Invited Review Paper

Probing and modeling of carrier motion in organic devices by electric-field-induced optical second-harmonic generation

, and

Published 19 September 2014 © 2014 The Japan Society of Applied Physics
, , Citation Mitsumasa Iwamoto et al 2014 Jpn. J. Appl. Phys. 53 100101 DOI 10.7567/JJAP.53.100101

1347-4065/53/10/100101

Abstract

By probing dielectric polarization originating from dipoles and electrons in materials, we can study dynamical carrier behaviors in materials and also in devices. Maxwell displacement current (MDC) measurement allows us to directly probe orientational dipolar motion in monolayers, while electric-field-induced optical second-harmonic generation (EFISHG) measurement allows dynamical electron and hole transport in solids to be probed directly. By probing nonlinear polarization induced in solids by coupling with incident electromagnetic waves of laser beam and dc electric field that originate from moving carriers, long-range carrier motion is visualized. Experiments using a time-resolved EFISHG technique reveal carrier transfer in organic devices such as organic field-effect transistors, organic light-emitting diodes, organic memory devices, and organic solar cells, and thus enable us to model the carrier transport mechanism. We anticipate that this novel technique using EFISHG can be a powerful tool for studying carrier behaviors in organic devices as well as in organic materials.

Export citation and abstract BibTeX RIS

1. Introduction

Probing and modeling of carrier transport in materials is a fundamental research subject in electronics and materials science.13) According to Maxwell's electromagnetic field theory,4) the total current flowing across inorganic and organic solids is the sum of the conduction current and Maxwell displacement current (MDC). The conduction current flows when electrons and holes are conveyed under the external electric field formed by applying potentials to electrodes, and it flows steadily. Accordingly, we measure current–voltage (IV) characteristics to study the carrier transport mechanism. In contrast, the MDC is generated when the electric flux density originating from electrons, holes, and dipoles changes with time, and it is basically a transient current. Therefore, we can study dynamical carrier motion and orientational change of dipoles in materials,5,6) if the change in electric flux density can be probed. Therefore, we pay attention to transient currents that are generated by applying external stimuli, such as pulse voltages, light illumination, and heating. For example, to probe the dynamical motion of dipoles in monolayers on the water surface, MDC measurement is employed with monolayer compression,79) alternating UV and visible light irradiation,6) and heating10) among others, where the change in the amount of charge induced on a suspended electrode above the monolayer on the water surface or on the solid substrate is recorded. Similarly, the time-of-flight (TOF) measurement2,5) is often used to probe carrier transport in solids sandwiched between two electrodes, to determine carrier mobility. What we measure using this TOF measurement is the change in the amount of charge induced on the electrode connected to an ammeter. Hence, the principle of TOF measurement is basically the same as that of MDC measurement used for studying monolayers on the water surface or solid substrate, in that the change in the amount of charge induced on the electrode connected to an ammeter is recorded. In TOF measurement, a long-range carrier motion in solids is recorded as a transient current, from which we can determine the transit time of carriers and thus estimate carrier mobility. However, the location of carriers traveling across organic materials cannot be determined directly from this observed transient current. TOF measurement requires an elaborate mathematical approach for analyzing traveling carriers, but this analysis leads to a puzzling situation in which there are many possible solutions to reproduce observed transient currents, particularly in the case of carriers being transported in the presence of multiple carrier traps. This means that it is quite a task to trace actual carrier behavior in solids by the TOF measurement. This situation is more complex in the case of e-TOF measurement using organic field-effect transistors (OFETs) with a three-electrode system.11,12) Electric flux density arising from traveling carriers falls on the three electrodes and induces charges on them. Thus, the charge induced on the electrode connected to an ammeter strongly depends on the location of carriers traveling along the channel and on the geometrical arrangement of electrodes. For example, the charge induced on the drain electrode is almost zero when carriers are located in the channel near the source electrode of OFETs. As such, it is impossible to observe the carrier behaviors over the entire region of the OFET channel, although we can determine the transit time by e-TOF measurement.11,13) The way to overcome this problem is to develop a technique that can directly probe traveling carriers. Direct probing of the carrier motion is indeed helpful in discussing the detailed carrier transport mechanism in organic devices as well as in organic materials.

Gauss's law, in Maxwell's electromagnetic field theory,14,15) suggests an insightful idea for probing the carrier motion in organic materials. That is, by probing the electric field diverging from moving electrons and holes, not by probing charges induced on electrodes, the carrier motion could be visualized if the polarization induced in an organic material surrounding the moving electrons and holes can be probed. That is, we can probe carrier motion by probing dielectric polarization induced by the electric field arising from moving carriers.16) Time-resolved microscopic optical second-harmonic generation (TRM-SHG) is a technique that enables us to probe the propagation of the dielectric polarization induced by electrons and holes that are moving across the organic materials,17) where the nonlinear polarization induced in solids by coupling with incident electromagnetic waves, e.g., laser beam, and dc electric field generated from moving carriers is probed in real time. Furthermore, using a charge-coupled device (CCD) imaging system, we can directly visualize their motion.18) In this paper, we review the technique of probing carrier motion in solids using this SHG, called electric-field-induced optical second-harmonic generation (EFISHG), and show our findings obtained by observation using EFISHG.

2. Methods

As mentioned in Sect. 1, to find a way to directly probe carrier motion in solids, it is very instructive to revisit Faraday's idea14,15) formulated as the Gauss law, where the basic concept of the presence of an electric field arising from charges is essential. Under the non-zero dc field E(0), the nonlinear polarization P is freshly induced when laser light is irradiated on organic materials, owing to the interaction between electromagnetic fields and electrons in the organic materials. This is a basic principle of EFISHG measurement. As the electron distribution of molecules is distorted by the presence of the dc field E(0), this type of nonlinear polarization P is induced even in a centrosymmetric molecular system, such as pentacene and phthalocyanine.19) That is, the induced nonlinear polarization is given by1921)

Equation (1)

where $\chi ^{(3)}(2\omega ;0,\omega ,\omega )$ represents the nonlinear optical susceptibility, ω stands for the angular wave frequency of incident electromagnetic wave, Ei(ω) and Ej(ω) represent the electric fields of incident light, and E(0) is the static electric field arising from moving carriers en, which satisfies the relation

Equation (2)

where εs and ε0 are the relative dielectric constant of organic materials and the dielectric constant of vacuum, respectively.

The induced polarization P(2ω) is a source of second-harmonic (SH) signals and results in the increase in SH signal I(2ω) in proportion to the square of P(2ω), as shown below:

Equation (3)

Here, I(2ω) is SH intensity. From Eqs. (2) and (3), it is clear that SH intensity changes along with the transport of carriers. Using a time-resolved EFISHG technique, we can directly probe carrier transport in organic materials by monitoring the propagation of polarization induced in solids along with carrier motion.18,22)

In electronic devices as well as in organic materials sandwiched between electrodes, there are many sources of dc electric fields. Among them are applied external voltages Eext(0), moving carriers Em(0), trapped charges Et(0), and work function difference between electrodes. Considering these, E(0) in Eq. (3) is replaced with Eq. (4) as follows:

Equation (4)

Equation (3) with Eq. (4) suggests that we can probe electric fields, such as Eext(0), Em(0), and Et(0) by EFISHG. For this purpose, a time-resolved method is helpful in SHG measurement,2327) as described in Sect. 3. Note that $\chi ^{(3)}(2\omega; 0,\omega ,\omega )$ is a material-dependent parameter, and it specifies the feature of materials. As a result, SHG is activated at a specific wavelength 2ω, depending on the material systems, and this fact allows us to selectively probe SHG from one of the layers in multilayers by choosing the appropriate laser wavelength. Furthermore, as SHG is a two-photon process, the generation of photocarriers can be neglected in the measurement. Consequently, we can measure carrier transfer nondestructively.

Note that there have been many reports on SHGs since the pioneering works by Bloembergen28) and Shen.20) However, those works focused on the molecular structure and chemicals from the viewpoint of spectroscopy for characterizing materials. As a result, resonance two-photon modes originating from nonsymmetric chemical structures are the main topics. The SHG measurement method employed in these works is different from the EFISHG measurement method that we use here. Spectroscopic SHG experiments are carried out focusing on the SHG signal generated from materials in a form below:

Equation (5)

where the main key parameter is the material-dependent $\chi ^{(2)}(2\omega ;\omega ,\omega )$ parameter, which depends on the arrangement and nonlinear susceptibility of molecules. This SHG is independent of the presence of the dc field E(0), but is strongly dependent on the $\chi ^{(2)}(2\omega ;\omega ,\omega )$ parameter. This situation is quite different from SHG described in Eq. (3), called EFISHG, where the dc field E(0) has a significant contribution.29) Probing E(0) in Eq. (3) by EFISHG measurement is indeed important for analyzing carrier behaviors and related phenomena in organic devices. Actually, all of the electric phenomena are related to carrier behaviors and thus with electric fields, which are generated from them. EFISHG provides a way for analyzing carrier behaviors in devices, such as OFETs, organic solar cells (OSCs), organic light-emitting diodes (OLEDs), and organic memory devices, and also for analyzing the electric field distribution in organic materials.

3. Experimental system

Figure 1(a) shows the basic arrangement of the experimental system used for the measurement of carrier transfer along the OFET channel, where a laser beam is vertically incident onto the top-contact OFET through the objective lens, and enhanced SHG images are captured using a CCD camera. The light source is an optical parametric oscillator, pumped by a third-harmonic light of a Q-switched Nd-YAG laser (Continuum SureliteII-10). To observe the SHG signal from the OFET, a fundamental wavelength is chosen, depending on the active layer of OFETs. For example, it was chosen to be 1120 nm for studying pentacene-OFETs. SH light generated from the OFET was filtered using a fundamental-cut filter and an interference filter to remove fundamental and other unnecessary light. Figure 1(b) illustrates an example of the timing chart of the laser pulse and voltage applied to the OFET, which is used for the time-resolved EFISHG measurement. The parameter τ stands for the delay time, and by choosing the delay time τ, we can visualize a snapshot of carrier motions at time t = τ after applying a step voltage at t = 0 and determine the electric field distribution in OFETs using the obtained carrier images. At t = 0, immediately after applying the step voltage, carrier injection starts but still without carrier injection in OFETs, while at t = τ, there are injected carriers along the channel of OFETs. That is, we can directly visualize the electric field distribution formed by external voltages in organic devices by choosing τ = 0, whereas we can map the injected carrier distribution with time. Furthermore, by choosing the waveform of applied voltages, we can study electric phenomena induced in organic devices.

Fig. 1.
Standard image High-resolution image
Fig. 1.

Fig. 1. (a) Optical setup of the experimental system used for the measurement of carrier transfer along the OFET channel and (b) example of timing chart of the laser pulse and voltage applied to the OFET.

Standard image High-resolution image

Note that EFISHG experiments are carried out in the same way as those for sandwich-type devices, using a laser beam that is obliquely incident on the devices.30) Furthermore, we have recently developed a novel method of detecting the electric field along the direction of the incident laser beam by introducing a radial polarized beam into the EFISHG measurement system,31) and this system makes it easy for us to study sandwich-type devices.

4. EFISHG measurement and modeling

4.1. Carrier motion in planar-type devices

4.1.1. Stationary state

As described in Sect. 2, we can measure the electric field distribution in organic devices and carrier motion in organic devices among others by probing the nonlinear polarization that generates EFISHG. Firstly, we discuss some examples and then show how carrier behaviors are modeled on the basis of experimental results.

Since the discovery of conducting organic materials, organic devices have attracted much attention in electronics. Among them are OLEDs, OFETs, and organic memory devices. In these electronic devices, organic semiconductor materials are used as active layers to realize their own device function, which are designed on the basis of general semiconductor device physics.32) However, the carrier injection and transport mechanisms in actual organic devices are not so simple, owing to the dielectric nature of active organic layers. Basically, the intrinsic carrier density is quite low in active organic layers. As a result, the function of devices with such layers is rather different from that of general semiconductor devices, and most organic devices are governed by the carrier injection process. We can see such examples of the device function of OFETs and OLEDs among others. Consequently, it is a very effective way to directly measure the electric field distribution in these devices for analyzing and modeling carrier behaviors on the basis of the dielectric physics approach.33,34) As already described in Sect. 2, EFISHG measurement can be used to detect electric fields formed in organic devices. Keeping in mind this idea, the electric field distribution formed along the channel of pentacene-OFETs has been measured. The electric field distribution formed in the pentacene-OFETs before and after hole injection into the channel has been well probed,35,36) where the electric field distribution before hole injection is in good agreement with the distribution analyzed by the rigorous solution of the Laplace equation. That is, the experimental electric field distribution well agrees with that determined theoretically by assuming the potentials of a three-electrode system of OFETs. Furthermore, the electric field distribution that originated from carriers has been determined from EFISHG measurements, and it has been shown that the charge distribution along the OFET channel is basically determined on the basis of the Maxwell–Wagner (MW) model, which describes carrier accumulation at the interface between two different materials. For OFETs, this model accounts for carrier accumulation and distribution on the gate-insulator/active organic layer interface.34) Figure 2 shows the EFISHG measurement of top-contact pentacene-OFETs with the Au source and drain electrodes.36) Under the condition of no carrier injection, EFISHG is shown to be activated to satisfy the rigorous Laplace solution of a three-electrode system of the OFETs,37) as shown in Fig. 2(a). On the other hand, under the condition of hole injection, the intensity of EFISHG decreases owing to the space charge field formed along the channel by injected holes, as shown in Fig. 2(b). Because we know the relationship between EFISHG intensity and electric field [see Eqs. (2) and (3)], it is easy to determine the carrier distribution along the channel, as plotted in Fig. 3.36) These results show that carrier accumulation along the channel is governed by the MW effect, and the distribution is basically determined on the basis of the MW model.34,35) Our further EFISHG study shows that the carrier distribution along the channel in the steady state follows a solution of the classical differential equation of current, which describes continuity of carrier transport along the channel, and which is modeled on the basis of a transmission line model, as an extension of the MW model.3840) Analytical results show that the potential distribution along the channel is not linear, but it changes on satisfying the square root dependence along the channel, as shown below:

Equation (6)

where x = 0 and L represent the edge of the source and the drain electrodes, respectively. Accordingly, the electric field distribution is given as

Equation (7)

EFISHG enables the probing of this electric field distribution and generates the signals that follow Eq. (3) with the relation Eq. (7), where E(0) in Eq. (3) is replaced with E(x) in Eq. (7). Figure 4(a) shows the plots of calculation (solid line) and experimental results (symbols) of the normalized electric field of the conduction carriers in the channel region of OFET in the on-state.37) The calculation is based on the rigorous solution of the differential equation, but a small difference is identified in experimental electric intensity and the calculated result of the electric field of the charged layer (indicated by solid lines), possibly owing to the presence of trapped carriers that contribute to the additional electric field [see Eq. (4)]. Figure 4(b) shows the electric field distribution obtained from SHG experiment plotted in the log–log scale, and the line represents the linear fit, suggesting the square root dependence of electric field along the channel [see Eq. (7)].

Fig. 2.

Fig. 2. EFISHG image obtained from top-contact pentacene-OFETs with application of (a) positive voltage to the gate electrode and (b) negative voltage to the gate and drain electrode. Reprinted with permission.36) © 2008 AIP Publishing LLC.

Standard image High-resolution image
Fig. 3.

Fig. 3. Example of the carrier distribution along the channel obtained from the EFISHG measurements. Reprinted with permission.36) © 2008 AIP Publishing LLC.

Standard image High-resolution image
Fig. 4.
Standard image High-resolution image
Fig. 4.

Fig. 4. Calculation (solid line) and experimental results (symbols) of the normalized electric field of the injected carriers in the channel region of OFET in the on-state and (b) the electric field distribution obtained from SHG experiment plotted in log–log scale. Reprinted with permission.38) © 2009 AIP Publishing LLC.

Standard image High-resolution image

As has been described above, it is clear that EFISHG measurement can be used for probing the steady-state carrier distribution along the channel of OFETs. Some organic materials show ambipolar behaviors,32) which means that electrons and holes can enter into materials from electrodes. In the field of electrical insulation engineering, this carrier behavior is called double injection.3) Consequently, probing injected electrons and holes and determining their distribution in organic devices are also important.35,37) For example, for organic light-emitting device applications, simultaneous electron and hole injection is a key to device operation. Using ambipolar organic light-emitting transistors (OLETs) with an active layer of the green-light-emitting polymer poly(9,9-di-n-octylfluorene-alt-benzothiadiazole) (F8BT), which has a photoluminescent peak at the wavelength of 560 nm, EFISHG measurement has been employed using a laser beam of 840 nm wavelength.4143) For the experiment, we need to remove the effect of photoluminescence using a band pass filter so that we can probe EFISHG at a wavelength of only 420 nm.

Figure 5(a) shows the structure of OLETs, where electron and hole transfer into the active layer from opposite electrodes is essential to emitting light. Figure 5(b) shows an image of electroluminescence (EL) from the F8BT OLETs, which is generated from the channel region and it is visualized as a bright band using a CCD camera.42) This emission band represents the meeting point of electrons and holes followed by recombination for EL, but this image does not directly show the distribution of holes and electrons along the channel. An EFISHG image shown in Fig. 5(c) clarifies this situation, because EFISHG is proportional to the electric field, as given by Eq. (3). Therefore, we can map the carrier distribution along the channel experimentally using the Gauss law [Eq. (2)]. Figure 6 shows the potential profile reconstructed from the EFISHG experiment.42) Closed circles represent the results, and we can see that there is a zero-potential position in the channel. It is easy to understand that electrons and holes accumulate, depending on the polarity of potentials. Interestingly, this distribution is in good agreement with the theoretical carrier distribution that is derived by solving an one-dimensional continuity equation for the current density J, in the same way as in the derivation of Eqs. (6) and (7). In this derivation, it is assumed that electrons and holes dominate carrier conduction, depending on the zero-potential position. That is, the idea of zero-potential position clearly developed from the study of ambipolar OFETs by EFISHG measurement.

Fig. 5.

Fig. 5. (a) Schematic image of OLETs and (b) EL image from F8BT OLETs, and (c) distribution of holes and electrons along the channel. Reprinted with permission.41) © 2014 Elsevier.

Standard image High-resolution image
Fig. 6.

Fig. 6. Potential profile in the OLET reconstructed from the EFISHG images. Reprinted with permission.42) © 2012 AIP Publishing LLC.

Standard image High-resolution image

4.1.2. Transient carrier behavior

As mentioned in Sect. 2, it is possible to probe carrier motion by TRM-SHG measurement. Using top-contact pentacene-OFETs with the Au source and drain electrodes, TRM-SHG measurement was carried out. Here, the channel length of the OFET is 40 µm. Figure 7(a) shows an image of TRM-SHG from the channel of pentacene-OFET with a 500-nm-thick SiO2 gate insulator, where the SHG image was obtained including carrier motions after a positive voltage pulse was applied. That is, the positive voltage pulse Vpulse = 70 V was applied to the source electrode with respect to the gate and drain electrodes that were shorted and grounded, i.e., Vds = Vgs = −70 V. At τ = 0 ns, the laser pulse coincides with the rising edge of the voltage pulse, and SHG signals are found near the edge of the source electrode, indicating that the Laplace electric field E0 parallel to the channel is formed only around the source electrode. Interestingly, as it is clearly shown in the image, the emission band of the SHG signal gradually moves in the channel from the source electrode to the drain electrode with time. The visualized emission band motion starting from the source electrode is the direct evidence of hole injection from the Au source electrode. That is, pentacene-OFET shows a p-type behavior, where the majority of carriers are holes injected from the source electrode and the carrier mobility is in the range of 0.1–0.2 cm2 V−1 s−1. Note that the SHG image is confined in the region within a radius of approximately 150 µm, owing to the spot size of the laser [see the region indicated by a dashed circle in Fig. 7(a)]. To further clarify the hole-transport mechanism across the channel, we plotted the SHG peak position with respect to the elapsed time, as shown in Fig. 7(b).44) From Figs. 7(a) and 7(b), it is clear that holes travel in the direction from the source Au electrode to the drain electrode, satisfying the square root time dependence.45,46) This time dependence reflects the interface charge propagation process [see Fig. 7(b) and Ref. 39], and well supports the RC ladder circuit model that is employed for analyzing transient carrier behaviors along the OFET channel.32,39,40) On the other hand, under the application of negative pulse voltage, the position of the SHG peak never moves; the SHG signals are concentrated around the edge of the source electrode during the measurement, suggesting that electron injection is prohibited.

Fig. 7.
Standard image High-resolution image
Fig. 7.

Fig. 7. (a) TRM-SHG image from the channel of pentacene-OFET at various delay times with application of positive voltage to the source electrode and (b) the SHG peak position with respect to the elapsed time.

Standard image High-resolution image

It is instructive here to note that carrier propagation is dependent on the interfacial condition of the channel formed on the surface of the gate insulator, e.g., the presence of carrier traps. Equation (4) suggests that TRM-SHG measurement can be used for monitoring the contribution of the surface states of the channel. To confirm the potentiality of TRM-SHG measurement, an experiment was carried out using pentacene-OFETs with a 100-nm-thick poly(methyl methacrylate) (PMMA) layer on the surface of the gate SiO2 insulator. Figures 8(a) and 8(b) show the results of the experiment carried out under the same experimental conditions as those for results shown in Fig. 7(a), where SHG is concentrated in the front region of the propagating SHG profile.47) This result indicates that the space-charge field Es formed along the channel is due to trapped holes and moving holes. Analysis based on a carrier trap model well accounts for the results. That is, TRM-SHG measurement is effective for probing interface states of the channel region. Furthermore, note that we can probe the effect of nanoparticles and dipoles on the carrier injection and transport by using the TRM-SHG measurements.48,49)

Fig. 8.

Fig. 8. SHG intensity distribution of pentacene-OFET with (a) SiO2 and (b) PMMA/SiO2 insulator. Reprinted with permission.47) © 2010 AIP Publishing LLC.

Standard image High-resolution image

As shown in Fig. 5, electrons and holes have a significant contribution in ambipolar OLETs. To understand the details of the carrier mechanism of EL, it is very informative to simultaneously probe the dynamical electron and hole carrier behaviors that lead to EL, after the electron and hole injection starts at the source and drain electrodes, respectively. Also, probing hole and electron motion simultaneously would be useful for understanding the electrical breakdown of materials50) and the recombination of injected carriers in organic devices.30) TRM-SHG measurement has the potentiality for visualizing electron and hole carrier motion simultaneously. Our group has already visualized such carrier motion in OFETs, by applying positive and negative pulse voltages to the source and drain electrodes with respect to the gate electrode. Also, we should note that the activation of SHG is material-dependent, owing to the material-dependent parameter $\chi ^{(3)}(2\omega ;0,\omega ,\omega )$ with the angular frequency ω in Eq. (3). Consequently, probing carrier motion in one of the layers in multilayer systems is possible. Actually, we demonstrated the detection of such carrier motions in OFETs with an active layer in a double-layer system, e.g., OFETs with a C60/pentacene active layer.5155) Furthermore, visualization of two-dimensional planar carrier motion in EFISHG measurement allows us to probe anisotropic carrier motions in organic layers, and this observation provides a new way for characterizing anisotropic carrier transport in organic thin films by employing a round-shaped electrode in the EFISHG measurement.56)

4.2. Carrier motion in sandwich-type devices

In Sect. 4.1, we focused on the probing of carrier motion in planar-type OFETs by EFISHG measurement. On the other hand, as mentioned in Sect. 2, this measurement is possible for measuring electric fields in the film-thickness direction as well as for probing carrier motions in sandwich-type devices, by introducing incident laser light obliquely onto such devices. In this section, we discuss topics on sandwich-type devices.

Sandwich-type devices such as EL and OPV are designed on the basis of general semiconductor device physics.57) However, owing to the dielectric nature of organic semiconductor layers, the complexity of interfaces, and the presence of traps among others, it is quite a task to describe the energy diagram of actual organic devices. For example, organic double-layer diodes composed of p- and n-type organic semiconductors show a rectification property, but the actual potential distribution formed in these diodes is far from the general semiconductor p–n diodes, owing to the dielectric nature of the organic semiconductor layers. Consequently, it is very effective to directly measure electric field and potential distributions in these devices for analyzing carrier behaviors in actual organic devices. As already discussed, EFISHG measurement enables the direct probing of the electric field distribution in organic layers and also enables the study of carrier behaviors in organic sandwich-type devices.5860) Results of EFISHG measurement well account for the rectifying property of organic double-layer diodes, e.g., diodes with a structure of ITO/PI/TIPS–pentacene/Au.6171) Here, we summarize some examples to show how carrier behaviors in sandwich-type devices are analyzed by EFISHG measurement.

4.2.1. Double-layer EL diodes

OLEDs have attracted much attention in electronics. Basically, an OLED is an injection-type device with two different electrodes that are separated by an organic multilayer system. Using the energy diagram of molecules, e.g., HOMO and LUMO states, many research studies have been carried out to improve device performance. Ideas of using low-work-function electrodes, depositing dipolar layers on the surface of electrodes, and utilizing multilayer systems comprising hole and electron transport, and electroluminescent layers among others have been presented.30,7274) These important ideas for achieving high-performance OLED devices are well accepted. However, owing to the ambiguities of energetics at the organic–metal and organic–organic interfaces, electron and hole behaviors in OLED devices are still not fully understood. To clarify the details, we need to pay attention to carrier injection from electrodes, carrier transport across active layers, and charge accumulation and recombination at the interface, which lead to EL. Hence, one of the most effective ways is to probe directly carrier motion in OLED devices. In particular, probing carrier behaviors simultaneously with charge accumulation at the interface is very important. EFISHG measurement can be applied for this purpose. It is instructive here to note that the MW effect well accounts for the charge accumulation at the interface between two adjacent materials;31,34) thus, the analysis of organic EL devices as a MW effect element in combination with the EFISHG measurement is helpful.75)

The EFISHG experiments on the indium zinc oxide (IZO)/N,N'-di-[(1-naphthyl)-N,N'-diphenyl]-(1,10-biphenyl)-4,4'-diamine (α-NPD)/tris(8-hydroxy-quinolinato) aluminum(III) (Alq3)/LiF/Al structure OLED device showed good charging and discharging of carriers on electrodes and carrier transit across α-NPD and Alq3 layers, and nonreversal charging and discharging processes resulted from the different carrier behaviors accompanied by EL. Briefly, a rectifying behavior was observed for the devices together with the enhancement of ordinary green EL from the Alq3 layer under forward bias condition (IZO positive). Figures 9(a)–9(d) respectively show the transient EFISHG intensity I(2ω) and EL intensity under application of a square wave voltage of Vex = 5 V.75) The SHG I(2ω) with a wavelength of 410 nm is selectively generated from the α-NPD layer by laser irradiation with a wavelength of 820 nm. Figure 9(b) shows the electric field in α-NPD, E1, plotted using Eq. (3). In Fig. 9, regions 1–3 and 4–6 respectively correspond to charging and discharging processes. In region 1, E1 increases with a response time τRC of 7 × 10−8 s. This value agrees well with the RsC* = 7.7 × 10−8 s with a lead resistance Rs of 170 Ω and a series capacitance $C^{*} = C_{1}C_{2}/(C_{1} + C_{2})$ of 4.5 × 10−10 F (C1: capacitance of α-NPD layer, C2: capacitance of Alq3 layer). E1 at the end of region 1 was estimated from the SHG intensity as 2.7 × 105 V/cm, and corresponded well to the Laplace electric field determined from the ratio of C1 to C2 as $E_{1} = C_{2}/(C_{1} + C_{2})\cdot V_{\text{ex}}/d_{1}$. Here, d1 is the thickness of α-NPD. Over the entire region 1, no EL enhancement was observed. These results show that region 1 reflects electrode charging by Vex. In region 2, SHG I(2ω) and E1 are nearly constant, and no EL enhancement was observed. In region 3, E1 decays with a relaxation time of 2 × 10−3 s, and finally saturates. This evidently shows that positive charge Qs is accumulated at the α-NPD/Alq3 interface. Note that this charge accumulation is due to the MW effect, and it is accompanied by the EL enhancement. At the end of region 3, the electric field is E1 = 2.2 × 105 V/cm, and the accumulated charge density is calculated as Qs = 6.8 × 10−8 C/cm2. The relaxation time corresponds well to the MW relaxation time of a double-layer system τMW = 2 × 10−3 s. Regions 4–6 are discharging processes. In region 4, E1 decreases with the time constant τRC = 7 × 10−8 s, in the same way as in region 1. This result indicates that the charge induced on electrodes at the end of region 1 is discharged through the external circuit. On the other hand, Qs (> 0) remains at the α-NPD/Alq3 interface, which gives rise to electric fields $E_{1} = - Q_{\text{s}}/(C_{1} + C_{2})\cdot 1/d_{1} < 0$ and $E_{2} = + Q_{\text{s}}/(C_{1} + C_{2})\cdot 1/d_{2} > 0$. At the end of region 4, the electric field saturated at E1 = −3.1 × 104 V/cm, suggesting that most of the charge Qs remain at the interface. The enhanced EL intensity is nearly constant over the entire region 4. In region 5, the electric field E1 in the α-NPD layer decreases and finally reaches E1 = 0, while the enhanced EL intensity does not change. In region 6, E1 = 0 but EL gradually decays with relaxation time τ = 3 × 10−4 s. As described above, in actual devices, charging and discharging processes are nonreversible, that is, the relaxation time τMW = 2 × 10−3 s for charging, whereas τMW = 5 × 10−6 s for discharging. The nonreversal behavior is mainly due to the difference in the contribution of the electric field arising from accumulated charges at the interface. In the charging process, accumulated positive charges hinder hole injection from ITO electrodes via the α-NPD layer. On the other hand, in the discharging process, accumulated positive charges assist electron injection from the Al electrode via the Alq3 layer. This model well accounts for two EL modes generated in these sandwich-type EL devices, depending on the frequency of the applied square voltage.7680) As mentioned above, we can monitor dynamical carrier behaviors in charging and discharging processes in EL devices by EFISHG measurement.

Fig. 9.

Fig. 9. (a) Transient EFISHG intensity I(2ω) under application of a square wave voltage Vex, (b) electric field in α-NPD layer, (c) estimated charge accumulated at the interface, and (d) transient EL intensity. Reprinted with permission.75) © 2010 American Chemical Society.

Standard image High-resolution image

Furthermore, note that EL has also received much attention as one of the prebreakdown phenomena of insulators in the field of electrical insulation engineering for the past 10 years, before the pioneering OLED work by Tang and Van Slyke,72) where the main research interest is to detect very weak EL signals for the diagnostics of insulators. EFISHG has an impact on the electrical insulation engineering community as a new method of detecting the prebreakdown phenomena of insulators as well as organic EL devices, because carrier motions leading to pre-electrical breakdown that results in EL can be probed before the EL as pre-breakdown phenomenon is initiated.50) This is also an advantage of using EFISHG measurement.

4.2.2. Double-layer memory devices

Ferroelectric polymers are widely used as memory materials, where the copolymer of vinylidene fluoride (VDF) and trifluoroethylene (TrFE), P(VDF–TrFE), stands out owing to its bistable and remanent polarization that can be repeatedly turned over by an external electric field.81,82) Recently, ferroelectric polymers have also been used in electronic devices such as OFETs83) and organic photovoltaic cells (OPVs),84) where a strong internal electric field induced in the active layer of these devices by the spontaneous polarization of ferroelectric polymers is well utilized for efficient device operations. As a result, carrier behaviors in the active semiconductor layer, e.g., pentacene, have been extensively studied in terms of the turn over of the spontaneous polarization of the ferroelectric layer.8588) For the p-type pentacene-OFETs with a ferroelectric P(VDF–TrFE) gate insulator, hole/electron injection, transportation, charging, and discharging at the P(VDF–TrFE) gate insulator interface and so forth have been clarified in response to the turn over of the spontaneous polarization of the ferroelectric layer. Similarly, organic memory MIS diodes with a ferroelectric P(VDF–TrFE) insulator have been studied. These studies are mainly carried out by electrical measurements such as displacement current measurement (DCM), but these are no longer sufficient to account for carrier behaviors in a variety of organic semiconductors being utilized with ferroelectric P(VDF–TrFE) layers. With EFISHG, the electric field change induced in organic devices using a ferroelectric layer gives much information on the carrier behavior in these devices.

Figure 10 shows a typical example of DCM measurements for MFM and MFSM [inset of Fig. 10(b)] devices, which were prepared on patterned IZO substrates.85) A P(VDF–TrFE) (72–28 mol %) layer of 200 nm thickness was spin-coated on the substrate, and then a pentacene layer (200 nm, for MFSM device) and gold electrode were thermally evaporated successively. For the DCM measurement, a ramp voltage was applied to the IZO electrode. Figure 10(a) shows the results of DCM for the MFM device. Two peaks (peaks 1 and 2) are clearly observed at symmetric positions at voltages corresponding to the coercive electric field Ec of 0.6 MV/cm, and show no dependence on sweeping frequencies of ramp voltage. These results suggest that the appearances of peaks 1 and 2 are due to the turn over of spontaneous polarization.

Fig. 10.
Standard image High-resolution image
Fig. 10.

Fig. 10. DCM for (a) MFM and (b) MFSM devices. Reprinted with permission.85) © 2011 AIP Publishing LLC.

Standard image High-resolution image

On the other hand, for MFSM devices [Fig. 10(b)], three peaks A, B, and C are clearly observed. The appearance of the third peak, peak C, at the non-symmetric position of peak A is a typical characteristic of the MFSM device. Results suggest a two-step polarization reversal of the ferroelectric layer in the pentacene/P(VDF–TrFE) double-layer device. However, this is a mere speculation. For studying carrier behaviors with the turn over of polarization, it is very helpful to measure the electric field formed in pentacene and PVDF layers. Figure 11 shows the result of EFISHG measurement, where the electric field across the pentacene layer is recorded.86) Results clearly verify the proposed two-step polarization reversal process by the two different hysteresis loops: one loop is shown in Fig. 11 corresponding to peaks A and B, and the other loop corresponds to peak C. Results of EFISHG suggest that the interaction of interfacial charge and ferroelectric polarization is responsible for this two-step process. The two-step turn over strongly depends on organic semiconductor materials, where the interaction at the interface between the ferroelectric layer and the semiconductor layer is different.8791) EFISHG measurement probes well such differences and is available for the development of memory devices.

Fig. 11.

Fig. 11. Transient changes in EFISHG intensity with application of a ramp voltage. Reprinted with permission.86) © 2012 The Japan Society of Applied Physics.

Standard image High-resolution image

4.2.3. Organic solar cells

There are many types of OSC, e.g., bilayer type and blended type. Among them, bulk heterojunction (BHJ) OSCs have been paid much attention in terms of their conversion efficiency. However, the basic mechanism of carrier behavior is so complex, and the development of techniques available for probing carrier behaviors in OSCs is anticipated, for obtaining a complete picture of carrier behaviors. In BHJ OSCs, electron donor and acceptor molecules are mixed to introduce a larger contacting area for creating excitons to be separated into electrons and holes efficiently. However, it is very speculative to identify carrier paths in such devices. The EFISHG measurement is very useful for investigating fundamental processes, not only in bilayer (double-active-layer) OSCs but also in blended-type OSCs.55,92100) The main reason is that the generation of the EFISHG depends on the material properties of the targeted sample. Consequently, it is possible to observe and study different material components in the BHJ layer individually by choosing two appropriate laser wavelengths. This is an advantage of EFISHG for the study of the carrier path in BHJ OSCs. For example, for OSCs codeposited with pentacene and C60, EFISHG measurement has been carried out at two laser wavelengths of 1000 and 860 nm, by which carrier behaviors inside the pentacene and C60 have been selectively probed. As a result, the two carrier paths for electrons and holes have been identified from the EFISHG response by the application of a laser pulse to the OSCs. That is, two types of electric field pointing in opposite directions are identified as a result of the selective probing of SHG activation from C60 and pentacene. Also, under open-circuit conditions, the transient process of the establishment of the open-circuit voltage inside the co-deposited layer has been directly probed in terms of the photovoltaic effect. Figure 12 shows an example of the EFISHG response. As we can see in the figure, we can identify the difference in the generation of EFISHG, suggesting two carrier paths in OSCs.92) The EFISHG provides an additional promising method for studying the carrier path of electrons and holes as well as the dissociation of excitons in BHJ OSCs.

Fig. 12.
Standard image High-resolution image
Fig. 12.

Fig. 12. (a) Schematic image of the suggested morphology and background electric field in pentacene and C60 components and (b) EFISHG responses from BHJs OSCs under open-circuit condition using illumination pulse. Reprinted with permission.55) © 2014 AIP Publishing LLC.

Standard image High-resolution image

As mentioned above, EFISHG measurement can be applied to studying carrier behaviors in organic devices, including OFETs, OSCs, and OLEDs. Obviously, this method has potential applications in the study of carrier behaviors in inorganic devices.

5. Conclusions

In this review paper, we described a novel method for studying carrier motions and electric fields in devices by optical second-harmonic generation, where nonlinear polarization induced in the presence of the dc electric field is well probed by coupling with a laser beam. As carriers are the source of electric field, we anticipate that this method can be broadly applied to studying not only organic but also inorganic devices. As a future scope, we also need to study carrier motions in energy space (k-space). If we can obtain information in actual space and also in energy space, it will be more effective. For this purpose, experimental methods coupled with EFISHG and other techniques for probing energetics would be powerful.52,101) Charge modulation spectroscopy, electron absorption spectroscopy, and others have potentiality, and experiments are underway.

Acknowledgement

A part of this work was financially supported by a Grant-in-Aid for Scientific Research (Nos. 22226007, 21686029, 24360118, 25709022) from the Japan Society for the Promotion of Science.

Please wait… references are loading.

Biographies

Mitsumasa Iwamoto

Mitsumasa Iwamoto graduated from Tokyo Institute of Technology and earned his B.E. degree in 1975. He earned his M.E. and D.E. degrees from Tokyo Institute of Technology in 1977 and 1981, respectively. He is a professor of Tokyo Institute of Technology, a fellow of JSAP and IEICE, and a senior member of IEEJ, among others. His current interests focus on organic electronics, mainly on the electrical, optical, physical, and dielectric properties of thin films, monolayers, liquid crystals, and molecules.

Takaaki Manaka

Takaaki Manaka graduated from Tokyo Institute of Technology and earned his B.E. degree in 1995. He earned his M.E. and D.E. degrees from Tokyo Institute of Technology in 1997 and 2000, respectively. Presently, he is an associate professor of Tokyo Institute of Technology, a member of JSAP and the IEEJ. His current interests focus on the optical properties of organic materials, organic and molecular electronics, and liquid crystals.

Dai Taguchi

Dai Taguchi graduated from Tokyo Institute of Technology and earned his M.E. and D.E. degrees from Tokyo Institute of Technology in 2005 and 2008, respectively. Presently, he is an assistant professor of Tokyo Institute of Technology and a member of JSAP and IEEJ. His current interests focus on organic electronics.

10.7567/JJAP.53.100101