Paper V

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Effect of Ion Concentration and Interfaces on Electrical Parameters of OECT’s

Vijay Kr. Lambaa, Archana Kumarib


a
Department of Physics & Electronics Engineering, Global College of Engineering & Technology, Kahnpur Khui, Punjab, 140117
b
Department of Physics and Engineering, Slippery Rock University of Pennsylvania, PA, 16057
ABSTRACT
The organic electrochemical transistors are potential candidate for biosensing application devices. In this work
we tried to investigate the modulation of drain current as a function of two interfacial capacitances, one at the
gate-electrolyte and other at the channel-electrolyte interface. Also, how its electrical properties depend on the
ion concentration. We used Nernst equations, with Poisson-Boltzmann to describe doping-dedoping phenomena
in the conductive PEDOT: PSS layer and distribution of ions/holes in a device. We used analytical model to
simulate effect of ion concentration and conductivity profiles to study the dependence of drain current on
change in gate voltage. we used gradual channel approximation. While performing simulation we restrict
ourselves to fix dimensions of the channel. We investigated these general equation for the carbon fiber gate
electrode functionalized with Active carbon and Nafion powder with immobilized Antibody.
KEYWORDS: Organic Electrochemical Transistor, Ion Concentration, Interface,
INTRODUCTION
An organic electrochemical transistor (OECT) is an electrochemical device that shows many technological
prospective applications as biosensors to sense biomolecules[1], monitoring and stimulating cell activity[2, 3],
used in neuromorphic devices[4] as a basic structures. An OECT basically consisting of source, drain and gate
electrodes; conjugated polymer thin-film channel patterned between source and drain electrodes; an electrolyte
that is in contact between the gate electrode and the channel as shown in fig 1. Basically during operations, the
source is grounded, and a constant voltage is applied between the source and the drain, a voltage at the gate
electrode determines the doping state of the channel and used to modulate the conductance of the channel. The
channel modulation which was obtained by the change in the gate voltage is due to an interaction between
electronic and ionic charge carriers.

K+ Gate (C)
Cl-
Electrolyte(KCL)
Source(Au)
Drain (Au)

PEDOT:PSS Film

Figure 1: OECT Model

Our OECT model consist of thin film conducting polymer (P-type), poly (3,4-ethylene
dioxythiophene)(PEDOT)doped with poly(styrene sulfonate)(PSS). PEDOT:PSS films having conductivities in
the range of 300 to 1000 Scm -1 , highly stable in aqueous environment [5]. Extensive research has been done on
the modulation of electrical properties of channel material and effects of electrolytes [6-8], device geometry [9-
11], gate electrode materials [12]. Since basic working of OECTs can be explained on same principal that of
OFETs, because of the structural similarity between OECTs and OFETs. In OFETs, a current will flows
through the channel and is collected at the drain, when we apply a voltage (𝑉𝐷) between the source and drain
electrodes, which is similar to the output current of the OECT which is defined as drain current (𝐼𝐷). Since the
channel conductivity is independent of VD, at lower values, so according to ohms law the value of ID is
proportional to the VD. Further the ID can control gate electrode, so when a potential is applied at the gate
electrode with respect to source, an electric field is generated due to which ions from the electrolyte starts
drifting into the channel as a result conductivity of channel decrease or increase depending upon on the
properties of material of the channel. The conductivity of channel decreases if it is of p-type and positive gate
voltages is applied, and increases if negative gate voltages is applied, and vice-versa for n-type channel. In our
model we consider OECT as p-type, depletion-mode transistors
COMPUTATIONAL DETAILS
We used Carbon paper with activated carbon as a gate electrode with 0.01 M KCl, 20 mM Borate Buffer
solution, (pH 10) as an electrolyte with Gold electrodes as drain & source and PEDOT:PSS channel. When we
apply a +ve bias at the gate, electric field is generated, due to which cations (K+) moves from electrolyte into
the channel layer, which forms ionic bonds with PSS- as a result it moves from the PEDOT causing reduction of
PEDOT+ to its neutral state, PEDOTo which is insulating in nature, which is described(de-doping processes)[13-
14] by the equation 1 and is reversible in nature:
𝑃𝐸𝐷𝑂𝑇 � : 𝑃𝑆𝑆 � + 𝐾 � ↔ 𝑃𝐸𝐷𝑂𝑇 � + 𝐾 � : 𝑃𝑆𝑆 � + ℎ� (1)
when the gate voltage, V𝐺 = 0 V, the K+ ions diffuse out of channel as a result PSS forms ionic bonds with
PEDOT resulting to oxidation to move into conductive state, representing by left side of eq. 1. In case of 𝑉𝐺 > 0
K+ ions form ionic bonds with the sulphonate group, resulting in decrease of concentration of holes, H+, in
channel. Thus the current through channel is controlled by the gate electrode to adjust the redox state of the
channel.
we calculated the conductivity[15, 16] of the conducting polymer films deposited by the spin coating processes
from a mixture of PEDOT: PSS (Sigma Aldrich), dodecyl benzenesulfonic acid (DBSA), ethylene glycol, and
(3 -glycidyloxypropyl)-trimethoxysilane (GOPS) of different thickness by depositing in steps is shown in Table
1 using four probe method, which shows that the conductivity in general increases with increase in film
Thickness. On further analysis conductivity increases exponentially for thickness varying from 50 to 100 nm
and linearly further which is due to transition from percolation(at lower thickness) to bulk-like charge transport(
as thickness increases)
Table 1: Conductivity as a function of PEDOT:PSS Film Thickness
Thickness (nm) Sheet Resistance Conductivity(S) Thickness (nm) Sheet Resistance Conductivity (S)
50 600 330 300 76 535
100 200 440 350 64 550
150 130 486 400 53 567
200 103 502 450 40 588
250 90 519 500 28 603

Gate electrode is immersed into conductive solution, there for it is necessary to study charge transport at the
interface. The charge transport process depends on surface morphology of electrode, chemical reactions at
surface, interface between electrode and electrolyte, etc. Assuming, a case of idle polarizable, electrode, so no
net reaction take place on its surface. As a result charges will drift towards an interface and mount up there
when potential is applied. As a result, current’s flow depends on the circuit-resistance(by ohm’s law), since no
electrochemical reaction take place at electrode (by assuming, it as a idle polarizable, electrode) so we consider
the electrode-solution interface[17] as a capacitor, whose capacitance is proportional to stored charge and a
function of applied potential. Since the amount of charge in the electrolyte solution is equal to charge on gate
electrode. Further the drop of potential near the electrode surface as a result of presence of a diffuse double. So
���
the effective voltage(𝑉� ) is given by
���
𝑉� = 𝑉� + 𝑉������ (2)
Voffset is the offset voltage defined as, potential drop at the electrolyte-channel and electrolyte-gate interface.
So, the change in surface area of gate electrode, and concentration of KCl in electrolyte, alter the effective
���
voltage (𝑉� ), as a result value of drain current changes. So by applying a gate voltage, cations moves from
electrolyte into the channel layer causing change in the drain current, so maximizing the change of the drain
current (ID) upon the application of a gate voltage (VG) [ 18, 19], now varying the gate geometry, keeping
channel /electrolyte interface fixed the transfer characteristics are plotted (Figure 2)
0

-10

-20

-30 Area = 2 mm2


ID (μA)

-40 Area = .2 mm2


Area = .12 mm2
-50

-60

-70
-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1
VG (V)
Figure 2: Change drain current (ID) at different gate potential (VG) by varying gate area.

From figure 2 we conclude that change in drain current depends on the area of gate electrode, as area increases
change in drain current is large, hence the value of transcondatance i.e. ∂I D/∂VG, so we can use over device as
an ion‐to‐electron converter. As Electrolyte/channel interface modulates the drain current because some ions get
penetrated into channel region causing drop in potential.
For calculation, we assume uniform thick (= 200nm) PEDOT:PSS layer, forms an interface with electrolyte
with 0.1 M of KCl, and no ion penetration take place inside channel, as a result ions starts accumulating at the
electrolyte-channel interface. So the local potential is influenced by the concentration of PEDOT:PSS and is
given by the Nernst eq. ( 5 ) . Using Poisson-Boltzmann equation for KCl layer
∇. 𝜀∇𝜑 = −𝑒([𝐾 � ] − [𝐶𝑙� ]) (3)
And for conductive layer (PEDOT:PSS) is
∇. 𝜀∇𝜑 = −𝑒([𝑃𝐸𝐷𝑂𝑇 � ] − [𝑃𝑆𝑆 � ] ) (4)
[𝐾 � ] and [𝐶𝑙� ], is the local concentrations of K+, 𝐶𝑙 � ions in electrolyte; while [𝑃𝐸𝐷𝑂𝑇 � ], and [𝑃𝑆𝑆 � ] is the
local concentrations of PEDOT+, 𝑃𝑆𝑆 � ions in the conductive layer (PEDOT:PSS) , and 𝜑 is the potential.
Using Nernst equation concentration for PEDOT+ is
[����� � ]�
[𝑃𝐸𝐷𝑂𝑇 � ] = � (5)
��� �(�� � )/��

Were [𝑃𝐸𝐷𝑂𝑇 � ]� is initial concentrations of PEDOT+ ,


t is temperature (in absolute scale), F is Faraday
constant, R is universal gas constant. Here we assumed concentration of K+ ions, and of PSS- is constant and
using concept of global electro-neutrality[10,11,18], potential and concentration profiles is analytically
calculated using eq 3, 4, & 5. The potential profile along device length is represented by graph 3. Due to diffuse
- double layer near gate electrode, we have a drop of the potential, with a flat equi-potential region inside
electrolytefurther drops slightly below 0 and then rises again to 0 at the source electrode, which is due to the
reduction of PEDOT+ to PEDOTo due to potential(applied), near the electrolyte-channel interface. Because of
absence of charge neutrality, so the potential distribution in channel is due to uniform distribution of PSS- in
absence of PEDOT+ ions.

1
0V
0.8 0.2 V
0.4 V
Electric Potential (V)

0.6 0.6 V
0.8 V
0.4 1V

0.2

-0.2
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9

1.1
1.2
1.3
1.4
1.5
1.6
1.7
1.8
1.9
0

X Coordinate (μm)

Figure 3: Potential profile inside PEDOT:PSS layer at different Gate voltage (VG) (varying from 0 to 1 V).
Further the drain current (ID) is synchronized by electrochemical reactions in the channel.[22] Assuming that
each K+ ion eliminates one hole from the PEDOT:PSS, as a result, there is decrease in the concentration of
holes in channel, and analytically calculated[21] as:
��(�)
= 𝑐𝑊 [𝑉� − 𝑉 (𝑥)] (6)
��
��(�) �� � �(�)
𝑝(𝑥 ) = 𝑝� �1 − � = 𝑝� �1 − � (7)
�� ����� ��

Where, 𝑑𝑄/𝑑𝑥 is the cationic charge density as a function of distance 𝑥 along the channel away from the source
electrode, 𝑐 is defined as ionic capacitance per unit area, 𝑊 and T as width, and thickness of channel, 𝑉(𝑥) is
the voltage at position 𝑥, 𝑝0 is the initial hole density in channel, and 𝑉𝑃 is the channel pinch-off voltage
defined as 𝑉𝑃 = 𝑞𝑝oT/𝑐.
Figure 4 represents the effect of applied potential on concentration of PEDOT+ calculated using eq 5. Assuming
zero ion penetration and reduction of PEDOT+ to PEDOTo occurs at electrolyte/channel interface only. So if we
increase applied potential the pinch off length increases due to decrease in hole concentration inside channel,
which is further explained by eq 7.

0.8
[PEDOT ]/[PEDOT:PSS]

0.6 0V
0.2 V
0.4 V
+

0.4
0.6 V
0.8 V
0.2
1V

0
1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2
X Co-ordinate (μm)
Figure 4: Change in Normalized PEDOT+ concentration within channel at different Gate voltage, VG. (Varying
from 0 to 1 V)
DRAIN CURRENT CALCULATION
We used Gradual channel approximation [23,24], to calculate drain current, because there is a variation of
potential steadily from VGS (at the Source) to VGS - VDS (at the drain). Here, drain current is due to drift of holes
under the applied electric field only, hence the current in the channel is analytically calculated using Nernst-
Planck equation:
��
𝑗 = − �� 𝐷����� � . [𝑃𝐸𝐷𝑂𝑇 � ]. ∇𝜑 (8)

Here, 𝐷����� � is the diffusion coefficient of PEDOT+ and defined [25, 26] as
������ � �� �
𝐷����� � = (9)

Further conductivity 𝜎 of PEDOT:PSS layer is defined as


𝜎 = 𝑞[𝑃𝐸𝐷𝑂𝑇 � ]. 𝜇����� � (10)
𝜇����� � is the mobility of PEDOT+ ion, and q is its electrical charge, since potential being varying gradually
along the channel length, 𝜎 is material dependent and function of potential applied. Using differential form of
��
ohm’s law (𝑗 = −𝜎(𝜑) �� ) and combining it with eq 8, 9 and 10 we get

𝑗𝑑𝑥 = −𝜎(𝜑)𝑑𝜑 (11)


� �� � ����
∫� 𝑗𝑑𝑥 = − ∫��� 𝜎(𝜑)𝑑𝜑 (12)
�� ��� ����
𝐼� = 𝑗𝑊𝑇 = − ∫��� 𝜎(𝜑)𝑑𝜑 (13)

where L is channel length.


The modeled OECT structure have T=100 nm, W=10 μm, L=200 μm, with initial density of PEDOT+ and PSS-
ion =1018 cm-3 and constant mobility μ=1·1 X 0-5 m2/V·s. The drain current (ID) for Drain-Source potential (VDS)
from 0V to -1V is plotted (Figure 4) for different Gate-Source potential (VGS) from -1V to 1V. From fig 4 we
conclude that:
• Current reaches to saturation level for high applied VDS at different VGS.
• Drain current decrease and highly dependent on applied Gate-Source potential
• On applying –ve VGS, high drain current is obtained.

7 V = -1V
6 V = -0.5V
Drain Current ID( X 10-5)A

V = 0V
5
V = 0.25V
4 V = 0.5V
3 V = 0.75V
V = 1V
2

0
-1 -0.8 -0.6 -0.4 -0.2 0
VDS (V)
Figure 4: Drain current(ID) profiles calculated analytically, ID dependence on applied Drain-Source
potential(VDS) for different Gate-Source potentials (VGS) .
From figures 5 it is clearly seen the difference in I-V profile for different thickness at different concentration.
Drain current is a function of channel thickness as conductivity of channel decreases with decreases of
thickness, which causes lowering of drain current.
Thus from figure 2, 3, 4, and 5th we conclude that effective area of gate electrode must be comparative equal to
that of channel/electrolyte interface area, the VG must be in between -0.5 V to -1 V, The thickness of
PEDOT:PSS layer must be greater than 200 nm, as below 200 nm, percolation occurs. The WT/L ration must be
greater than 1.
0

-10

-20

-30
ID (μA)

200 nm
-40 100 nm
50 nm
-50

-60

-70
-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1
VGS (V)

Figure 5: Analytical model for different channel thickness (T), width: 10 μm; and L = 200μm

CONCLUSION
In present work we tried analytical model for biosensors based on Organic Electrochemical Transistor, in order
to optimize dimensions for the fabrication of devices for specific purpose. In the first part we tried to explain the
effect of, change in surface area of gate electrode, and change in concentration of electrolyte on drain current,
and how it modulates the effective voltage. The effect on potential near the electrolyte-polymer interface, and In
second part, how the drain current changes with change in thickness of channel thickness and at what value of
gate potential, OECT can give the optimized results. Thus we can say that analytical modeling is a powerful
tool to characterize a device, and can be used for optimal and efficient way of device creation.
REFERENCES

1. M. Pappa, D. Ohayon, A. Giovannitti, I. P. Maria, A. Savva, l. Uguz, J. Rivnay, I. McCulloch, R. M.


Owens, S. Inal, "Direct metabolite detection with an n-type accumulation mode organic electrochemical
transistor", Sci. Adv. Vol. 4, no. 6, eaat0911 (2018)
2. D. T. Simon, E. O. Gabrielsson, K. Tybrandt, M. Berggren, Organic bioelectronics: Bridging the
signaling gap between biology and technology. Chem. Rev. 116, 13009–13041 (2016)
3. V. F. Curto, B. Marchiori, A. Hama, A. M. Pappa, M. P. Ferro, M. Braendlein, J. Rivnay, M. Fiocchi, G.
G. Malliaras, M. Ramuz, R. M. Owens, "Organic transistor platform with integrated microfluidics for in-
line multi-parametric in vitro cell monitoring". Microsyst Nanoeng 3, 17028 (2017)
4. Y. Van De Burgt, A. Melianas, S. T. Keene, G. Malliaras, A. Salleo, "Organic electronics for
neuromorphic computing". Nature Electronics, 1(7), 386-397, (2018).
5. P. Lin, F. Yan, J. Yu, H. L. W. Chan, M. Yang, "The Application of Organic Electrochemical
Transistors in Cell-based Biosensors", Adv.Mater. 22, 3655−3660, (2010)
6. Bubnova, O.; Berggren, M.; Crispin, X. Tuning the Thermoeletric Properties of Conducting Polymers in
an Electrochemical Transistor. J. Am. Chem. Soc., 134, 16456−16459 (2012).
7. M. M. Alam, J. Wang, Y. Guo, S. P. Lee, H. R. Tseng, "Electrolyte-Gated Transistors Based on
Conducting Polymer Nanowire Junction Arrays", J. Phys. Chem. B, 109, 12777−12784 (2005).
8. G. Tarabella, G. Nanda, M. Villani, N. Coppedé, R. Mosca, G. G. Malliaras, C. Santato, S. Iannotta, F.
Cicoira, "Organic Electrochemical Transistors Monitoring Micelle Formation", Chem. Sci. 3,
3432−3435 (2012).
9. S. Chao, M. S. Wrighton, "Solid-State Microelectrochemistry: Electrical Characteristics of a Solid-State
MicroelectrochemicalTransistor Based on Poly(3-methylthiophene)", J. Am. Chem. Soc., 109,
2197−2199, (1987).
10. M. Berggren, R. Forchheimer, J. Bobacka, P. O. Svensson, D. Nilsson, O. Larsson, A. Ivaska,
"PEDOT:PSS-based Electrochemical Transistors for Ion-to-Electron Transduction and Sensor Signal
Amplification. In Organic Semiconductors in Sensor Applications", Springer: Berlin/ Heidelberg,
Chapter 9, pp 263−280 (2008).
11. F. Cicoira, M. Sessolo, O. Yaghmazadeh, J. A. DeFranco, S. Y. Yang, G. G. Malliaras, "Influence of
Device Geometry on SensorCharacteristics of Planar Organic Electrochemical Transistors". Adv.Mater.
22, 1012−1016 (2010).
12. G. Tarabella, C. Santato, S. Y. Yang, S. Iannotta, G. G. Malliaras, F. Cicoira, "Effect of the Gate
Electrode on the response of Organic Electrochemical Transistors", Appl. Phys. Lett. , 97,123304(2010).
13. X. Strakosas, M. Bongo, R. M. Owens, “The organic electrochemical transistor for biological
applications”, J. Appl. Polym. Sci., 132, 41735(2015).
14. J. Rivnay, P. Leleux, M. Sessolo, D. Khodagholy, T. Hervé, M. Fiocchi, G. G. Malliaras, “Organic
electrochemical transistors with maximum transconductance at zero gate bias ”Adv. Mater., 25,
7010(2013).
15. M. Nikolou, G. G. Malliaras, “Applications of poly(3,4-ethylenedioxythiophene) doped with
poly(styrene sulfonic acid) transistors in chemical and biological sensors ”, Chem. Rec., 8, 13(2008).
16. J. Rivnay, S. Inal, A. Salleo, R. M. Owens, M. Berggren, G. G. Malliaras, Nat. Rev. Mater., 3,
17086(2018)
17. Bard, Allen; Faulkner, Larry Electrochemical Methods. Fundamentals and Applications (2nd ed.).
Hoboken, NJ: John Wiley & Sons(2001).
18. D. A. Bernards, D. J. Macaya, M. Nikolou, J. A. DeFranco, S. Takamatsu, G. G. Malliaras, “Enzymatic
sensing with organic electrochemical transistors”, J Mater Chem, 18, 116– 120(2008).
19. D. A. Bernards, G. G. Malliaras, “Steady‐State and Transient Behavior of Organic Electrochemical
Transistors”, Adv Funct Mater , 17, 3538– 3544(2007).
20. M. J. Panzer, C.D. Frisbie, “Exploiting Ionic Coupling in Electronic Devices: Electrolyte‐Gated Organic
Field‐Effect Transistors”. Advanced Materials, 20(16): p. 3177-3180(2008).
21. I. Loi, I. Manunza, A. Bonfiglio, “Flexible, organic, ion-sensitive field-effect transistor”. Applied
Physics Letters, 86(10): p. 103512 (2005).
22. A. M. Nardes, “On the conductivity of PEDOT: PSS thin films”. Technische Universiteit Eindhoven.
Eindhoven, p. 132 (2007)
23. M. Weis, “Gradual channel approximation models for organic field-effect transistors: The space-charge
field effect”. Journal of Applied Physics, 2012. 111(5): p. 054506.
24. E. Stavrinidou, “A simple model for ion injection and transport in conducting polymers”. Journal of
Applied Physics, 113(24)(2013).
25. A. j. Bard, “ Electrochemical methods. Fundamentals and applications”. John Wiley & Sons. 833(2001).
26. H. j. Butt, K. Graf, M. Kappl, “Physics and chemistry of interfaces”. John Wiley & Sons(2006).

You might also like