxide films in multilayered transparent electrodes

because of their lower ductility compared to metal films. In addition, to ensure high flexibility

and transparency, it is advantageous to reduce the thickness of the metal layer, which also

contributes to the conductivity of the electrodes.

Ag layer tends to form isolated islands separated by voids during the early growth stages, that

is, the Volmer–Weber growth mode, due to the poor wettability of Ag on heterogeneous

substrates.[50-53] The Volmer–Weber growth mode causes a significant delay in the

formation of continuous Ag layers, resulting in a rough morphology and low conductivity.

Various methods have been introduced to secure the high flexibility and transparency of

OMO transparent electrodes by employing nucleation-inducing seed layers and optimizing the

gas conditions during the deposition process. Schubert et al. produced continuous, smooth,

and conductive Ag thin films using Ca, Al, and Au as the seed layers for Ag.[54] Wang et al.

produced AgOx by injecting a small amount of O2 into the metal layer during the production

of Ag thin film. Further, they noted that the conductivity and optical reflectivity decreases as

the oxygen flow rate increases.[55] Zhao et al. aimed to overcome these challenges by

replacing Ag with Cu, which has a higher wettability, as the metal in an OMO transparent

electrode, producing CuNx by injecting a small amount of N2 into the metal layer during the

Cu thin film deposition.[56] However, these studies are not suitable for application in flexible

electronic devices as the resulting electrodes exhibit a relatively low conductivity, metal

oxidation, target poisoning due to oxidation-based reactive sputtering, and reddish color due

5
to the intrinsic color of Cu, respectively.[57, 58] Therefore, in spite of the aforementioned

efforts, further innovative improvements on the flexibility and transparency of Ag layers in

multilayered electrodes are needed.

This study aims to develop a stable, cost-effective, and practical method for obtaining metal

thin films that can maintain their morphology at a low thickness in multilayered electrodes.

Although the transmittance was improved by introducing a thinner Ag film to the

multilayered transparent electrode in a previous study, island growth of Ag films could

not be suppressed, which necessitated the deposition of thick Ag films.[45] In this study,

we investigate the electrical, optical, and surface properties of Ag thin films under different

mixed gas conditions to improve the electrical and optical properties, and flexibility of

multilayered transparent electrodes. The N2-mixed inert gas is introduced to inhibit the 3D

island growth mode, thereby obtaining a flat film with smaller thickness. In addition, we

measure the resistance change of the optimized multilayered electrodes using several

laboratory-designed bending test systems to demonstrate the flexibility of the fabricated

transparent electrodes. The optical, electrical, and flexible properties of the multilayered

transparent electrodes are investigated as their thicknesses are reduced from 100 nm to less

than 50 nm. As a representative of the application of multilayered transparent electrodes, we

fabricate transparent flexible heaters that can be attached to glass windows using the

fabricated multilayered transparent electrodes, and investigate the correlation between the

heating performance of the fabricated multilayered OMO and ITO heaters.

2. Material and methods

2.1 Growth of Al-doped ZnO /Ag/ Al-doped ZnO multilayered thin films on glass and flexible

colorless polyimide substrates

6
Double-sided, optically polished, soda-lime glass substrates with a thickness of 1 mm and

area of 2 in2 were used as the substrates. Before the deposition of the multilayered thin films,

the substrates were cleaned in acetone followed by isopropyl alcohol in an ultrasonic bath for

10 min. The substrates were then rinsed with deionized water and dried with N2 gas. Colorless

polyimide (CPI, Kolon industry, LTD) substrates, which exhibit flexible properties, were used

for the bending tests. Al-doped ZnO /Ag/ Al-doped ZnO (AZO/Ag/AZO) multilayered thin

films were deposited by magnetron sputtering (iSAC Research, iRV dX3) without breaking

the vacuum. The sputtering chamber was initially evacuated to a base pressure of 8.0 × 10−7

Torr. The target-to-substrate distance was maintained at 15 cm. The substrate holder was

scanned from side to side during deposition at speeds of 1000, 2000, and 3000 rpm. The AZO

ceramic target (99.99%, 125 mm × 450 mm), with 2 wt.% Al2O3 and Ag target (99.97%, 125

mm × 450 mm) was employed as the sputtering target. AZO deposition was performed at 30

mTorr in a pure Ar atmosphere (99.99% purity) under a supplied power of 1 kW at room

temperature (26–30 °C) using the AZO target. Ag was deposited at 5 mTorr under a supplied

power of 250 W at room temperature using an Ag target. In the Ag deposition process,

optimized growth conditions were determined by exploring different gas mixtures of pure Ar

and N2 with varying sputtering parameters, such as oxygen and N2 gas flow rates.

2.2 Characterization of Ag thin films and AZO/Ag/AZO multilayered thin films

The electrical properties of the films were measured using a four-point probe (Advanced

Instrument Technology, CMT-100S) lined up at 1 mm intervals, following ASTM F390-09

standard. The electrical resistivity, carrier concentration, and mobility of the thin films with

indium contacts at the four corners were measured by the Van der Pauw Hall effect

measurements (Accent, HL5500IU Hall System). The optical properties of the electrodes and

substrates were measured using a UV-vis spectrometer (Perkin Elmer, Lambda 750) in the

7
wavelength range of 380–780 nm, according to ISO 20508 standard. The N2 levels were

determined using secondary ion mass spectrometry (CAMECA, IMS 7f) depth profiling

measurements at the National Nanofab Center (Daejeon, Republic of Korea). Surface nano-

images were obtained using an atomic force microscope (Park Systems, XE-100). X-ray

photoelectron spectroscopy measurements were performed using a VG-ESCA system with an

Al Kα X-ray source. The transmission electron microscopy (TEM) images were obtained

using a scanning TEM (JEOL, JEM-ARM200F) at an acceleration voltage of 200 kV. The

mechanical properties of the multilayered thin films were measured using a lab-made bending

system (Figure 5a, inset).

2.3 Evaluation of AZO/Ag/AZO-based thin film heaters

AZO/Ag/AZO-based thin film heater was fabricated on a 50 mm × 50 mm CPI substrate, and

Ag pastes were used at the edges of the multilayered thin film for a two-terminal contact. A

voltage of up to 10 DCV was applied to the AZO/Ag/AZO-based thin film heater through the

Ag contact electrode through a power supply. The temperature was measured using a k-type

thermocouple mounted on the surface of the AZO/Ag/AZO-based thin film heater. The

uniformity of the heat distribution was confirmed by observing infrared thermal images using

an infrared camera (Therm-App, TH). For the reliability test, a heating–cooling cycle was

conducted 10 times and the saturated temperature of the AZO/Ag/AZO multilayered thin film

heater was obtained for 1 h with a supplied voltage of 10 V.

3. Results and discussion

3.1 Influence of nitrogen doping during the sputtering process

8
Figure 1a illustrates the schematics of the apparatus for the film fabrication, which enabled a

continuous plate-to-plate sputtering process using AZO and Ag targets to fabricate

multilayered transparent electrodes with an AZO/Ag/AZO structure. In the continuous

sputtering process, the bottom AZO layer, Ag interlayer, and top AZO layer were sequentially

deposited on glass and 80 μm CPI substrates. As the sputter shared one load-lock chamber,

the transparent electrodes were deposited without a vacuum break. Figure 1a also shows the

interior of the chamber during the fabrication of the transparent electrodes. We proposed a

method to reduce the amount of target consumed to fabricate transparent electrodes with a

large area by decreasing the target size relative to the sample size, unlike conventional plate-

to-plate sputtering systems. The substrate (width 200 mm; length 200 mm) scanned the

bottom of a smaller target (width: 100 mm; length: 450 mm). This approach can be used for

the mass production of transparent electrodes during roll-to-roll sputtering deposition. As

shown in Figure 1b, the multilayered transparent electrodes coated on the CPI substrate have a

symmetric structure. The AZO layers reduced the reflection loss from the optical interference

with the Ag films, whereas the Ag films promoted the conductivity of the electrode. The

optimized multilayered transparent electrode exhibited high flexibility, transmittance, and

applicability to a large area, as shown in Figure 1c. The high transmittance and superior

flexibility of the OMO multilayered transparent electrodes coated on the CPI substrate

allowed for a clear observation of the objects behind the transparent electrode. Furthermore,

its high uniformity and reproducibility demonstrated the feasibility of the continuous roll-to-

roll process in fabricating transparent electrodes.

9
Figure 1. (a) Schematic of the continuous sputtering system for the fabrication of oxide–

metal–oxide (OMO) multilayered films on flexible colorless polyimide (CPI) substrates. (b)

Structure of the fabricated OMO multilayered transparent films on an 80 μm thick CPI

substrate. (c) Photograph of a 200 × 200 cm2 flexible and transparent OMO film. [AZO: Al-

doped ZnO]

Figure 2a illustrates the cross-sectional view of the Ag films grown at room temperature. The

electrical and optical characteristics of multilayered transparent electrodes rely on the surface

morphology of the thin metal films inserted between the oxide films. Conventional OMO

electrodes use a continuous and smooth Ag layer with a thickness greater than 10 nm, which

results in a low optical transmittance. To overcome this low transmittance while maintaining

high conductivity, it is necessary to inhibit the island growth of the Ag layer at a smaller

thickness. Therefore, a unique Ag:N growth mechanism was proposed to suppress the

Volmer–Weber growth mode, which is typical for pure Ag films. The proposed Ag:N layer

exhibited a completely flat and smooth morphology with a thickness as small as 9 nm,

10
whereas the pure Ag layer had a discontinuous and irregular granular morphology at an

identical thickness (Figure 2a). In this study, the N2 level of the Ag:N film was controlled by

introducing N2 gas mixed with Ar gas during the sputtering process to inhibit the uncontrolled

clustering and agglomeration in the Ag thin film, which increases its surface roughness. The

injection of N2 gas at gas flow rate ratios of 0%, 0.07%, 0.20%, and 0.47% into an Ar

atmosphere during the sputtering process decreased the surface roughness of the Ag thin films.

Figure 2. (a) Influence of N2 doping during the sputtering process. (b,c) Atomic force

microscopy (AFM) images of the 9 nm thick Ag films (b) without and (c) with N2 injection

during the sputtering process with a scan domain of 1 × 1 μm2. (d) Root-mean-square surface

roughness of the Ag films with increasing N2 flow rate determined from the AFM scans of at

least five different domains (1 × 1 μm2) for each condition. (e) Nitrogen doping level of the

Ag films determined by secondary ion mass spectrometry depth profiling. [AZO: Al-doped

ZnO]

11
The abrupt morphological transition from a highly granular and discontinuous pure Ag film

(N2 gas flow rate ratio of 0%) to a continuous and smooth Ag:N film (N2 gas flow rate ratio

0.07%) was confirmed by the atomic force microcopy images of the 9 nm thick Ag and Ag:N

films, as shown in Figure 2b and 2c, respectively. Furthermore, this morphological difference

was confirmed by the quantitative comparison of the root mean square surface roughness of

the Ag and Ag:N layers at various N2 gas flow rate ratios, as shown in Figure 2d. The root

mean square value of the Ag thin film fabricated without N2 gas injection was ~1.7 nm, which

was reduced to half (~0.8 nm) upon the feeding of N2 gas, due to the inhibited island growth

of the Ag films. The incorporation of N2 in the Ag film induced a dramatic change

resulting in the formation of agglomerated islands immediately after the nucleation

stage. Contrastingly, the pure Ag film exhibited the well-known 3D-island growth

behavior via uncontrolled cluster migration, reported for the noble metal-films on oxide

surfaces. The formation of such connections between adjacent islands can be attributed

to the morphological transformations that occur in order to reduce the surface energy.

The N2 is likely to form bridges between Ag islands, resulting in the formation of 9 nm

thick continuous Ag:N films. These films fill the voids at the grain boundary grooves,

decreasing the root mean square surface roughness value.[56] In addition, using

secondary ion mass spectrometry measurements, different amounts of N2 were detected in the

Ag and Ag:N films, as shown in Figure 2e and S1. The results confirmed the incorporation of

N2 in the Ag:N films at low concentrations. Further, we expect that any noticeable

degradation in the crystallinity of the metallic Ag films was not caused by the minimal doping

levels of N2.

12
3.2 Electrical and optical characteristic of the multilayered transparent electrodes

Figure 3a shows the electrical characteristics of the multilayered transparent electrodes with 9

nm thick Ag and Ag:N thin films, respectively, under various N2 flow rate ratios. Owing to

the anti-reflection effect generated in the multilayered structure, the multilayered transparent

electrodes demonstrated a high transmittance in the visible wavelength region. The

multilayered transparent electrode with an Ag:N layer (N2 gas flow rate ratio of 0.07%)

exhibited the lowest sheet resistance of 8.5 Ω sq.−1, while that of the electrode with a pure Ag

layer (N2 gas flow rate ratio of 0%) was 12.9 Ω sq.−1. The sheet resistance of the multilayered

transparent electrodes increased as the N2 flow rate ratio increased during the sputtering

process. These results demonstrate that the injection of N2 gas could achieve the formation of

a continuous smooth film with enhanced conductivity; however, excessive N2 gas injection

compromised the conductivity of the Ag films. Figure 3b shows the Hall measurement results

of the multilayered transparent electrodes prepared with Ag and Ag:N films at different N2

gas flow rate ratios. The lowest carrier mobility and highest carrier concentration of the

multilayered transparent electrodes were noted in the Ag films without N2 feeding. This low

mobility was related to carrier scattering because of the rough surface, grain boundaries, and

discontinuous and granular microstructure of the Ag thin films deposited without N2 gas

injection. In contrast, the increase in the N2 gas flow rate lowered the carrier concentration,

implying that the free electrons generated in the metal plasma were suppressed. The optical

characteristics of the multilayered transparent electrodes are compared in Figure 3c, where the

smooth continuous flat Ag:N thin films exhibited increased average transmittance in the

visible wavelength region owing to the minimal surface scattering. The N2 gas flow rate ratios

of 0%, 0.07%, 0.20%, and 0.47% achieved Ag and Ag:N thin films with the average

transmittances of 73.6%, 82.4%, 82.1%, and 80.5%, respectively, in the visible wavelength

regions. Particularly, the multilayered transparent electrode with an Ag:N layer deposited at

13
an N2 gas flow rate ratio of 0.07% exhibited the highest transmittance of 88.4% at a

wavelength of 550 nm (96.0%, excluding the substrates) owing to the formation of flat

surfaces.

Figure 3. (a) Sheet resistance and resistivity, and (b) mobility and carrier concentration of the

oxide–metal–oxide (OMO) multilayered films on a glass substrate at increasing N2 flow rates.

(c) Optical transmittance and (d) Figure of merit values calculated from the sheet resistance

and optical transmittance of the OMO multilayered films.

For a better comparison of the Ag films and Ag:N films deposited at various N2 gas flow rate

ratios, we used a widely accepted figure of merit (FoM), as shown in Figure 3d, which is

expressed as:

(1)

where T is the average transmittance in the visible wavelength region and Rsh is the sheet

resistance.[59] The multilayered transparent electrode with the optimized Ag:N layer (N2 gas

14
flow rate ratio of 0.07%) exhibited the highest FoM of 218.2, whereas the FoM of the

multilayered transparent electrode with pure Ag layer has the lowest value of 88.2. The

electrical and optical properties of the OMO multilayered transparent electrodes on a glass

substrate at different N2 gas flow rate ratios are summarized in Table S1. The electrical and

optical characteristics of the Ag and Ag:N thin films exhibited the same tendency as the

multilayered transparent films, as shown in Figure S2 and Table S2. Accordingly, the Ag:N

thin film deposited with an N2 gas flow rate ratio of 0.07% exhibited the highest FoM value of

70.6, whereas pure Ag thin film achieved the lowest FoM value of 10.7. Based on previous

studies, the optical characteristics of multilayered transparent electrodes with OMO structures

are superior to those of Ag thin films because of the optical interference of the upper and

lower AZO layers that covered the middle Ag thin film.[60] Further, the AZO thin film

inhibited surface plasmon coupling at the interface between the metal and oxide owing to its

large refractive index difference with the Ag thin film, thereby increasing the optical

transmittance in the multilayered OMO structure.

The effect of N2 gas injection during sputtering on the surface morphology is shown in Figure

4a. Irregular and broken islands grew on the substrate in the absence of N2 gas, resulting in

the formation of voids between the islands, which negatively affects the electrical

characteristics of the multilayered transparent electrodes. Moreover, the agglomeration of the

islands increased the surface roughness of the electrode, thereby compromising its optical

characteristics. In contrast, the injection of N2 gas during the sputtering process eliminated the

number of voids, achieving smooth and flat surfaces during film growth. As shown in Figure

4b, the cross-sectional TEM images of the fabricated multilayered transparent electrodes

demonstrated the fabrication of uniform Ag:N thin films with the injection of N2 gas during

the sputtering process. There were no remarkable defects, such as voids and discontinuous

regions, developed in the Ag:N layers. As AZO is an n-type material with a work function of

15
~5.0 eV, while Ag has a work function of 4.4 eV, the bonding of AZO with Ag leads to ohmic

contact and follows a linear current–voltage relationship that obeys Ohm’s law. Therefore, the

absence of voids in the metal thin film increased the conductivity of the multilayered

transparent electrode with OMO structures.

Figure 4. (a) Schematic of the sputtered Ag:N layer. (b) Transmission electron microscopy

images of the optimized oxide–metal–oxide (OMO) multilayered film with the injection of N2

gas during sputtering. [AZO: Al-doped ZnO]

A conventional roll-to-roll sputtering deposition method was modified to scan the bottom of a

target for large-area sputtering deposition (Figure 1a). The scan speed of the substrate along

the bottom of the target was varied between 1000, 2000, and 3000 rpm. The resulting

electrical and optical characteristics of the multilayered transparent electrodes were analyzed,

as shown in Figure S3. The electrical and optical properties of the OMO transparent

16
electrodes with 11 nm thick Ag thin film deposited without N2 gas injection (Figure S3a and

S3b) were compared with those with a 9 nm thick Ag:N thin film deposited with N2 gas

injection (N2 gas flow rate ratio of 0.7%) (Figure S3c and S3d). All transparent electrodes

exhibited a flat and smooth film, regardless of the presence or absence of N2 gas injection.

During the sputtering process, the thickness of the Ag thin films without N2 gas injection

increased from 9 nm to 11 nm and a smooth film was achieved, while the thickness of the

Ag:N thin films with N2 gas injection was maintained at 9 nm. The multilayered transparent

electrode with a pure Ag layer deposited without N2 gas injection exhibited relatively larger

deviations in the transmittance and sheet resistance at different scan speeds, as shown in

Figure S3, Table S3, and Table S4. In contrast, the multilayered transparent electrodes with

an Ag:N layer deposited with N2 gas injection exhibited relatively smaller deviations in the

transmittance and sheet resistance at varying scan speeds. Particularly, the best electrical and

optical properties were obtained at a scan speed of 2000 rpm. These results demonstrate that

Ag:N thin films are less sensitive to fluctuations of the scan speed, which is an important

variable in achieving uniformity over a large area. In addition, the injection of N2 gas

achieved a smooth and flat Ag thin film with low surface roughness; thus, adjustments in the

scan speed had a minimal effect on the surface characteristics. This aspect is highly preferred

for large-area processing and industrialization. Although pure Ag thin film deposited without

N2 gas injection exhibited an extremely low sheet resistance of 7.0 Ω sq.−1, the transmittance

was relatively low at 75.1%, resulting in a modest FoM value of 174.9. In contrast, the 9 nm

thick Ag:N thin film fabricated with N2 gas injection had a slightly higher sheet resistance of

8.5 Ω sq.−1 but a significantly better transmittance of 82.4%, exhibiting an improved FoM

value of 218.2. In addition, the electrical and optical characteristics of the Ag and Ag:N thin

films exhibited the same tendencies as those of the multilayered transparent electrodes, as

shown in Figure S4, Table S5, and Table S6.

17
To demonstrate the outstanding flexibility and durability of the multilayered transparent

electrodes coated on CPI substrates, we carried out mechanical tests using a laboratory-made

bending test setup. As shown in Figure S5 and Table S7, the electrical and optical properties

of the multilayered transparent electrodes coated on the flexible CPI substrate are almost

identical to those of the multilayered transparent electrodes coated on the glass substrate. The

difference in the sheet resistance, which is slightly higher for the multilayered transparent

electrode on the CPI substrate, is negligible as further reductions in the sheet resistance were

expected using suitable surface treatments for the CPI substrate. Further, the transmittance

and sheet resistance of the multilayered transparent electrodes with 9 and 13 nm thick Ag:N

thin films deposited with N2 gas injection during the sputtering process are compared, as

shown in Figure S6 and Table S8.

3.3 Flexibility of the multilayered transparent electrodes

The characteristics required for transparent electrodes in different applications can

significantly vary. For example, touch panel electrodes in a small area require a sheet

resistance of 400–700 Ω sq.−1, whereas large-area flat displays and thin-film solar cells

require a sheet resistance of <10 Ω sq.−1. Therefore, the characteristics of the multilayered

transparent electrodes fabricated in this study were analyzed by increasing the Ag:N thin film

thickness to investigate their applicability in various electronic devices. The flexibility of the

fabricated multilayered transparent electrodes with 9 and 13 nm thick Ag:N thin films was

compared with that of a conventional ITO transparent electrode with a sheet resistance of 10

Ω sq.−1, as shown in Figure 5a. The change in resistance was evaluated as the electrodes were

bent to different curvature radii. The ITO transparent electrode exhibited a change in

resistance up to a curvature radius of 20 mm. Particularly, the resistance sharply increased

18
rapidly from the curvature radius of 10 mm. In contrast, the multilayered transparent films

with a thickness of 110 nm (50 nm AZO/9 nm Ag/50 nm AZO) did not exhibit any

fluctuations in the resistance change even at small curvature radii of up to 5 mm,

demonstrating their superior flexibility. The resistance was also evaluated during continuous

repeated bending of the films from a curvature radius of 5 mm over 5,000 cycles, as shown in

Figure 5b. Cracks might be formed in the ITO transparent electrode at the beginning of the

cycling and the resistance exponentially increased. Meanwhile, the multilayered transparent

electrodes based on 9 and 13 nm Ag:N thin films did not exhibit any resistance changes

throughout the full 5,000 cycle test, indicating that there were no cracks formed at the surface

or interface of the fabricated multilayered transparent electrodes during mechanical testing.

The conventional ITO electrodes are not suitable for use in flexible devices as cracks are

formed easily in ITO because of its rigidity arising from its ionic bonds. The cracks in

ITO lead to electrical failure of devices. Therefore, the multilayered transparent

electrodes fabricated in this study are suitable for use in next-generation flexible

electronic devices.

19
Figure 5. (a) Resistance changes during the bending tests of the multilayered films with 9 and

13 nm thick Ag:N thin films, and 20 nm thick Al-doped ZnO (AZO) thin films coated on a

colorless polyimide substrate, and indium tin oxide (ITO) film at decreasing bending radius.

The inset shows the sample experiencing tensile strain. (b) Repeated bending tests as a

function of the bending cycles of the multilayered films with 9 and 13 nm thick Ag:N thin

films, and 20 nm thick AZO thin films. (c) Resistance changes during the bending test of the

multilayered films with 9 nm thick Ag:N thin films, and various thicknesses of the upper and

lower AZO thin films. (d) Repeated bending tests as a function of the bending cycles of the

multilayered films with 9 nm thick Ag:N thin films, and various thicknesses of the upper and

lower AZO thin films.

To achieve superior flexibility, decreasing the thickness of AZO films is essential in

multilayered transparent electrodes because of the lower ductility of AZO films compared to

metal films. Therefore, mechanical bending tests of the multilayered transparent electrodes at

various AZO thicknesses were conducted while maintaining the thickness of the Ag:N film at

20
9 nm. The optical and electrical properties of the multilayered transparent electrodes with

thicknesses of 110, 70, and 50 nm are shown in Figure S7 and Table S9, in which only minor

changes in the sheet resistance and transmittance were observed. Figure 5c shows the

resistance changes of the multilayered transparent electrodes as a function of the bending

radius during the bending test. Regardless of the thickness of the AZO thin films, the

multilayered transparent electrodes exhibited a constant resistance change even at a bending

radius of 3 mm. Repeated bending tests were performed at a constant bending radius of 3 mm,

as shown in Figure 5d. There was no resistance change in the OMO transparent films with 20

nm thick AZO thin films due to its superior flexibility, whereas the increase of the thickness

of the AZO thin films led to an increase in the resistance change, indicating the formation of

cracks and physical separation or delamination of the films from the substrate. A thin

metallic Ag layer with a thickness of 10 nm or less exhibits good flexibility, whereas

AZO, a hard ceramic material, has relatively low flexibility compared to the Ag layer.

Therefore, by reducing the thickness of AZO and maintaining a low resistance through

the metal Ag layer, excellent electrical and mechanical properties can be secured

simultaneously. Based on the bending test results, we found that a decrease in the oxide and

metal thickness, and formation of continuous and smooth Ag films of the multilayered thin

films play an important role in their application as flexible electrodes.

3.4 Thermal response of the multilayered transparent heaters

To investigate the viability of the multilayered transparent electrodes for flexible thin film

heaters, we fabricated typical transparent thin film heaters with a size of 50 mm × 50 mm on a

CPI substrate. To efficiently apply DC voltage to the multilayered transparent electrode, silver

paste electrodes were painted on the multilayered and ITO-based thin film heaters. The

21
temperature profile of the transparent film heaters was monitored using a thermocouple

mounted on the surface of the transparent and flexible film heaters. Heat was generated by

Joule heating when power was applied to the transparent film heaters. At the same time, small

amounts of heat were released by convection and radiation from the surfaces of the

transparent film heaters, neglecting conduction due to the extremely thin thickness of the CPI

substrate. The quantity of the generated and released heat eventually equalized and reached an

equilibrium temperature, where the temperature of the transparent film heaters was saturated.

In a previous study, we derived an expression for the time-dependent temperatures of heating

films as a function of electrical and thermal parameters, as expressed in Eq. 2,

( ) ( ) (2)

where V is the input bias voltage, R0 is the initial resistance, α is the temperature coefficient of

the resistance, ΔT is the temperature variation, hf is the convection coefficient of the films, A

is the surface area of the films, Tf is the film temperature, T0 is the environmental temperature,

ε is the surface emissivity, σ is the Stefan–Boltzmann constant (5.67 × 10−8 Wm−2K−4), Cf is

the heat capacity, Δt is the time variation, and Tf’ is the next temperature. Based on Eq. 2, the

proposed simulation results were compared with the experimental results, demonstrating their

high accuracy.

Figure 6a and 6b show the temperature variations of the multilayered and ITO-based thin

film heaters with respect to time when a 10 V DC bias was applied to the film heaters on the

CPI substrates for 120 s. The experimental and simulated data using Eq. 2 were in good

agreement. The curve fitting of the multilayered Ag:N films with 9 and 13 nm thick Ag:N

layers, and ITO-based thin film heaters showed convection coefficients (hf) of 35.7, 25, and

51.5 W m−2 K−1, respectively, which are directly related to the convection heat loss during the

heating and cooling processes. In other words, the thin film heaters with a higher convective

heat transfer coefficient result in a lower saturation temperature owing to the loss of more heat
22
at the same operating voltage. The convective heat transfer coefficient of the multilayered thin

film heaters is more than twice that of ITO-based thin film heaters. Further, the multilayered

film heaters achieved a saturation temperature of over 150 °C at a relatively low DC voltage

and large film area owing to their low convective heat transfer coefficient.

Figure 6. Thermal response and simulation results of the (a) multilayered film heaters with 9

and 13 nm thick Ag:N layers, and (b) indium tin oxide (ITO) film heaters using heat transfer

equations under an applied voltage of 10 V. (c) Heating–cooling temperature profiles of the

multilayered film heaters for 10 repeated cycles. (d) Durability of the multilayered film

heaters and their ability to maintain their saturation temperatures during heating for 1 h.

[OMO: oxide–metal–oxide]

To demonstrate the stability of the multilayered thin film heaters, heating and cooling tests of

10 consecutive cycles and a 1 h heating test were conducted, as shown in Figure 6c. The

23
saturation temperature and heating–cooling profiles of the multilayered thin film heaters were

maintained during the repeated cycle of the application and removal of 10 DCV. Figure 6d

shows the durability of the multilayered thin film heaters and their ability to maintain their

saturation temperature after heating for 1 h. These temperature profiles demonstrate that the

multilayered thin films are more suitable for flexible heater applications than conventional

ITO thin films.

3. Conclusion

In this study, we investigated the feasibility of ITO-free multilayered transparent films as

flexible transparent conducting electrodes for high-performance flexible thin film heaters. To

develop high-performance thin film heaters, the properties of the Ag thin films as a metal

layer in the multilayered thin films at varying N2 gas flow rate ratios during the sputtering

deposition were investigated. Based on the optical, electrical, morphological, and structural

properties of the multilayered transparent electrodes, we achieved an optimized multilayered

thin film with a high optical transmittance of 82.44% in the visible wavelength region

(95.99% at a wavelength of 550 nm without substrates) and low sheet resistance of 8.45 Ω

sq.−1. These superb properties were attributed to the effect of the N2 gas injection on the flat

surface and interface morphology of the Ag layer. Moreover, the multilayered thin films were

found to be mechanically stable with their sheet resistance maintained even after 5,000

bending cycles owing to the good interface between the Ag and AZO layers. The flexibility

improved by more than half by reducing the thickness of the multilayered transparent

electrodes from the conventional 100 nm to 50 nm. The temperature profiles of the

multilayered thin film heaters were compared with those of the ITO-based thin film heaters.

The multilayered thin film heaters reached high saturation temperatures over 150 °C due to

24
their lower sheet resistance and heat convection coefficient. Therefore, the optimized

multilayered thin films are a more promising candidate than conventional ITO transparent

electrodes in application in high-efficiency thin film heaters and optoelectronic devices owing

to their high transmittance, low sheet resistance, and flexibility.

Funding: This work was supported by Electronics and Telecommunications Research Institute

(ETRI) grant funded by the Korean government and the Korea Institute of Energy Technology

Evaluation and Planning (KETEP) grant funded by the Korea government Ministry of

Knowledge Economy (Grant No. 21ZB1160 and 20183010013820).

25
Declaration of interests
The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.

References

[1] P. Cataldi, F. Bonaccorso, A. Esau del Rio Castillo, V. Pellegrini, Z. Jiang, L. Liu, N.

Boccardo, M. Canepa, R. Cingolani, A. Athanassiou, Advanced Electronic Materials, 2

(2016) 1600245.

[2] S. Kim, H.Y. Jeong, S.K. Kim, S.-Y. Choi, K.J. Lee, Nano letters, 11 (2011) 5438-5442.

[3] Y. Wang, H. Guo, J.-j. Chen, E. Sowade, Y. Wang, K. Liang, K. Marcus, R.R. Baumann,

Z.-s. Feng, ACS applied materials & interfaces, 8 (2016) 26112-26118.

[4] Y. Wang, R. Yang, Z. Shi, L. Zhang, D. Shi, E. Wang, G. Zhang, ACS nano, 5 (2011)

3645-3650.

[5] Z. Zhong, H. Lee, D. Kang, S. Kwon, Y.-M. Choi, I. Kim, K.-Y. Kim, Y. Lee, K. Woo, J.

Moon, ACS nano, 10 (2016) 7847-7854.

[6] J. Gao, K. Kempa, M. Giersig, E.M. Akinoglu, B. Han, R. Li, Advances in Physics, 65

(2016) 553-617.

[7] W. Gao, L. You, Y. Wang, G. Yuan, Y.H. Chu, Z. Liu, J.M. Liu, Advanced Electronic

Materials, 3 (2017) 1600542.

[8] R. Gupta, K. Rao, K. Srivastava, A. Kumar, S. Kiruthika, G.U. Kulkarni, ACS applied

materials & interfaces, 6 (2014) 13688-13696.

[9] C. Guillen, J. Herrero, Thin solid films, 480 (2005) 129-132.

[10] T.M. Barnes, M.O. Reese, J.D. Bergeson, B.A. Larsen, J.L. Blackburn, M.C. Beard, J.

Bult, J. Van de Lagemaat, Advanced Energy Materials, 2 (2012) 353-360.

[11] T. Hauger, A. Zeberoff, B. Worfolk, A. Elias, K. Harris, Solar Energy Materials and

Solar Cells, 124 (2014) 247-255.

26
[12] Z. Chen, B. Cotterell, W. Wang, E. Guenther, S.-J. Chua, Thin Solid Films, 394 (2001)

201-205.

[13] Y. Leterrier, L. Medico, F. Demarco, J.-A. Månson, U. Betz, M. Escola, M.K. Olsson, F.

Atamny, Thin Solid Films, 460 (2004) 156-166.

[14] W. Wallace, J. Van Zanten, W. Wu, Physical review E, 52 (1995) R3329.

[15] J. Forrest, K. Dalnoki-Veress, J. Stevens, J. Dutcher, Physical review letters, 77 (1996)

2002.

[16] J. Forrest, K. Dalnoki-Veress, J. Dutcher, Physical Review E, 56 (1997) 5705.

[17] S. Yu, W. Zhang, L. Li, D. Xu, H. Dong, Y. Jin, Acta materialia, 61 (2013) 5429-5436.

[18] S. Yu, L. Li, X. Lyu, W. Zhang, Scientific reports, 6 (2016) 1-8.

[19] H.J. Kim, U. Kim, H.M. Kim, T.H. Kim, H.S. Mun, B.-G. Jeon, K.T. Hong, W.-J. Lee, C.

Ju, K.H. Kim, Applied Physics Express, 5 (2012) 061102.

[20] C. Zhang, A. Khan, J. Cai, C. Liang, Y. Liu, J. Deng, S. Huang, G. Li, W.-D. Li, ACS

applied materials & interfaces, 10 (2018) 21009-21017.

[21] X. Chen, W. Guo, L. Xie, C. Wei, J. Zhuang, W. Su, Z. Cui, ACS applied materials &

interfaces, 9 (2017) 37048-37054.

[22] X. Chen, X. Wu, S. Shao, J. Zhuang, L. Xie, S. Nie, W. Su, Z. Chen, Z. Cui, Scientific

reports, 7 (2017) 1-8.

[23] T. Wang, C. Luo, F. Liu, L. Li, X. Zhang, Y. Li, E. Han, Y. Fu, Y. Jiao, Langmuir, 33

(2017) 4702-4708.

[24] Y. Huang, Y. Tian, C. Hang, Y. Liu, S. Wang, M. Qi, H. Zhang, Q. Peng, ACS Applied

Nano Materials, 2 (2019) 2456-2466.

[25] H. Hu, S. Wang, S. Wang, G. Liu, T. Cao, Y. Long, Advanced Functional Materials, 29

(2019) 1902922.

[26] I. Jeon, J. Yoon, U. Kim, C. Lee, R. Xiang, A. Shawky, J. Xi, J. Byeon, H.M. Lee, M.

Choi, Advanced Energy Materials, 9 (2019) 1901204.


27
[27] I. Jeon, C. Delacou, H. Okada, G.E. Morse, T.-H. Han, Y. Sato, A. Anisimov, K.

Suenaga, E.I. Kauppinen, S. Maruyama, Journal of Materials Chemistry A, 6 (2018) 14553-

14559.

[28] A. Hussain, Y. Liao, Q. Zhang, E.-X. Ding, P. Laiho, S. Ahmad, N. Wei, Y. Tian, H.

Jiang, E.I. Kauppinen, Nanoscale, 10 (2018) 9752-9759.

[29] K.S. Kim, Y. Zhao, H. Jang, S.Y. Lee, J.M. Kim, K.S. Kim, J.-H. Ahn, P. Kim, J.-Y.

Choi, B.H. Hong, nature, 457 (2009) 706-710.

[30] S. Pang, Y. Hernandez, X. Feng, K. Müllen, Advanced Materials, 23 (2011) 2779-2795.

[31] Z. Liu, P. You, C. Xie, G. Tang, F. Yan, Nano Energy, 28 (2016) 151-157.

[32] H. Kang, S. Jung, S. Jeong, G. Kim, K. Lee, Nature communications, 6 (2015) 1-7.

[33] M. Vosgueritchian, D.J. Lipomi, Z. Bao, Advanced functional materials, 22 (2012) 421-

428.

[34] J.G. Tait, B.J. Worfolk, S.A. Maloney, T.C. Hauger, A.L. Elias, J.M. Buriak, K.D. Harris,

Solar Energy Materials and Solar Cells, 110 (2013) 98-106.

[35] M. Wu, H. Zheng, X. Li, S. Yu, Ceramics International, 46 (2020) 4344-4350.

[36] S. Yu, X. Liu, H. Dong, X. Wang, L. Li, Ceramics International, 47 (2021) 20379-20386.

[37] M. Theuring, M. Vehse, K. von Maydell, C. Agert, Thin Solid Films, 558 (2014) 294-

297.

[38] J. Yun, Advanced Functional Materials, 27 (2017) 1606641.

[39] W.S. Cheong, Y.H. Kim, J.M. Lee, C.H. Hong, H.Y. Choi, Y.J. Kwak, Y.J. Kim, Y.S.

Kim, Advanced Materials Technologies, 4 (2019) 1800550.

[40] S.H. Lee, G. Kim, J.W. Lim, K.-S. Lee, M.G. Kang, Solar Energy Materials and Solar

Cells, 186 (2018) 378-384.

[41] S.-H. Park, S.-M. Lee, E.-H. Ko, T.-H. Kim, Y.-C. Nah, S.-J. Lee, J.H. Lee, H.-K. Kim,

Scientific reports, 6 (2016) 1-12.

28
[42] T.-H. Kim, S.-H. Park, D.-H. Kim, Y.-C. Nah, H.-K. Kim, Solar Energy Materials and

Solar Cells, 160 (2017) 203-210.

[43] E.-H. Ko, H.-J. Kim, S.-J. Lee, J.-H. Lee, H.-K. Kim, RSC advances, 6 (2016) 46634-

46642.

[44] S. Yu, W. Zhang, L. Li, D. Xu, H. Dong, Y. Jin, Thin Solid Films, 552 (2014) 150-154.

[45] G. Kim, J.W. Lim, C. Yeon, T. Kim, H.C. Lee, S.J. Yun, Journal of Alloys and

Compounds, 774 (2019) 1092-1101.

[46] G. Kim, J.W. Lim, J. Kim, S.J. Yun, M.A. Park, ACS applied materials & interfaces, 12

(2020) 27122-27130.

[47] G. Kim, J.W. Lim, M. Shin, S.J. Yun, Solar Energy, 159 (2018) 465-474.

[48] J.W. Lim, G. Kim, M. Shin, S.J. Yun, Solar Energy Materials and Solar Cells, 163

(2017) 164-169.

[49] J.W. Lim, M. Shin, G. Kim, S.J. Yun, Solar Energy Materials and Solar Cells, 183

(2018) 92-100.

[50] G. Abadias, L. Simonot, J. Colin, A. Michel, S. Camelio, D. Babonneau, Applied Physics

Letters, 107 (2015) 183105.

[51] K. Ernst, A. Ludviksson, R. Zhang, J. Yoshihara, C. Campbell, Physical Review B, 47

(1993) 13782.

[52] M. Jose-Yacaman, C. Gutierrez-Wing, M. Miki, D.-Q. Yang, K. Piyakis, E. Sacher, The

Journal of Physical Chemistry B, 109 (2005) 9703-9711.

[53] M. Ohring, Materials science of thin films, Elsevier, 2001.

[54] S. Schubert, J. Meiss, L. Müller‐Meskamp, K. Leo, Advanced Energy Materials, 3

(2013) 438-443.

[55] W. Wang, M. Song, T.S. Bae, Y.H. Park, Y.C. Kang, S.G. Lee, S.Y. Kim, D.H. Kim, S.

Lee, G. Min, Advanced Functional Materials, 24 (2014) 1551-1561.

29
[56] G. Zhao, S.M. Kim, S.G. Lee, T.S. Bae, C. Mun, S. Lee, H. Yu, G.H. Lee, H.S. Lee, M.

Song, Advanced Functional Materials, 26 (2016) 4180-4191.

[57] S. Yu, L. Zhao, R. Liu, C. Zhang, H. Zheng, Y. Sun, L. Li, Solar Energy Materials and

Solar Cells, 183 (2018) 66-72.

[58] S. Yu, Y. Liu, H. Zheng, L. Li, Y. Sun, Optics letters, 42 (2017) 3020-3023.

[59] S.B. Sepulveda-Mora, S.G. Cloutier, Journal of Nanomaterials, 2012 (2012).

[60] K. Hong, K. Kim, S. Kim, I. Lee, H. Cho, S. Yoo, H.W. Choi, N.-Y. Lee, Y.-H. Tak, J.-L.

Lee, The Journal of Physical Chemistry C, 115 (2011) 3453-3459.

30

You might also like