• 沒有找到結果。

Morphological study on pentacene thin-film transistors: the influence of grain boundary on the electrical properties

N/A
N/A
Protected

Academic year: 2021

Share "Morphological study on pentacene thin-film transistors: the influence of grain boundary on the electrical properties"

Copied!
7
0
0

加載中.... (立即查看全文)

全文

(1)

This content has been downloaded from IOPscience. Please scroll down to see the full text.

Download details:

IP Address: 140.113.38.11

This content was downloaded on 25/04/2014 at 02:37

Please note that terms and conditions apply.

Morphological study on pentacene thin-film transistors: the influence of grain boundary on the

electrical properties

View the table of contents for this issue, or go to the journal homepage for more 2010 J. Phys. D: Appl. Phys. 43 405103

(http://iopscience.iop.org/0022-3727/43/40/405103)

(2)

J. Phys. D: Appl. Phys. 43 (2010) 405103 (6pp) doi:10.1088/0022-3727/43/40/405103

Morphological study on pentacene

thin-film transistors: the influence of

grain boundary on the electrical

properties

Fang-Chung Chen

1,3

, Ying-Pin Chen

1

, Yu-Jen Huang

1

and

Shang-Chieh Chien

2

1Department of Photonics and Display Institute, National Chiao Tung University, Hsinchu 30010, Taiwan 2Department of Photonics and Institute of Electro-optical Engineering, National Chiao Tung University,

Hsinchu 30010, Taiwan

E-mail:fcchen@mail.nctu.edu.tw

Received 25 June 2010, in final form 4 August 2010 Published 21 September 2010

Online atstacks.iop.org/JPhysD/43/405103

Abstract

We have prepared organic thin-film transistors (OTFTs) featuring pentacene molecules deposited at various substrate temperatures onto either hexamethyldisilazane (HMDS)- or poly(α-methylsyrene) (PαMS)-treated SiO2surfaces. As a result, we obtained different grain boundary densities in the conducting channel. Since the surface-modified devices featured similar grain boundary densities in their active layers, but displayed different electrical performances, we suspected that different trap states probably existed at the grain boundaries for the two different kinds of OTFTs. In addition, the surface morphologies of the initial layers featured grain boundaries that were rather blurred for the thin films prepared on the PαMS-treated substrates, whereas shallow boundaries appeared for the pentacene layers on the HMDS-treated surfaces. Therefore, we deduced that the different surface treatment processes resulted in different Schwoebel (step-edge) barriers, and hence, different morphologies. These results suggested that different trap states existed at the grain boundaries of the two types of surface-treated devices, leading to variations in the electrical performance, even though the grain boundary densities were similar.

(Some figures in this article are in colour only in the electronic version)

1. Introduction

Organic thin-film transistors (OTFTs) are promising candidates for use in next-generation electronics because they possess many advantageous properties, including low-cost fabrication, mechanical flexibility and light weight. Indeed, they have been incorporated into several potential devices, including smart cards, radio frequency identification tags and flexible displays [1–4]. Recent reports have described OTFT hole mobilities of greater than 1 cm2V−1s−1, suggesting electrical performance that can be compared with that of amorphous Si [4–7]. Among the various organic semiconductors, pentacene (C14H22), a planar aromatic molecule comprising five benzene rings, is 3 Author to whom any correspondence should be addressed.

one of the most promising candidates for use in p-channel devices [8,9]. Because of its high molecular symmetry, it has a strong tendency to form highly ordered thin films, which would, in principle, increase the hopping rates of charge car-riers between the molecules. Furthermore, the device perfor-mance of pentacene-based OTFTs is highly correlated with the pentacene layer’s morphology, which is significantly affected by many processing parameters, such as the surface proper-ties of the dielectric, the substrate temperature, the deposition rate and the basal pressure of the processing chamber [10–13]. For example, Knipp et al reported that an increase in dielectric roughness reduced the charge carrier mobility; meanwhile, a lower growth temperature led to smaller grains and influenced the trap distribution [12]. On the other hand, although many

(3)

J. Phys. D: Appl. Phys. 43 (2010) 405103 F-C Chen et al 0.000 0.004 0.008 0.012 -60 -40 -20 0 20 10-11 10-10 10-9 10-8 10-7 10-6 10-5 10-4 Si CH3 CH3 CH3 Si H3C N CH3 CH3 H HMDS [-I D (A)] 1/2 20OC 50OC 70OC 90OC

(a)

-ID (A) VG(V) -60 -40 -20 0 20 10-11 10-10 10-9 10-8 10-7 10-6 10-5 10-4 0.000 0.004 0.008 0.012 (b) -ID (A) [-I D (A)] 1/2 20OC 50OC 70OC 90OC VG (V) C CH3 C H2 n PaMS

Figure 1. Transfer characteristics of OTFTs featuring pentacene deposited at various substrate temperatures onto (a) HMDS- and (b) PαMS-treated substrates. The drain voltage (VD) was

maintained at−60 V.

reports have connected the morphological and structural prop-erties of pentacene thin films to the corresponding device per-formance, thorough investigations still remain rare [14,15].

In this study, we systematically investigated the influence of grain boundary on the performance of pentacene-based OTFTs. We used two types of modified dielectric surfaces to obtain different dielectric/channel interfaces and then deposited pentacene films onto these surfaces at various substrate temperatures. As a result, we obtained different grain boundary densities in the conducting channel. Since the surface-modified devices featured similar grain boundary densities in their active layers, but displayed different electrical performances, we suspect that different trap states probably existed at the grain boundaries for the two different kinds of OTFTs. Hence, to explain these observations we propose herein a pentacene growth pattern that takes into consideration the Schwoebel barrier effect on the dielectric surfaces.

2. Experimental details

Heavily n-doped Si wafers featuring a 200 nm-thick thermally grown SiO2 layer were used as substrates. The SiO2 surface was further modified with either hexamethyldisilazane (HMDS) or poly(α-methylsyrene) (PαMS) (see the insets to figure 1) to improve the device performance. HMDS modification was performed in an oven at 150◦C; no further post-annealing process was employed. On the other hand, PαMS modification was performed through spin-coating from

a dilute toluene solution (0.1 wt%); the resulting film was annealed at 80◦C for 1 h [16]. Next, pentacene (60 nm) was thermally evaporated to form the semiconductor layer. Using an infrared heater, the substrate could be heated from room temperature (20◦C) to 90◦C during the deposition processes in the thermal evaporator. The deposition rate was controlled at 0.5 Å s−1 using a quartz oscillator. Finally, Au was evaporated, through a shadow mask, to function as the source and drain electrodes. The channel width (W ) and length (L) were 2000 µm and 90 µm, respectively. Electrical characterization, at room temperature in an atmosphere of air, was performed using a semiconductor parameter analyzer (Keithley 4200). The surface morphology was measured using a Digital Instruments Dimension 3100 atomic force microscope. The x-ray diffraction (XRD) patterns were obtained using a thin film x-ray diffractometer (M18XHF-SRA, MAC Science) and Cu Kα (λ= 1.5406 Å) radiation.

3. Results and discussion

3.1. Electrical properties

Figure1displays the representative transfer characteristics of the OTFTs deposited on the HMDS- and PαMS-modified SiO2 surfaces at various substrate temperatures. The device mobility (µ) was extracted from the conventional field-effect transistor model in the saturation regime [17]:

ID= W Ciµ

2L (VG− VT)

2, (1)

where IDis the source–drain current, Ciis the capacitance per unit area of the dielectric, VG is the gate voltage and VT is the threshold voltage. Figure 2 summarizes the calculated mobilities. For the HMDS-treated devices, the mobility increased from 0.21 to 0.33 cm2V−1s−1 upon increasing the substrate temperature from 20 to 70◦C. The mobilities of the PαMS-treated devices increased similarly, reaching an even higher value of 0.72 cm2V−1s−1 at 70C, but decreased dramatically to 0.10 cm2V−1s−1when the substrate temperature was 90◦C. Meanwhile, the VT shift proceeded pronouncedly in the negative direction (figure 1(b)). We suspected that the morphology of the semiconducting layer changed significantly at such a high temperature.

3.2. Morphological study on the pentacene thin films

To study the effect of substrate temperature during the deposition of pentacene, we used atomic force microscopy (AFM) to observe the resulting surface morphologies. Figures 3 and 4 display AFM images of the 60 nm-thick pentacene layers deposited on the HMDS- and PαMS-treated substrates, respectively. Note that no stable films were obtained when the substrate temperature was above 110◦C, presumably because of a higher rate of pentacene desorption relative to absorption at that temperature. In general, the grain size increased upon increasing the deposition temperature for both kinds of devices. With larger amounts of thermal energy, the pentacene molecules were able to diffuse longer 2

(4)

10 30 50 70 90 0.2 0.4 0.6 0.8 Temperature (ºC) Mobility (cm 2/Vsec) PαMS-treated HMDS-treated Tg of PαMS

Figure 2. Variation in the mobility calculated from the I –V characteristics in figure1as a function of the substrate temperature for the pentacene-based OTFT devices prepared on the HMDS- and PαMS-modified SiO2substrates.

Figure 3. AFM images of 60 nm-thick pentacene films deposited at various deposition temperatures at a fixed flux rate (0.5 Å s−1) onto HMDS-treated SiO2substrates; image size: 3× 3 µm2. The

substrate temperatures were (a) 20, (b) 50, (c) 70 and (d) 90◦C.

on the dielectric surface, thereby having a higher probability of finding energetically preferred sites for nucleation and facilitating grain growth. For the PαMS-treated samples, although the grain size continued to grow upon increasing the temperature to 70◦C, it decreased when the substrate temperature reached 90◦C; interestingly, we observed poor connectivity between the grains (figure 4(d)). Because the glass transition temperature (Tg)of PαMS is ca 85◦C, the polymer chains became rubbery when the substrate temperature was greater than Tg. The higher level of thermal motion led to a shorter diffusion length of pentacene molecules, and therefore, diminished grain sizes. Accordingly, we observed a dramatic change in morphology when the pentacene thin film was deposited onto the PαMS-modified samples at a

Figure 4. AFM images of 60 nm-thick pentacene films deposited at various deposition temperatures at a fixed flux rate (0.5 Å s−1) onto PαMS-treated SiO2substrates; image size: 3× 3 µm2. The

substrate temperatures were (a) 17, (b) 50, (c) 70 and (d) 90◦C.

substrate temperature of 90◦C.

Next, we used XRD to examine the crystalline phase of the pentacene thin films. For the thin film deposited on the HMDS-treated substrate at 20◦C (figure 5(a)), we observed one set of diffraction peaks, having a d spacing of 15.4 Å. This set of peaks, indexed as (0 0 l) reflections, has been identified as a thin film phase of pentacene [18]. For higher substrate temperatures, another set of (0 0 l) peaks appeared; this second phase, having a d spacing of 14.4 Å, was consistent with the bulk phase of pentacene [18]. A similar phase transition has been reported for pentacene films deposited on bare SiO2 surfaces [18]. Previous reports have suggested that this change in phase is related to the relaxation of the initially strained thin film [19]. In contrast, the thin film phase of the pentacene crystals on the PαMS-treated substrate remained almost unchanged after increasing the substrate temperature (figure 5(b)). Thus, treatment with PαMS suppressed the phase transition. We inferred that the softer texture and flexibility of the PαMS polymer could aid in releasing the strain, thereby stabilizing the thin film phase. On the other hand, because the bulk phase was thermodynamically stable, the greater mobility of pentacene at higher substrate temperatures facilitated the molecules to reposition themselves to the lowest energy configuration [18]. Hence, some bulk phase was observed for the films deposited on the PαMS-treated substrates at higher temperatures.

3.3. Effect of grain boundaries on the device performance The effective mobility (µeff) in polycrystalline TFTs can be determined using the equation [20]

(5)

J. Phys. D: Appl. Phys. 43 (2010) 405103 F-C Chen et al

5

10

15

20

(003 ) (003 ') (002 ) (002 ') (001 ) (001 ')

70oC

50oC

90oC

20oC

log(Intensity) (a.u.)

2

θ

(º)

2

θ

(º)

(a)

5

10

15

20

(003 ) (003 ') (002 ) (002 ')

log(Intensity) (a.u.)

90oC 50oC 30oC 17oC (001) (001 ')

(b)

Figure 5. XRD spectra of 60 nm-thick pentacene films deposited at various deposition temperatures onto the (a) HMDS- and

(b) PαMS-treated substrates. The peak at a value of 2θ of ca 14◦ represents the diffraction of SiO2.

where µ0 is the bulk mobility and µGBis the grain boundary mobility. Thus, we know that the device mobility is related to the bulk and grain boundary status. Furthermore, if µGBis much less than µ0, the effective mobility will be close to µGB. In other words, the charge transport is limited mainly by the grain boundaries. Assuming that the charge transport at grain boundaries is governed by thermionic emission, µGBcan be further expressed as µGB= µGB0 exp  −EB kT  , (3)

where EBis the barrier height and µGB0is the trap-free mobility in the grain boundary. To see the effect of grain boundary, we developed a program (appendix) to estimate the ‘length’ of the grain boundaries from the AFM images in figures3and4and, thereby, to calculate the density of the grain boundary per unit area. From the data, we found, surprisingly, that the mobility was generally inversely proportional to the grain boundary

1

3

5

7

0.1

0.3

0.5

0.7

Mobility (cm 2 /Vs)

Grain Boundary density (µm-1) HMDS-treated PαMS-treated

90oC 90oC

Figure 6. Calculated mobility plotted with respect to the grain boundary density for devices prepared on the two types of surface-treated dielectric surfaces. The grain boundary density was calculated from the AFM images of the 60 nm-thick films in figures3and4.

density (figure6). This result suggests that the effect of grain boundary is the dominant factor in determining the degree of charge transport in the devices. In addition, the mobilities of the PαMS-treated devices were always greater than those of the HMDS-treated ones, even though the holes were being transported through the same number of grain boundaries. We inferred that the difference was due to the presence of different trapping states at the grain boundaries. Notably, we observed two exceptions to the devices’ performances: (i) for the HMDS-treated OTFT device prepared with pentacene deposition at 90◦C, when serious phase transition occurred, and (ii) for the PαMS-treated device prepared with pentacene deposition at 90◦C, when the polymer turned rubbery. In these two cases, other factors became important; the derivation from the general trend in figure6was, therefore, expected.

3.4. Analysis of the origins of the different traps states at the grain boundaries

To understand the origins of the different trap states at the grain boundaries of the two different surface-treated devices, figure7displays the morphologies of the corresponding 8 nm-thick pentacene films (ca five monolayers (MLs)) deposited at various temperatures. For the thin films deposited on the HMDS-treated substrates, we clearly observe grain boundaries within the first few MLs. On the other hand, for the pentacene layers deposited on the PαMS-treated substrates, many of the grain boundaries were too blurry to recognize, in particular for the thin film prepared at a substrate temperature of 70◦C. We suspected that the growth of the pentacene layers on the HMDS-treated substrates was similar to the ‘mound growth mode’, whereas that on the PαMS-treated substrates favoured the layer-by-layer mode [21,22] (figure8). Furthermore, the difference in the growth modes probably resulted from the presence of the Schwoebel barrier effect (see below) [21].

The growth mechanism of pentacene on various substrates has been investigated extensively. In addition to diffusion-limited aggregation in the lateral direction, some previous 4

(6)

Figure 7. AFM images of 8 nm-thick pentacene films deposited at various temperatures onto HMDS- (top panel) and PαMS-treated (bottom panel) SiO2substrates; image size: 3× 3 µm2.

Figure 8. Schematic representation of the structures of pentacene films on the (a) HMDS- and (b) PαMS-treated substrates. The higher Schwoebel barriers on the HMDS-treated surfaces resulted in deeper crevices.

reports have also indicated that the Schwoebel barrier disrupts the desired epitaxial growth of pentacene molecules [21]. The presence of such a step-edge barrier on a terrace will prevent the landing molecules from hopping down the edge. Hence, they prefer to remain on the islands, thereby leading to a three-dimensional (3D) growth mode. In short, the presence of such a barrier leads to a fractal shape. That is, the pentacene layers approach a dendrite-like morphology (figure7). In contrast, in the absence of such a Schwoebel barrier, the molecules will jump down to a lower layer, resulting in a two-dimensional (2D) growth mode [21,22].

Since the Schwoebel barrier effect can be strengthened or alleviated after surface treatment [21], we expected different phenomena to occur for the devices prepared on our two different substrates. Since a potential energy (Schwoebel) barrier existed in the HMDS-treated devices, the apparent uphill flux of the pentacene molecules led to fast upward

growth, causing the deep crevices that are apparent in figure8(a). On the other hand, for the PαMS-treated devices, either the Schwoebel barrier was relatively small or thermal energy could easily overcome the edge barrier. Therefore, we observed layer-by-layer growth of the pentacene molecules and the grain boundaries became relatively blurred (figure8(b)). From the schematic plot of the structures of the pentacene films, we deduced that the barrier height (EB) for the grain boundaries on the HMDS-treated surfaces was higher, because charge transport is believed to occur only in the first few layers of pentacene near the dielectric/semiconductor interfaces. In short, the two different growth modes led to grain boundaries possessing various barrier heights. As a result, the PαMS-modified devices exhibited higher mobilities, even though the grain boundary densities, which we measured from the surfaces of 60 nm-thick pentacene films, were the same as those of the HMDS-treated devices (figure6).

(7)

J. Phys. D: Appl. Phys. 43 (2010) 405103 F-C Chen et al

4. Conclusion

We have prepared pentacene thin-film transistors on either HMDS- or PαMS-treated SiO2surfaces at different substrate temperatures. AFM analysis of the surface morphologies revealed that the grain size increased upon increasing the substrate temperature. Electrical analysis indicated that the device performance was highly related to the grain boundary density. In addition, using AFM to investigate the surface morphologies of the corresponding initial layers (ca 5 MLs), we found that the grain boundaries were rather blurry for the thin films deposited on the PαMS-treated substrates, whereas shallow boundaries existed for the pentacene layers deposited on the HMDS-treated surfaces. We deduced that the different surface treatment processes resulted in different Schwoebel (step-edge) barriers, and hence, different morphologies. As a result, different trap states existed at the grain boundaries for the two different surface-treated devices, leading to dissimilar hole mobilities even though the grain boundary densities were similar. Our results suggest that surface treatment has a pronounced effect on the barrier heights of grain boundaries. We conclude that grain boundary is the dominant factor in determining the electrical performance for OTFT devices incorporating pentacene.

Acknowledgments

This study was supported by the National Science Council (grant no NSC 98-2221-E-009-028) and the Ministry of Education through the ATU Programme.

Appendix

Matlab Program to define the grain boundary density from the N*N matrix of AFM images:

function GB count (AFM file) rgb=imread (AFM file); rgb=imresize(rgb, [512 512]); r=rgb(:,:,1);

gray2log=(r>139); figure; imshow(gray2log);

[u,v]=meshgrid(-255:256, -255:256); % to define a space for Fourier transform

highpass=1-1./(1+(sqrt(u.ˆ2+v.ˆ2)/60).ˆ4); % highpass filter F=fftshift(fft2(double(gray2log)));

y=real(ifft2(fftshift(highpass.*F)));

Y=(y>0.1); % to show the pattern obviously

image i = Y;

dimension = size(image i);

target = 0; % to count how many white pixels is for i=1 : dimension(1,1)

for j = 1 : dimension(1,2) if image i(i,j) == 1 target = target + 1; end; end; end; ratio = target/(dimension(1,1)*4)

References

[1] Muccini M 2006 Nature Mater.5 605

[2] Zaumseil J and Sirringhaus H 2007 Chem. Rev.107 1296 [3] Sung C F, Kekuda D, Chu L F, Lee Y Z, Chen F C, Wu M C

and Chu C W 2009 Adv. Mater.21 4845

[4] Facchetti A, Yoon M H and Marks T J 2005 Adv. Mater. 17 1705

[5] Lin Y Y, Gundlach D J, Nelson S F and Jackson T N 1997 IEEE Electron Devices Lett.18 606

[6] Klauk H, Halik M, Zachieschang U, Schmid G, Radlik W and Weber W 2002 J. Appl. Phys.92 5259

[7] Roberts M E, Mannsfeld S C B, Queralt´o N, Reese C, Locklin J, Knoll W and Bao Z 2008 Proc. Natl Acad. Sci. USA105 12134

[8] Ruiz R et al 2004 Chem. Mater.16 4497

[9] Gundlach D J, Nichlos J A, Zhou L and Jackson T N 2002 Appl. Phys. Lett.80 2925

[10] Klauk H, Gundlach D J, Nichlos J A, Zhou L and Jackson T N 1999 IEEE Trans. Electron Devices46 1258

[11] Shtein M, Mapel J and Forrest S R 2002 Appl. Phys. Lett. 81 268

[12] Knipp D, Street R A and V¨olkel A 2003 Appl. Phys. Lett. 82 3907

[13] Chuang C S, Chen F C and Shieh H P D 2007 Org. Electron. 8 767

[14] Guo D, Ikeda S, Saiki K, Miyazoe H and Terashima K 2006 J. Appl. Phys.99 094502

[15] Lee H S, Kim D H, Cho J H, Park Y D, Kim J S and Cho K 2006 Adv. Funct. Mater.16 1859

[16] Chen F C, Chuang C S, Lin Y S, Kung L J, Chen T H and Shieh H P D 2006 Org. Electron.7 435

[17] Dimitrakopoulos C D and Malenfant P R L 2002 Adv. Mater. 14 99

[18] Dimitrakopoulos C D, Brown A R and Pomp A 1996 J. Appl. Phys.80 2501

[19] Lee J, Kim J H and Im S 2004 J. Appl. Phys.95 3733 [20] Horowitz G, Hajlaoui M E and Hajlaoui R 2000 J. Appl. Phys.

87 4456

[21] Zorba S, Shapir Y and Gao Y 2006 Phys. Rev. B74 245410 [22] Ruiz R, Nickel B, Koch N, Feldman L C, Haglund R F,

Kahn A and Scoles G 2003 Phys. Rev. B67 125406

數據

Figure 1. Transfer characteristics of OTFTs featuring pentacene deposited at various substrate temperatures onto (a) HMDS- and (b) PαMS-treated substrates
Figure 4. AFM images of 60 nm-thick pentacene films deposited at various deposition temperatures at a fixed flux rate (0.5 Å s −1 ) onto PαMS-treated SiO 2 substrates; image size: 3 × 3 µm 2
Figure 6. Calculated mobility plotted with respect to the grain boundary density for devices prepared on the two types of surface-treated dielectric surfaces
Figure 7. AFM images of 8 nm-thick pentacene films deposited at various temperatures onto HMDS- (top panel) and PαMS-treated (bottom panel) SiO 2 substrates; image size: 3 × 3 µm 2 .

參考文獻

相關文件

In section 4, based on the cases of circular cone eigenvalue optimization problems, we study the corresponding properties of the solutions for p-order cone eigenvalue

In this study, we compute the band structures for three types of photonic structures. The first one is a modified simple cubic lattice consisting of dielectric spheres on the

Basing on the observation and assessment results, this study analyzes and discusses the effects and problems of learning the polynomial derivatives on different level students

Teacher then briefly explains the answers on Teachers’ Reference: Appendix 1 [Suggested Answers for Worksheet 1 (Understanding of Happy Life among Different Jewish Sects in

“Polysilicon Thin Film Transistors Fabricated at 100℃ on a Flexible Plastic Substrate,” IEEE Electron Device Meeting, p. “Polysilicon Thin Film Transistors

Based on the different recreational choices of tourists, we obtain that under different fame effects the benefits of firms and tourists are different that result from the

Meanwhile, the study also explores the influence of students’ participation intention of Flipped Classroom based on their family background, learning achievement of

Deformation twins did not exhibit in the samples tested at 250°C and the recrystallized grain with a grain size of about 2.3 µm was observed at the