J. Semicond. > 2016, Volume 37 > Issue 6 > 064001

SEMICONDUCTOR DEVICES

Comparison between Pt/TiO2/Pt and Pt/TaOX/TaOY/Pt based bipolar resistive switching devices

Patrick W. C. Ho, Firas Odai Hatem, Haider Abbas F. Almurib and T. Nandha Kumar

+ Author Affiliations

 Corresponding author: Patrick W. C. Ho Email: kecx2pha@nottingham.edu.my

DOI: 10.1088/1674-4926/37/6/064001

PDF

Abstract: Nonvolatile memories have emerged in recent years and have become a leading candidate towards replacing dynamic and static random-access memory devices. In this article, the performances of TiO2 and TaO2 nonvolatile memristive devices were compared and the factors that make TaO2 memristive devices better than TiO2 memristive devices were studied. TaO2 memristive devices have shown better endurance performances (108 times more switching cycles) and faster switching speed (5 times) than TiO2 memristive devices. Electroforming of TaO2 memristive devices requires~4.5 times less energy than TiO2 memristive devices of a similar size. The retention period of TaO2 memristive devices is expected to exceed 10 years with sufficient experimental evidence. In addition to comparing device performances, this article also explains the differences in physical device structure, switching mechanism, and resistance switching performances of TiO2 and TaO2 memristive devices. This article summarizes the reasons that give TaO2 memristive devices the advantage over TiO2 memristive devices, in terms of electroformation, switching speed, and endurance.

Key words: resistive switching devicesTiO2 devicesTaO2 devicesnon-volatile memory devices

In 1971,Leon Chua postulated the theory of a memristor as the fourth basic circuit element[1]. Later in 2007,a thin film stack comprising titanium dioxide TiO2 as the bulk layer material sandwiched between platinum (Pt) electrodes was proposed as a candidate for nonvolatile resistive random-access memory (RRAM) that exhibits resistance switching behavior[2]. In 2008,a nanoscale resistance switching device was successfully fabricated,and named memristive devices[3]. These memristive devices exhibited nonvolatile resistance switching abilities that make them a promising alternative to flash memory[4],dynamic random-access memory (DRAM)[5] and static random-access memory (SRAM)[6] applications. Its growing reputation for being a nanoscale nonvolatile device has attracted researchers into developing RRAM applications[5, 6],improved logic gates[7],designing memristive neuromorphic systems[8],and improving memory density[9].

Although preliminary experimental results show that memristors can replace memory devices in the future,the endurance of memristors is still less than that of a typical DRAM (1012 cycles)[10]. The highest number of endurance cycles achieved by a TiO2 memristive device is 2×106 cycles using a TiN electrode[11],and 104 cycles using Pt electrodes[12],where both are far from the endurance benchmark of current established memory devices[13].

Recently,researchers have shown that other metal oxides could also be used as the bulk material to fabricate nanoscale memristive devices,such as zinc tin oxide[14],vanadium oxide[15],silicon oxide[16],copper oxide[17],hafnium oxide[18],and tantalum oxide (TaO2) [19]. However,TaO2 is considered as one of the prospective resistive switching materials due to its two stable phases (TaO2 and Ta2O5)[20],which can control the stability of its high and low resistance states. In addition,the TaO2 physical memristive devices have exhibited better resistance switching characteristics than TiO2 memristive devices,in terms of switching speed (5 times faster) and better endurance (108 times more switching cycles)[21],longer retention (10 years) and larger resistance ratio. Due to these statistics,a comparison between TaO2 and TiO2 physical memristive devices was conducted to understand the requirements and factors to fabricate better performing memristive devices. TiO2 memristive devices were selected as the benchmark because the study and fabrication of TiO2 memristive devices have been extensive.

Several review articles have provided a general overview of the memristive devices that are found in References [22-24]. This research work differs from other memristor review articles because this article provides a direct comparison between two types of memristive devices that differ only in the substrate. The comparisons performed in this article are in the aspect of physical structures,electroformation,switching mechanism,switching speed,switching energy,resistance retention,and device endurance. The review in this article investigates the physical and chemical properties of TiO2 and TaO2-based memristors,and the changes that occur in both memristive devices during electroformation and switching. The properties of the device structures that could lead to possible future discovery of new bulk layer materials or improved structures were also studied. This article provides a review on the comparison of the performances of TiO2 and TaO2 physical memristive devices in the literature,and discusses the methods which could improve retention,endurance and switching performance of memristive devices. The material and geometric factors that affect switching speed and resistance ratios are also explained. This research work provides several recommendations,summarized from previous work to fabricate better performing memristive devices. This article concludes the physical and material factors that improve the electrical performance of memristive devices.

This article is organized as follows: Section 2 describes the electroformation process in general and also specifically for both structure types of memristive devices; Section 3 explains the switching mechanisms,and compares the switching speeds and energy requirements; Section 4 discusses resistance state non-linearity,resistance state stability and resistance ratio; Section 5 explains the difference in retention capability and endurance of the memristive devices; Section 6 studies the effect of varying electrode cross-sectional area and bulk thickness; and Section 7 summarizes this article. This article is also an extension of the article in Reference [25].

TiO2 and TaO2 memristive devices have different physical structures (Figure 1). The TiO2 memristive device has a bulk layer of TiO2-X sandwiched between two Pt electrodes,resembling metal-insulator-metal (MIM) structures with the Pt electrodes being metal and TiO2 as an insulator[2]. For TaO2 memristive devices,the TaO2-X bulk layer is formed by reactive sputtering of a Tantalum (Ta) layer in oxygen,and a Ta2O5-X layer is fabricated by using oxygen plasma with Ta metal above the TaO2-X bulk layer[26]. The oxygen plasma conditions ensure that the Ta metal is oxygen rich to form the Ta2O5-X layer. The top Pt electrode is directly deposited over the Ta2O5-Xlayer[27],giving the substrate of the TaO2 memristive devices two layers instead. The purpose of the insulating layer of Ta2O5 is to convert the device into a bipolar resistive switching device[28, 29] and also results in TaO2 memristive devices resembling metal-insulator-semiconductor-metal (MISM) structures.

Figure  1.  Typical construction of memristive devices.
(a) Titanium oxide memristive device with TiO2 bulk layer. (b) Tantalum oxide memristive device with TaO2 bulk layer and Ta2O5 insulator layer.

However,by using the Ta2O5/TaO2 RRAM structure instead of MIM structure and by selecting each of these layers to have an appropriate resistance,two important factors can be controlled: the programming current and device stability[26]. The MISM memristor structure reduces the low resistance state current by maintaining a medium resistance during switching between MISM and metal-metal-semiconductor-metal (MMSM),which is the low resistance state of the MISM structure[26]. The low-current operation devices (<100μA) are necessary for more effective RRAM applications[20]. In addition,the Ta2O5 MIM structure is not suitable for bipolar resistive switching due to the highly resistive Ta2O5 forming a very high Schottky barrier with the metallic electrodes. Thus,a reverse bias will not be able to overcome the high Schottky barrier. Therefore,in an MISM structure,the interface between the metallic electrode and semiconductor TaO2 bulk layer is reported to form another Schottky barrier that will decrease the high Schottky barrier,which assists resistance switching,and dominates the relationship between current and voltage for MISM memristor structures[30].

In the TaO2 memristive devices,the TaO2 layer functions as an oxygen vacancies supply layer whereas Ta2O5 is an oxygen vacancies accumulation layer[31]. Although the fabrication of TaO2 memristive devices requires an additional layer of Ta2O5,but it serves as an advantage against other MIM structured memristive devices because the variation of the thickness of the Ta2O5 insulator layer provides an easier manipulation of the resistance ratio of TaO2 memristive devices[26].

Despite having different device structures and bulk layer materials,both TiO2 (MIM) and TaO2 (MISM) memristive devices exhibit similar resistive switching behavior. This is evident by the fact that both devices are able to switch between high and low resistance states and the switching of states depends on the direction of current (demonstrating bipolar switching behavior). However,the switching mechanisms are different due to the different device structures. In short,the difference in electrical performance is mainly due to the physical structure and chemical properties of bulk layer material of TiO2 and TaO2 memristive devices.

*Please note: hereafter,TiO2 and TaO2 memristive devices will respectively be written as TiO2 and TaO2 devices.

Post-fabricated MIM structures cannot conduct electricity because insulator bulk layers have high resistance against electrical conduction under normal operating conditions. To form a device capable of conducting current,electroformation is performed,which is an irreversible one-time process that forms a device capable of conducting current,usually by creating conducting channels through a high resistance layer. This is done by causing physical changes to the device using electrical potential. It is essential for MIM structures to undergo electroformation in order to exhibit memristive behavior[32, 33].

The magnitude of electroforming voltage is usually much larger than the switching voltage magnitude required for subsequent resistance switching to take place[34]. The electroforming voltage magnitude also usually exceeds the breakdown potential of the device,causing a build-up or depletion of oxygen ions at the electrodes,depending on the polarity of electroforming voltage. The subsequent physical switching mechanism takes place around these erupted regions because it was observed that conducting channels are formed and collapse near these regions[35].

During electroformation,the potential bias applied at the electrodes energizes the oxygen ions of negative charge to drift to one of the electrodes of the device. Oxygen ions will combine to form oxygen gas molecules. The oxygen gas molecules are trapped in the interface between the electrode and bulk layer. Gas pressure builds up as more oxygen gas

molecules are formed. This eventually causes an eruption through the top electrode and a permanent physical deformation occurs in the memristive device[12].

Although there is permanent physical change to the device during electroformation,the subsequent resistance switching after electroformation is not affected. This is due to the subsequent conducting channels that are responsible for the conduction of current being formed and collapsed near the electroformed conducting channel[36, 37]. Thus,no further deformations at other regions are required[37].

There are various mechanisms proposed to explain the theory of electroformation in MIM devices. According to Reference [34],the electroforming voltage is given by:

VFΔGMXxF

(1)

where ΔGMX is the Gibbs free energy of formation of the bulk layer material and F is Faraday's constant. The proposed theory by Greene et al. states that the electroforming voltage is independent of bulk layer thickness[34]. This theory can be used to compare two devices of similar physical dimensions but different bulk layer material. According to Equation (1),the electroforming voltage required is proportional to the Gibbs free energy of the formation of the bulk layer material. The Gibbs free energy of the formation of metal oxides commonly used in the bulk layer of MIM resistive switching devices is shown in Table 1[38]. From Table 1,it was noticed that the Gibbs free energy of the formation of TaO2 and Cu2O is among the lowest whereas HfO and TiO2 is among the highest. TaO2 is more widely used over Cu2O due to its more stable resistance states as seen in the current-cycle and resistance-cycle graphs of TaO2 and Cu2O memristive devices[17, 30, 39]. Other researchers have also shown that lower differences of the Gibbs free energy of the formation between TaO2 and Ta2O5 show more stable behavior[20].

Table  1.  Gibbs free energy of the formation of metal oxides used in MIM resistive switching devices.
Metal oxideGibbs free energy of formation,Δf G° (kJ/mol)
Hafnium oxide (HfO2)1088:2
Zirconium oxide (ZrO2)1042:8
Titanium oxide (TiO2)888:8
Silicon oxide (SiO2)856:4
Vanadium oxide (VO2)446:4
Nickel oxide (NiO)211:7
Tantalum oxide (TaO2)209:0
Copper oxide (Cu2O)149:0
DownLoad: CSV  | Show Table

Although memristive devices with bulk layers of high Gibbs free energy of formation may also exhibit resistive switching behavior,these devices require an additional one-time electroformation step. This is reflected by TiO2 devices requiring an electroformation process that requires a large electroforming potential before TiO2 devices can exhibit resistive switching behavior. As TiO2 devices mimic an MIM structure,many publications have reported that other MIM structures of different materials also requires an electroforming step before it exhibits memristive behavior[40, 41]. Conversely,TaO2 devices with similar physical dimensions as TiO2 devices would require ∼ 4.5 times less energy for the electroforming process because the Gibbs free energy of the formation of TaO2 is ∼4.5 times smaller than the Gibbs free energy of the formation of TiO2. Moreover,if the Gibbs free energy of the formation is very small,the electroforming voltage may become smaller than the switching voltage of the memristive device[42]. The damage to the device is insignificant and the electroformation process could be part of the first switching process of the device. This avoids the need of a large voltage for a separate electroforming process.

Research also shows that the electroforming voltages of physical TiO2 and TaO2 devices are proportional to bulk layer thickness[28, 43]. This may be in contradiction to Greene's theory,but it is noted that if devices are of the same physical dimensions,then Greene's theory is completely applicable. If two devices have different physical dimensions but similar bulk layer material,then using the Dearnaley theory would be more appropriate.

Dearnaley et al. proposed that the electroforming voltage is proportional to the insulator thickness in MIM devices[44]. They reasoned that this is due to the insulator-metal surface not being completely smooth. Thus,there will be certain regions where the bulk layer is thinner than other regions. Voltage applied at the electrodes is applied across the entire bulk layer surface. The electric field applied during electroforming,EF is the amount of voltage supplied per insulator thickness,given by:

EF=VL

(2)

where V is the voltage applied across the device and L is the bulk material thickness of the device. Electric field is inversely proportional to insulator thickness,thus the electric field is larger at those thinner bulk layer regions. At these high electric field regions,electroformation is more likely to occur and hence forms a conducting channel at these regions. This theory also explains why conducting channels in TiO2 and TaO2 devices are not spanned across the entire electrode area[45] This is discussed further in section 4 of this article.

The electroforming field applied across the bulk layer is inversely proportional to the bulk layer thickness. Hence,insulators thicker than 1 μm cannot be electroformed due to insufficient electroforming field across the bulk layer[2, 34],whereas devices thinner than 10 nm may suffer from permanent dielectric breakdown due to the high electric field[46],or due to large amounts of current flow[47, 48].Therefore,to avoid a permanent breakdown or excess physical damage during electroformation,current compliance methods are commonly used[49].

In order to avoid electroformation in TiO2 devices,two methods were proposed: (i) by thinning the bulk layer to a few nanometers thick[12],and (ii) depositing the bulk layer by reactive sputtering in an oxygen rich ambient[50]. Reducing the device size also decreases the physical damage caused by oxygen gas eruption because smaller devices increase the rate of the release of oxygen molecules out of the device[12].

To avoid electroformation in TaO2 devices,the top electrode Pt is proposed to be replaced with Ta[21]. Although a different electrode is used,the device still behaves with similar memristive properties[42]. Using the magnetron sputtering method,TaO2 devices are able to exhibit memristive behavior without electroformation[31]. This shows that TaO2 devices can exhibit memristive behavior without the need for an additional electroformation process. It also reduces energy consumption by avoiding the need of a large voltage application to electroform the device.

During the fabrication process,oxygen ions are injected into the memristive devices in order to dope the bulk layer. This process is uneven and the oxygen ions are asymmetrical in the bulk layer[51, 52]. It is the asymmetric distribution of oxygen ions (or vacancies) in the TiO2 bulk layer that contributes to the memristive properties of the device. The defect states induced by oxygen vacancies act as the n-type dopant in TiO2 and the highly resistive TiO2 layer becomes more conductive. The asymmetric distribution of oxygen vacancies also causes the TiO2 layer to have regions of high and low resistances. The highly-doped region is less resistive,while the lowly-doped region is more resistive. Due to the asymmetric doping of the bulk layer,only one of the electrodes dominates the switching mechanism,which is the electrode that is nearer to the lowly-doped region of the bulk layer. The formation and collapsing of conducting channels takes place nearer to this electrode[53]. The region where switching cycles will take place,or the electrode that dominates the switching cycles,is determined during fabrication[26, 54].

The switching mechanism in TiO2 devices involves the dislocation of the mobile charged species ions or vacancies towards or away from an electrode. This process either forms or collapses conducting channels[55]. Drifting of oxygen vacancies from the region of highly-doped TiO2 towards the oxygen-deficient region will form areas of lower resistances in the lowly-doped region of the bulk layer[56, 57]. These areas of lower resistances will then combine to form a chain of low resistance regions,which creates the conducting channels. This resembles the switching mechanism of other MIM memristive devices[58].

Positive charge oxygen vacancies are migrated in the bulk layer during switching (Figure 2(a)) forming a filament-like structure,protruding the entire bulk layer (Figure 2(b))[59]. The formation of the filament-like structure conducting channel is localized and does not take place throughout the entire cross-section of the bulk layer[45]. This filament-like structure in the bulk layer is the conducting channel consisting of the TI4O7 phase[41]. It is less resistive than the non-doped regions of the bulk layer,which are composed of the TiO2 phase[55]. The opposing movement of oxygen vacancies transforms the TI4O7 phase in the conducting channel into other Ti-O phases,mainly the TiO2 phase (Figure 2(c)). The high resistance TiO2 phase does not conduct current and the conducting channel becomes highly-resistive. The collapse of the conducting channel eliminates the current conduction path from the bulk layer,so large amounts of current cannot tunnel through the device (Figure 2(d)).

Figure  2.  (Color online) Switching mechanism of TaO2 devices.
(a) A negative bias is applied on the top electrode to switch the device ON,repelling negatively charged oxygen ions away from the Ta2O5 layer towards the TaO2 layer. (b) The Ta2O5 phase is reduced to TaO and Ta. After low-resistance TaO and Ta phases align,it forms conducting channels in the Ta2O5 layer. (c) Positive bias is applied on the top electrode to switch the device OFF,attracting oxygen ions towards the Ta2O5 layer. (d) The oxygen ions oxidize TaO and Ta phases to become Ta2O5 phases,and cause the collapse of conducting channels.

In TaO2 devices,the resistive switching mechanism is assisted by a very thin insulating layer of a highly resistive Ta2O5 phase fabricated between the top electrode and bulk layer[21, 28, 30]. The insulating layer of Ta2O5 consists of both Ta2O5 with a very small percentage of the TaO2 phase.

To turn the device ON,a negative bias is applied at the top electrode and the bottom electrode is grounded. The negative bias at the top electrode repels negative charged oxygen ions away from the top electrode (Figure 3(a)). Ta2O5 molecules are reduced to TaO[60] and Ta[61] molecules by releasing oxygen ions. The oxygen ions drift towards the TaO2 bulk layer. After a number of TaO and Ta molecules align in the Ta2O5 layer to form a conducting channel that is TaO and Ta-rich,large amounts of current can tunnel through the insulating layer. This switches the device into LRS (Figure 3(b)).

To turn the TaO2 device off,a positive bias is applied at the top electrode while the bottom electrode remains grounded. This causes the attraction of oxygen ions from the TaO2 bulk layer towards the top electrode (Figure 3(c)). As the oxygen ions accumulate in the insulating layer,oxidation takes place and TaO and Ta molecules are oxidized into Ta2O5 molecules. The number of Ta2O5 molecules in the conducting channel increases and the resistance of the conducting channel increases. The conducting channel collapses when a number of Ta2O5 molecules disconnect the conducting channel. This prevents current from flowing through the bulk layer and Ta2O5 layer,thus the device is in the HRS (Figure 3(d)).

Figure  3.  (Color online) Switching mechanism of TaO2 devices.
(a) A negative bias is applied on the top electrode to switch the device ON, repelling negatively charged oxygen ions away from the Ta2O5 layer towards the TaO2 layer. (b) The Ta2O5 phase is reduced to TaO and Ta. After low-resistance TaO and Ta phases align, it forms conducting channels in the Ta2O5 layer. (c) Positive bias is applied on the top electrode to switch the device OFF, attracting oxygen ions towards the Ta2O5 layer. (d) The oxygen ions oxidize TaO and Ta phases to become Ta2O5 phases, and cause the collapse of conducting channels.

The switching mechanisms of TiO2 and Ta2O5 devices represent the general switching mechanisms of MIM and MISM devices respectively. It was shown that both MIM and MISM devices use the redox reactions of metal oxides. Metal oxide phases have varying resistivity characteristics. Thus,by performing oxidation and reduction in memristive devices,the conductivity of the bulk layer can be altered.

The major difference between the two types of switching mechanisms is the manner of the formation of the conducting channels. In MIM devices,the conducting channel spans the entire length of the bulk layer of the device,protruding from one electrode to the other. As for MISM devices,the conducting channels are formed in the higher resistance region of the bulk layer[62]. The lower resistance of the bulk layer does not require the formation of a conducting channel in order to be able to conduct current.

Apart from the mechanism formation and collapsing of conducting channels in the substrate of memristive devices,the location of the formation of conducting channels is also different in MIM and MISM devices. In MIM devices,switching takes place at the interface of metal-insulator that is less doped with oxygen vacancies[53, 63]. Researches have also shown that insertion of nano-interface layers into MIM devices improves resistance switching by acting as an oxygen sink[64-66],as well as to reduce switching currents[60]. As for MISM devices,there are also theories that claim the metal-insulator (or metal-semiconductor) interface is responsible for memristive behavior. This metal-insulator (or metal-semiconductor) interface forms a Schottky barrier that is crucial for resistance switching in MISM devices[26, 57, 67]. This type of switching mechanism is further described in References [68, 69].

Generally,switching causes the movement of charged species (e.g. oxygen ions or vacancies) along the length of the device in both switching mechanisms. A conducting channel is formed as soon as there are sufficient oxygen vacancies at a localized switching region. Conversely,if the number of oxygen ions to metal ions (oxygen vacancies in our case) becomes larger than some critical value,Mott-type metal to insulator transition occurs in the switching layer and the conducting channels are collapsed. Due to the charged species having finite mobility,the application of potential difference at the electrodes must exceed a magnitude (known as the switching voltage) for the mobile charged species to have sufficient energy to overcome potential barriers. The potential difference must also be applied longer than a certain amount of time for sufficient charged species to be displaced. Hence,the switching speed of memristive devices is defined as the rate of the device to completely switch resistance states. Switching time is measured starting from when voltage is applied across the device electrodes until switching is complete.

The mobility of charged species through the switching layer is dependent on the material type and the forming field across that layer. The mobility of the mobile charged species in both TiO2 and TaO2 devices is compared between the devices. The oxygen vacancies mobility μ in TiO2 is 1010 cm2/(V⋅s)[3] and in TaO2 is 5.461010 cm2/(V⋅ s)[70]. Due to the vacancy mobility in TaO2 being ∼5 times faster than that in TiO2,it is deduced that the migration of mobile charged species in TaO2 is ∼ 5 times faster than that in TiO2 for similar device dimensions and voltage applied. This means that the time taken to form a conducting channel in TaO2 devices is ∼ 5 times faster than that for TiO2 devices,provided the same amount of voltage is applied and devices are of a similar size. Thus,it is concluded that the average switching speed of TaO2 devices is ∼ 5 times faster than that of TiO2 devices.

The amount of electric field applied and charge carrier mobility are considered when estimating the energy consumed during the switching cycles in the devices. The drift velocity of charge carriers,vd is given as:

vd=μEL

(3)

In order to maintain similar switching speeds (fixed drift velocity of charge carriers) the required application of electric field across the length of the bulk layer is inversely proportional to the charge carrier mobility in the bulk material. Assuming TiO2 and TaO2 devices of similar physical dimensions,it is postulated that to obtain similar switching speeds,the energy requirement of TiO2 devices is larger than that of TaO2 devices. This is due to the charge carrier mobility in TiO2 being smaller than that in TaO2.

Since the electrodes and substrate of TiO2 and TaO2 devices are in series,the instantaneous resistance of \,the devices is the summation of the resistance of each layer. The resistance of the electrodes is always constant because it does not take place in the switching mechanism. The device substrate has different resistance behavior due to the different switching mechanisms in MIM and MISM devices. Although the resistance changes differently depending on device structure,the physical condition of the bulk layer still depends on the history of charge flow through the devices,i.e. HRS or LRS. Thus,the instantaneous resistance of a memristive device depends on both the history of charge flow and its switching mechanism.

Among the results from fabricated memristive devices,the current-voltage curve of each device varies. It is also noticed that the fabricated memristive devices have varying fabrication processes and device dimensions. This tells us that it may be difficult to apply one current-voltage equation for all memristive devices. The current-voltage equation would also have to include other factors such as device size and material properties. Various theories were proposed to simulate the current-voltage characteristics of the devices and in this article,the types of theories that were proposed hitherto and the justifications for those theories are briefly explained.

Drifting of oxygen vacancies along the bulk layer of MIM devices causes the resistance of the bulk layer to change. Initial models assumed that the bulk layer is divided into regions of high and low resistances. The length of the low resistance region is termed as the barrier width,w and is modeled to sweep along the bulk length,$L[55]. The drifting of oxygen vacancies or barrier width along the bulk length was proposed to be linear[3]. The rate of change of barrier width is defined as:

dw(t)dt=μRONLi(t)f(t)

(4)

where i(t) is the amount of current passing through the device and f(t) is a window function with respect to time. The purpose of the window function is to model the behavior of the barrier width when approaching electrode boundaries. A state equation was proposed by Strukov[3] to relate barrier width w(t) and instantaneous resistance of the device:

R(t)=RONx(t)+ROFF[1x(t)].

(5)

where RON is the resistance of the device when it is in LRS and ROFF for when the device is in HRS. x represents the ratio of the thickness of the low resistance region in the bulk layer (w) to the total thickness of the bulk layer (L),giving x=w/L,where x ranges from 0 to 1. The assumption for this equation is that the bulk layer only contains TiO2 and TI4O7 phases. The authors of Reference [3] also assumed that the device could be in a state of intermediate resistance,with conductivity between LRS and HRS.

However,non-linear simulation models have exhibited current-voltage curves which are closer to the physical devices than linear models. The drifting of oxygen vacancies is non-linear in physical devices because the electric field across the device is not linear along the bulk layer due to the asymmetrical distribution of oxygen vacancies in the bulk layer[71]. Furthermore,the drift velocity of charge carriers is strongly affected by the non-linear resistance along the bulk layer. At the electrode boundaries,the charge carriers move into another medium of different charge and material density. This gives rise to a change in charge mobility and causes the drift velocity of the charge carriers to change as they approach or leave electrodes,which adds to the factor of non-linearity[72]

Window functions f(x) were subsequently proposed to model the mobility of charge carriers approaching electrode-bulk layer boundaries. The window functions proposed are summarized in Table 2,in chronological order. The window function to simulate the linear drift model was proposed in 2008[3]. The window function was improved by introducing a parameter p for generalization of non-linear behavior[71]. However,both of these two window functions will cause the simulation of barrier width to be stuck at 0 or 1,and cannot be adjusted further. To avoid this,a window function was proposed to be dependent on current[73],where

Table  2.  Window functions proposed for the behavior of charge carriers approaching electrode boundaries.
Author(s) Window function,f(x)
Strukov et al.[3]f(x)=x(L-w)/L2
Joglekar et al.[71]f(x)=1-(2x-1)2p
Biolek et al.[73]f(x)=1-(x-stp(-i))2p
Prodromakis et al.[74]f(x)=1-[(x-0.5)2+0.75]p
Kvatinsky et al.[76] fon(x)=exp[exp(xaonp)]
foff(x)=exp[exp(xaoffp)]
DownLoad: CSV  | Show Table

stp(i)={1,i00,i<0

(6)

Although the window function of Reference [73] has solved the problem issues and has parameter p for curve-fittings,it lacks scalability. Prodromakis etal. proposed a new window function that improves scalability[74] giving a range of fmax where 0fmax(x)1 .

Up to this time,there are no physical devices that correlate to any of the proposed windows functions. Conversely,Yao et ,al.[16] and Rossel et al.[75] suggested close relations between physical memristive devices and the Simmons tunneling current density equation. Due to the window functions hitherto not fitting the Simmons tunneling current density equation,a window function which correlates to the Simmons equation was proposed[76]. The Kvatinsky window function fits the Simmons equation but it lacks scalability and does not guarantee symmetrical behavior. This means that the rate of change of drift velocity of charge carriers moving towards one electrode may be different from moving towards the opposite electrode. Two more fitting parameters were also introduced in the form of aon and aoff,which are used to fit the model to match the Simmons equation. The Simmons tunneling current density equation is given as:

J=Em0D(Ex)ξdEx,

(7)

where ζ is defined as:

ξ=4πqm2h30[f(E)f(E+eV)]dEr,

(8)

and DEx is a function of the probability of an electron being able to overcome the potential barrier posed by the insulator layer and penetrate through the substrate from one electrode to another. f(E) and f(E+eV) are the Fermi-Dirac functions for electrons following and going against the potential barrier respectively,while dEr is the integral with respect to the effective electron direction. The Simmons tunnel equation is the most accurate due to its consideration of distribution and drifting of electrons in both directions. However,the probability function DEx is ambiguous due to a lack of experiments performed in obtaining it and may be difficult to obtain[12, 77]. Based on these findings,more study on MIM devices is required in order to produce a generic current-voltage equation for the behavior of TiO2 devices.

J=AT2exp(αTdqϕT)exp(qϕBkT)(expqVηkT1).

(9)

Equation (9) involves the thermionic-emission for the Schottky barrier and includes a tunneling probability term, exp(αTdqϕT) . A* is the Richardson constant,q is electron charge,V is voltage applied,T is temperature,k is the Boltzmann constant, ϕT is the effective barrier height,d is insulator layer thickness,and αT is the base transport factor term. αT approaches 1 if the effective mass in the insulator is equivalent to the free electron mass. The Schottky barrier height ϕB term exists in the tunneling current density equation due to the movement of charge carriers between metal and semiconductor through a very thin insulator layer mimicking that of a Schottky-barrier diode. Ideality factor, η is close to the value of 1 when the amount of doping in the insulator layer is small,and becomes a larger value when doping increases. Using this equation,simulation models have been developed and they match the characteristics displayed by physical TaO2 devices[68, 79].

The switching resistance ratios (ROFF/RON) of TiO2 devices range from 500 to 105,[50, 51, 53, 80, 81],while the switching resistance ratios of TaO2 devices range from 10 to 106,[19, 30, 42]. It is estimated that the average resistance ratio of future fabricated TaO2 devices will be higher than that of TiO2 devices due to the resistance ratios of TaO2 devices reaching 106,while TiO2 devices reach up to only 105 of resistance ratios. Comparing band gaps of TiO2 and TaO2/Ta2O5,the literature reported that wider band gaps are advantageous for resistive switching[82]. The band gap of TaO2 is 4.0-4.2 eV[83],Ta2O5 is 3.9-5.3 eV[84, 85],and TiO2 is 3.0-3.2 eV[82, 86]. Therefore,this adds weight to the choice of TaO2 devices over TiO2 devices.

In order for memristive devices to be able to replace physical memory devices,they need to be able to consistently switch between resistance states. Although a slight deviation of the resistance value is allowed,a stable ROFF/RON ratio is required. In both devices,the HRS will vary much more than the LRS,with an example of resistance switching given in Figure 4. The HRS is formed by collapsing conducting channels,which is not a consistent process because the amount of charge carriers needed to collapse a conducting channel may vary during each switching. In the LRS,the conducting channel effectively shorts the device,so the only resistance along the device is the resistance of the electrodes,which is constant[87, 88]. Most of the TiO2 and TaO2 devices have exhibited an almost consistent ROFF/RON resistance ratio[30, 89]. Therefore,both TiO2 and TaO2 devices are two of the suitable candidates of memristive devices to replace physical memory devices in terms of stable resistive states.

Figure  4.  Resistance switching of memristive devicesŒ.[87]

Retention and endurance are common measures of the reliability of memory and memristive devices. Retention is the ability of memory and memristive devices to retain logic information,measured by the amount of time that the memory is still retrievable. Endurance is the measurement of the number of times a device can perform a successful read and write process. It is equivalent to the number of switching cycles that the device can perform successfully.

The retention time of a memristive device is defined as the time taken for the device to deteriorate from its resistance state (HRS or LRS) by a certain amount of percentage,R%. The threshold for the percentage of change,R% is determined by researchers or manufacturers. Smaller values of R% yield a better quality of retention in exchange for a shorter retention period. Conversely,larger values of R% result in a longer retention period but yield a device with a poor retention quality.

The retention time of memory devices at room temperature or normal operating conditions is predicted using extrapolation of the Arrhenius plot. This method has been used by many other researchers[90-92]. The Arrhenius equation defines the rate constant τ of a chemical reaction as:

τ=A\meEa/kT,

(10)

where Ea is the activation energy of the reaction. A is a pre-exponential factor or frequency factor that depends on the frequency of collision between molecules,while the rate constant τ is the frequency of a successful reaction between reactants. Taking the natural logarithm,Equation (10) becomes:

lnτ=EAk(1T)+lnA.

(11)

This method uses temperature-accelerated degradation of memristive devices and the percentage of change in resistance or conductivity,ΔR is observed. The device is degraded past R% at a specific constant temperature T. A graph of ΔR is plotted against time,and the gradient is the rate constant τ for temperature T (Figure 5). A similar procedure is repeated to obtain rate constants at different temperatures.

Figure  5.  Temperature-accelerated degradation plot at temperature T.

Temperature-accelerated degradations are usually performed at temperatures much higher than room temperature or normal operating conditions,in the range between 150 ℃ to 250 ℃[90-92]. The Arrhenius plot is then drawn,where lnτ is plotted against T1 (Figure 6). The graph is then extrapolated until the rate constant at room temperature or at normal operating conditions is found.

Figure  6.  Arrhenius plot to find the rate constant at room temperature.
40.0×103-1 corresponds to a room temperature of 25 °C,while 4.0,5.0 and 6.7 (×10-3) °C-1 corresponds to the temperature-accelerated degradation performed at 250,200,and 150 °C respectively.

This approach was used to obtain the retention time for TaO2 devices operating between 25 ° and 85 °,which is of over 10 years[30]. This retention time meets the standards of manufacturers of memory devices[90]. At the time of writing this article,the retention of TiO2 devices has not yet been obtained using extrapolation of the Arrhenius plot. There were reports stating that TiO2 devices have a retention time as long as 10 years but there are no experimental data or methods that could provide sufficient evidence to support this statement. Although different fabrication methods and/or slightly different electrodes were tried,experiment data could still not provide sufficient evidence for retention of 10 years[11, 93-95]. Due to the lack of robust methods in obtaining the retention period of TiO2 devices,it cannot be concluded yet that TiO2 devices have the retention ability of 10 years.

The endurance of memristive devices is measured by the number of resistance switching cycles that can be successfully performed until the resistance ratio converges towards unity,or when the resistance ratio becomes less than a required amount. The minimum resistance ratio is determined by the application that uses the memristive device. The method to test device endurance is by switching the device continuously until it fails. Recent publications have shown that TaO2 devices (1010 cycles[31]) have better endurance than TiO2 devices (104 cycles[96]). It was also shown that after 104 cycles,the resistance states of TiO2 converge and no longer exhibit memristive properties[21],while endurance of TaO2 devices can reach up to 1012 cycles via increasing the device size[26]. Increasing the device size improves endurance due to the larger space available to produce more conducting channels. In larger devices,if for any reason a conducting channel fails to reconnect while turning-ON,then the turning-ON voltage will electroform another new conducting channel in the bulk layer. Thus,ensuring another switching cycle can occur.

TaO2 devices exhibit better endurance than TiO2 devices due to the multiple stable Magnéli phases of titanium oxide (Ti-O)[97] compared to the tantalum oxide (Ta-O) series that only has two stable Magnéli phases[98]. Figures 7 and 8 show the phase diagrams of Ti-O and Ta-O. Magnéli phases of oxides appear when crystallographic shear occurs that changes the stoichiometry of the cation (metal) but still maintaining the coordination requirements of the cation. The homologous series of metal-oxide phases are made up of Magnéli phases of metal-oxides[99]. Due to the unstable TI4O7 phase and multiple Magnéli phases in the Ti-O series,the doping of TiO2 results in the formation of oxide phases other than TI4O7[41]. After a number of switching cycles,the redox reaction that takes place in the bulk layer forms Ti-O phases other than TiO2 and TI4O7. The other phases of Ti-O are also in thermal equilibrium with TiO2 and TI4O7,which increases the reluctance of the other phases to react. With the increase of other higher resistive phases in the bulk layer,the device will not be able to switch back to its initial RON and ROFF states. This means that the TiO2 device has reached its endurance limit.

Figure  7.  Ti-O phase diagram[97].
Figure  8.  Ta-O phase diagram[98].

For TaO2 devices,the Magnéli series consist of only two stable phases: TaO2 and Ta2O5. This means that every chemical reaction in the bulk layer will only involve either of these two Ta-O phases,which improves device endurance. Therefore,it is concluded that TaO2 devices have better endurance than TiO2 devices due to the lesser number of stable Magnéli phases in the Ta-O series.

Apart from using materials with a lesser number of stable Magnéli phases,the endurance of resistive switching devices can be improved by regulating the electroformation process[29]. By minimizing the electroforming process,physical damages are minimized. Thus,preserving the original structure of the device and improving the endurance of the device in the physical aspect,as previously discussed in section 3 of this article.

The thick bulk layer in MIM memristive devices prevents conducting channels from being formed due to the large amount of electric field required to breakdown the thick bulk layer. This is shown by the decrease in probability for the formation of a conducting channel with an increase in bulk layer thickness[89]. The range of bulk layer that consistently exhibits memristive behavior is below 500 nm[82],while bulk layers thinner than 50 nm[41,\thinspace 100] or even 10 nm[101] can also show memristive properties. Switching voltages are observed to be independent with varying bulk layer thickness[34]. This is due to the forming and deforming of the conducting channel that occurs only at localized areas. The majority of the applied potential difference is dropped across the switching region,which is almost constant in size and independent of bulk layer thickness. Thus,the switching bias remains fairly constant with varying bulk layer thickness[41, 43, 98].

The resistivity of the HRS of MIM memristive devices increases with the thicker bulk layer due to the charge carriers needing to drift through a greater distance along the device. There are no conducting channels in the HRS of the device,thus the greater the distance,the higher the resistivity of the bulk layer. Conversely,the resistivity of LRS remains fairly constant because the device is fundamentally shorted in the LRS. This is in accordance to the published results in Reference [53].

In TaO2 devices,the substrate is composed of two layers,the bulk layer with the majority TaO2 phase and the insulator layer with the majority Ta2O5 phase. Because there are two layers with different phases of Ta-O,two types of comparison are performed: (i) devices with same layer ratio but different substrate thickness,and (ii) devices with different layer ratios.

TaO2 devices with different substrate thickness but constant thickness ratio between TaO2 and Ta2O5 layers are discussed here. It was observed that the trend is similar to that of TiO2 devices and the switching bias is fairly constant for a constant layer ratio between layers[102]. The reason for this trend is due to the similar characteristics of HRS and LRS of TaO2 and TiO2 devices.

For TaO2 devices with different layer ratios[28],there are two types of comparison: (i) different TaO2 bulk layer thickness with constant Ta2O5 insulator layer,and (ii) different Ta2O5 layer thickness with constant TaO2 bulk layer.

Varying TaO2 layer thickness (fixed Ta2O5 layer thickness) does not change the amount of current flow in the LRS due to the device being shorted in LRS. In the HRS,it is observed that the current flow is less in the device with the thicker TaO2 layer 10-8A compared to 10-6A),because a longer device increases the overall device resistance. Due to switching regions located inside the Ta2O5 insulator layer,thus varying the TaO2 layer does not affect the switching-ON bias (-1.5 V). However,the switching-OFF bias is larger with the thicker TaO2 layer. This is due to the longer distance that the oxygen ions need to travel in the thicker TaO2 layer. Since the Ta2O5 layer is constant and assuming the switching region is at fixed locations,this makes the drifting of oxygen ions from the TaO2 layer towards the Ta2O5 layer more difficult due to the longer distance. Therefore,a larger switching-OFF bias is required to collapse the conducting channel,and it is concluded that varying the TaO2 layer affects the HRS but not the LRS of the device[28].

Varying the Ta2O5 layer thickness (fixed TaO2 layer thickness) changes the current flow in the LRS and HRS of the device. Increasing the thickness of the Ta2O5 layer decreases the amount of current flow through the device in both LRS and HRS. Although it is documented that the device is short-circuited-like in the LRS,it is not entirely true within the Ta2O5 layer. This is because the MISM structure gives a similar effect to that of a Schottky barrier. This is also mathematically shown in Equation (9) that current density J is inversely proportional to insulator layer thickness d. Another observation is that both switching-ON and switching-OFF biases remain constant with different Ta2O5 thickness. Since the thickness of the TaO2 bulk layer is fixed,the distance between the switching regions in the Ta2O5 layer to the TaO2 layer is constant. Therefore,the switching-ON and switching-OFF biases are constant due to a similar amount of electric field being applied across the device[28]. It is concluded that the Ta2O5 layer affects the amount of current in the LRS and HRS of the devices. This conclusion is also based on the fact that the thickness of the Ta2O5 layer is the dominating factor in manipulating the device resistance ratio[26].

According to Ohm's law,the resistance of a conductor is inversely proportional to the cross-section area of the conductor. Increasing the memristive device size (defined by the cross-section area of electrodes) would decrease ROFF only. In the HRS of the devices,there are no conducting channels,thus the electric field is applied across the entire cross-section of the bulk layer. Hence,ROFF decreases with increasing device size.

Increasing the device size of memristive devices does not linearly decrease $R_{\mathrm{ON of TiO2 and TaO2 devices[26]. When a memristive device is in LRS,it forms a thin conducting channel with the diameter much smaller than the electrode cross-section area. The conducting channels also do not span across the entire electrode[103]. This observation was also reported for other memristive devices with similar device

structures[14, 104]. Increasing device size does not necessarily mean that the conducting channels increase in size. However,it increases the probability of more conducting channels being formed[41],but not necessarily all conducting channels are switched on when the memristive device is switched ON. Thus,since larger devices do not contribute to a large increase in conductivity,then it would be a waste to fabricate larger devices. In summary,switching parameters may be affected by device size,but the switching mechanism remains the same and overall the memristive behavior is still exhibited,which means memristive devices are scalable devices.

Electrode material does not take place in the switching process,but it affects the manner chemical reactions taking place in the bulk layer[105]. Using the same bulk layer material,it is possible to fabricate either a unipolar resistive switching device or a bipolar resistive switching device by using different electrode materials[106, 107].

For TiO2 devices,one of the electrodes may be replaced with other materials that still allow bipolar resistive switching to take place,such as: silver,Ag[86],aluminum,Al[108],and tungsten,W[109]. For TaO2 devices,other electrode materials that allow bipolar resistive switching to take place are such as: Al[110],copper,Cu[19],nickel,Ni[111],and W[112]. Among the electrodes used to fabricate TiO2 and TaO2 devices,a common trend could not be found and the effect on the performance of the devices varies,albeit several similar electrode materials used to fabricate TiO2 and TaO2 devices.

Electrode material selection is vital to enable resistive switching to take place in a device[113]. Resistive switching cannot take place in a device if both top and bottom electrodes create ohmic contacts with the bulk layer. Resistive switching can only take place when at least one of the electrodes creates a Schottky barrier with the bulk layer. This also explains why certain electrode materials can be used to fabricate memristive devices with different bulk layer materials,such as tungsten,W which was used as the electrode material to fabricate memristive devices with bulk layers of TiO2[109],TaO2[112],silicon oxide,SiO2[114],and hafnium oxide,HfO2[115].

In this article,a review on the comparison between TiO2 and TaO2 memristive devices was presented. The comparison showed that TaO2 memristive devices have shown better electrical properties than TiO2 memristive devices that are of a similar physical size. This is due to the following material factors: (1) oxygen vacancies charge carrier mobility in TaO2 is higher than that in TiO2,(2) fewer stable Magnéli phases in the Ta-O Magnéli series than in the Ti-O Magnéli series,and (iii) smaller Gibbs free energy of the formation of TaO2.

For memristive devices with similar dimensions and voltage applied,the switching speed of TaO2 memristive devices is faster than TiO2 memristive devices because the charge carrier mobility in TaO2 is higher than that in TiO2. Electroformation energy requirements for TaO2 devices are also lesser due to the smaller Gibbs free energy of formation. TaO2 devices have demonstrated memristive behavior without the electroformation process,while TiO2 devices require an electroformation process unless additional fabrication steps are taken. Steps such as thinning the bulk layer or fabricating in high-pressure oxygen ambient allows MIM devices to exhibit memristive behavior without electroformation.

The retention capability of memristive devices has been predicted to exceed 10 years. Using temperature-accelerated degradation and the Arrhenius plot,TaO2 memristive devices exhibit good retention past 10 years at room temperatures. Although TiO2 devices exhibit excellent retention properties at initial stages,no temperature-accelerated degradation experiment was performed to provide such results. The endurance of memristive devices can be improved by reducing electroformation effects or using a bulk layer material that has fewer stable Magnéli phases,such as the Ta-O Magnéli series. The electrical properties of both TiO2 and TaO2 devices show similar scalability on changing device dimensions. However,memristive devices are usually thinner than 10 nm because increasing the device size does not contribute much to an improvement of electrical properties.

This article concludes that lesser Magnéli phases contribute to higher endurance,smaller Gibbs free energy of formation leads to smaller electroformation energies,and faster oxygen vacancies mobility provides faster resistance switching of memristive devices. For memristive devices to replace current memory applications,electrical characteristics should be calculable pre-fabrication. At the moment,accurate electrical properties of most memristive devices are only obtainable post-fabrication. Moving forward,memristive devices have great potential in increasing the density of nanoscale devices,but more research is needed to ensure a more stable resistance switching and increased endurance performance.



[1]
Chua L. Memristor-the missing circuit element. IEEE Trans Circuit Theory, 1971, 18(5):507
[2]
Schroeder H, Jeong D S. Resistive switching in a Pt/TiO2/Pt thin film stack-a candidate for a non-volatile ReRAM. Microelectron Eng, 2007, 84(9-10):1982
[3]
Strukov D B, Snider G S, Stewart D R, et al. The missing memristor found. Nature, 2008, 453(1):80
[4]
Yang J J, Strukov D B, Stewart D R. Memristive devices for computing. Nat Nanotechnol, 2013, 8:13
[5]
Bez R, Pirovano A. Non-volatile memory technologies:emerging concepts and new materials. Mater Sci Semicond Process, 2004, 7(4-6):349
[6]
Duan S, Hu X, Wang L, et al. Memristor-based RRAM with applications. Sci China Inf Sci, 2012, 55(6):1446
[7]
Batas D, Fiedler H. A memristor SPICE implementation and a new approach for magnetic flux-controlled memristor modeling. IEEE Trans Nanotechnol, 2011, 10(2):250
[8]
Jo S H, Chang T, Ebong I, et al. Nanoscale memristor device as synapse in neuromorphic systems. Nano Lett, 2010, 10(4):1297
[9]
Vontobel P O, Robinett W, Kuekes P J, et al. Writing to and reading from a nano-scale crossbar memory based on memristors. Nanotechnology, 2009, 20(42):1
[10]
Eshraghian K, Cho K, Kavehei O, et al. Memristor MOS content addressable memory (MCAM):hybrid architecture for future high performance search engines. IEEE Trans VLSI, 2011, 19(8):1407
[11]
Yoshida C, Tsunoda K, Noshiro H, et al. High speed resistive switching in Pt/TiO2/TiN film for nonvolatile memory application. Appl Phys Lett, 2007, 91(22):223510
[12]
Yang J J, Miao F, Pickett M D, et al. The mechanism of electroforming of metal oxide memristive switches. Nanotechnology, 2009, 20(21):215201
[13]
Boboila S, Desnoyers P. Write endurance in flash drives:measurements and analysis. USENIX Conference on File and Storage Technologies, 2010:115
[14]
Rajachidambaram J S, Murali S, Conley J F, et al. Bipolar resistive switching in an amorphous zinc tin oxide memristive device. J Vac Sci Technol B, 2013, 31(1):01A104
[15]
Driscoll T, Kim H T, Chae B G, et al. Phase-transition driven memristive system. Appl Phys Lett, 2009, 95(4):043503
[16]
Yao J, Sun Z, Zhong L, et al. Resistive switches and memories from silicon oxide. Nano Lett, 2010, 10(10):4105
[17]
Chen A, Haddad S, Wu Y J, et al. Non-volatile resistive switching for advanced memory applications. IEEE International Electron Devices Meeting, 2005:746
[18]
Zhang X. Resistive switching characteristics of Ni/HfO2/Pt ReRAM. J Semicond, 2012, 33(5):054011
[19]
Sakamoto T, Banno N, Iguchi N, et al. A Ta2O5 solid-electrolyte switch with improved reliability. IEEE Symposium on VLSI Technology, 2007:38
[20]
Prakash A, Jana D, Maikap S. TaO_x-based resistive switching memories:prospective and challenges. Nanoscale Res Lett, 2013, 8(1):418
[21]
Yang J J, Zhang M X, Strachan J P, et al. High switching endurance in TaOX memristive devices. Appl Phys Lett, 2010, 97(23):232102
[22]
Akinaga H, Shima H. Resistive random access memory (ReRAM):a metal oxide memory cell. Proc IEEE, 2010:98(12):2237
[23]
Fujisaki Y. Overview of emerging semiconductor non-volatile memories. IEICE Electron Express, 2012, 9(10):908
[24]
Mazumder P, Kang S M, Waser R. Memristors:devices, models, and applications. Proc IEEE, 2012, 100(6):1911
[25]
Ho P W C, Hatem F O, Almurib H A F, et al. Comparison on TiO2 and TaO2 based bipolar resistive switching devices. IEEE International Conference on Electronic Design (ICED), 2014:249
[26]
Lee M J, Lee C B, Lee D, et al. A fast, high-endurance and scalable non-volatile memory device made from asymmetric Ta2O5-x/TaO2-x bilayer structures. Nat Mater, 2011, 10(8):625
[27]
Yoo H K, Kang B S, Lee S B. Forming time of conducting channels in double-layer Pt/Ta2O5/TaOx/Pt and single-layer Pt/TaOx/Pt resistance memories. Thin Solid Films, 2013, 540:190
[28]
Yoo H K, Lee S B, Lee J S, et al. Conversion from unipolar to bipolar resistance switching by inserting Ta2O5 layer in Pt/TaOX/Pt cells. Appl Phys Lett, 2011, 98(18):183507
[29]
Lee S B, Yoo H K, Kim K, et al. Forming mechanism of the bipolar resistance switching in double-layer memristive nanodevices. Nanotechnology, 2012, 23(31):315202
[30]
Wei Z, Kanzawa Y, Arita K, et al. Highly reliable TaOx ReRAM and direct evidence of redox reaction mechanism. IEEE International Electron Devices Meeting, 2008:1
[31]
Fukuda N, Fukuju K, Yogosawa I, et al. Forming-free resistance random access memory using Ta2O5/TaOx bi-layer prepared by magnetron sputtering method. MRS Proc, 2012, 1430
[32]
Pagnia H, Sotnik N. Bistable switching in electroformed metal-insulator-metal devices. Phys Status Solidi, 1988, 108(1):11
[33]
Collins R A, Bowman G, Sutherland R R. Electroforming of MIM thin-film devices in air at atmospheric pressure. J Phys D Appl Phys, 1971, 4(11):49
[34]
Dearnaley G, Stoneham A M, Morgan D V. Electrical phenomena in amorphous oxide films. Reports Prog Phys, 1970, 33(3):1129
[35]
Thurstans R E, Oxley D P. The electroformed metal-insulator-metal structure:a comprehensive model. J Phys D Appl Phys, 2002, 35(8):802
[36]
Strachan J P, Strukov D B, Borghetti J, et al. The switching location of a bipolar memristor:chemical, thermal and structural mapping. Nanotechnology, 2011, 22(25):254015
[37]
Münstermann R, Yang J J, Strachan J P, et al. Morphological and electrical changes in TiO2 memristive devices induced by electroforming and switching. Phys Status Solidi-Rapid Res Lett, 2010, 4(1/2):16
[38]
Dean J A. Lange's handbook of chemistry. 15th ed. New York, USA:McGraw-Hill, Inc, 1999
[39]
Lv H, Wang M, Wan H, et al. Endurance enhancement of Cu-oxide based resistive switching memory with Al top electrode. Appl Phys Lett, 2009, 94(21):213502
[40]
Nauenheim C, Kuegeler C, Ruediger A, et al. Investigation of the electroforming process in resistively switching TiO2 nanocrosspoint junctions. Appl Phys Lett, 2010, 96(12):122902
[41]
Kwon D H, Kim K M, Jang J H, et al. Atomic structure of conducting nanofilaments in TiO2 resistive switching memory. Nat Nanotechnol, 2010, 5:148
[42]
Miao F, Strachan J P, Yang J J, et al. Anatomy of a nanoscale conduction channel reveals the mechanism of a high-performance memristor. Adv Mater, 2011, 23(47):5633
[43]
Hosotani K, Park S G, Nishi Y. Electric field dependent switching and degradation of resistance random access memory. IEEE International Integrated Reliability Workshop Final Report, 2009:11
[44]
Dearnaley G, Morgan D V, Stoneham A M. A model for filament growth and switching in amorphous oxide films. J Non Cryst Solids, 1970, 4:593
[45]
Kim K M, Choi B J, Hwang C S. Localized switching mechanism in resistive switching of atomic-layer-deposited TiO2 thin films. Appl Phys Lett, 2007, 90(24):242906
[46]
Sire C, Blonkowski S, Gordon M J, et al. Statistics of electrical breakdown field in HfO2 and SiO2 films from millimeter to nanometer length scales. Appl Phys Lett, 2007, 91(24):242905
[47]
Murali S, Rajachidambaram J S, Han S Y, et al. Resistive switching in zinc-tin-oxide. Solid State Electron, 2013, 79:248
[48]
Yang M K, Park J W, Ko T K, et al. Resistive switching characteristics of TiN/MnO2/Pt memory devices. Phys Status Solidi-Rapid Res Lett, 2010, 4(8/9):233
[49]
Lu Y M, Noman M, Chen W, et al. Elimination of high transient currents and electrode damage during electroformation of TiO2-based resistive switching devices. J Phys D Appl Phys, 2012, 45(39):395101
[50]
Salaoru I, Prodromakis T, Khiat A, et al. Resistive switching of oxygen enhanced TiO2 thin-film devices. Appl Phys Lett, 2013, 102(1):013506
[51]
Huang J J, Kuo C W, Chang W C, et al. Transition of stable rectification to resistive-switching in Ti/TiO2/Pt oxide diode. Appl Phys Lett, 2010, 96(26):262901
[52]
Hudec B, Hranai M, Hušeková K, et al. Resistive switching in RuO2/TiO2/RuO2 MIM structures for non-volatile memory application. International Conference on Advanced Semiconductor Devices & Microsystems, 2010:255
[53]
Choi B J, Jeong D S, Kim S K, et al. Resistive switching mechanism of TiO2 thin films grown by atomic-layer deposition. J Appl Phys, 2005, 98(3):033715
[54]
Do Y H, Kwak J S, Hong J P, et al. Al electrode dependent transition to bipolar resistive switching characteristics in pure TiO2 films. J Appl Phys, 2008, 104(11):114512
[55]
Strachan J P, Pickett M D, Yang J J, et al. Direct identification of the conducting channels in a functioning memristive device. Adv Mater, 2010, 22(32):3573
[56]
Yang J J, Pickett M D, Li X, et al. Memristive switching mechanism for metal/oxide/metal nanodevices. Nat Nanotechnol, 2008, 3(7):429
[57]
Sawa A. Resistive switching in transition metal oxides. Mater today, 2008, 11(6):28
[58]
Yasuhara R, Fujiwara K, Horiba K, et al. Inhomogeneous chemical states in resistance-switching devices with a planar-type Pt/CuO/Pt structure. Appl Phys Lett, 2009, 95(1):012110
[59]
Sun P, Su L, Ling L, et al. Simulation study of conductive filament growth dynamics in oxide-electrolyte-based ReRAM. J Semicond, 2014, 35(10):104007
[60]
Park G S, Kim Y B, Park S Y, et al. In situ observation of filamentary conducting channels in an asymmetric Ta2O5-x/TaO2-x bilayer structure. Nat Commun, 2013, 4:2382
[61]
Strachan J P, Medeiros-Ribeiro G, Yang J J, et al. Spectromicroscopy of tantalum oxide memristors. Appl Phys Lett, 2011, 98(24):242114
[62]
Bi J, Han Z. Characteristics of HfO2/Hf-based bipolar resistive memories. J Semicond, 2015, 36(6):064010
[63]
Kim K M, Choi B J, Shin Y C, et al. Anode-interface localized filamentary mechanism in resistive switching of TiO2 thin films. Appl Phys Lett, 2007, 91(1):012907
[64]
Lin C Y, Wu C Y, Wu C Y, et al. Modified resistive switching behavior of ZrO2 memory films based on the interface layer formed by using Ti top electrode. J Appl Phys, 2007, 102(9):094101
[65]
Prakash A, Maikap S, Chiu H C, et al. Enhanced resistive switching memory characteristics and mechanism using a Ti nanolayer at the W/TaOX interface. Nanoscale Res Lett, 2014, 9(1):125
[66]
Zhuo V Y, Jiang Y, Zhao R, et al. Improved switching uniformity and low-voltage operation in TaO-based RRAM using Ge reactive layer. IEEE Electron Device Lett, 2013,34(9):1130
[67]
Seong D, Jo M, Lee D, et al. HPHA effect on reversible resistive switching of Pt/Nb-doped SrTiO3 Schottky junction for nonvolatile memory application. Electrochem Solid-State Lett, 2007, 10(6):168
[68]
Hatem F O, Ho P W C, Kumar T N, et al. Modeling of bipolar resistive switching of a nonlinear MISM memristor. Semicond Sci Technol, 2015, 30(11):115009
[69]
Kim S, Choi S, Lu W. Comprehensive physical model of dynamic resistive switching in an oxide memristor. ACS Nano, 2014
[70]
Mattsson M S, Niklasson G A, Forsgren K, et al. A frequency response and transient current study of ��-Ta2O5:methods of estimating the dielectric constant, direct current conductivity, and ion mobility. J Appl Phys, 1999, 85(4):2185
[71]
Joglekar Y N, Wolf S J. The elusive memristor:properties of basic electrical circuits. Eur J Phys, 2009, 30(4):661
[72]
Strukov D B, Williams R S. Exponential ionic drift:fast switching and low volatility of thin-film memristors. Appl Phys A, 2009, 94(3):515
[73]
Biolek Z, Biolek D, Biolková V. SPICE model of memristor with nonlinear dopant drift. Radio Engineering, 2009, 18(2):210
[74]
Prodromakis T, Peh B P, Papavassiliou C, et al. A versatile memristor model with non-linear dopant kinetics. IEEE Trans Electron Devices, 2011, 58(9):3099
[75]
Rossel C, Meijer G I, Breìmaud D, et al. Electrical current distribution across a metal-insulator-metal structure during bistable switching. J Appl Phys, 2001, 90(6):2892
[76]
Kvatinsky S, Friedman E G, Kolodny A, et al. TEAM:threshold adaptive memristor model. IEEE Trans Circuits Syst I, 2013, 60(1):211
[77]
Zhang L, Chen Z, Yang J J, et al. A compact modeling of TiO2-TiO2-x memristor. Appl Phys Lett, 2013, 102(15):153503
[78]
Sze S M, Ng K K. Physics of semiconductor devices. 3rd ed. New Jersey, USA:John Wiley & Sons, 2007
[79]
Ji H H, Lee M J, Chang L B, et al. Modeling for bipolar resistive memory switching in transition-metal oxides. Phys Rev B, 2010, 82(15):155321
[80]
Pickett M D, Strukov D B, Borghetti J L, et al. Switching dynamics in titanium dioxide memristive devices. J Appl Phys, 2009, 106(7):074508
[81]
Zhang F, Gan X, Li X, et al. Realization of rectifying and resistive switching behaviors of TiO2 nanorod arrays for nonvolatile memory. Electrochem Solid-State Lett, 2011, 14(10):H422
[82]
Acharyya D, Hazra A, Bhattacharyya P. A journey towards reliability improvement of TiO2 based resistive random access memory:a review. Microelectron Reliab, 2014, 54(3):541
[83]
Franke E, Trimble C L, DeVries M J, et al. Dielectric function of amorphous tantalum oxide from the far infrared to the deep ultraviolet spectral region measured by spectroscopic ellipsometry. J Appl Phys, 2000, 88(9):5166
[84]
Kukli K, Aarik J, Aidla A, et al. Properties of tantalum oxide thin films grown by atomic layer deposition. Thin Solid Films, 1995, 260(2):135
[85]
Fleming R M, Lang D V, Jones C D W, et al. Defect dominated charge transport in amorphous Ta2O5 thin films. J Appl Phys, 2000, 88(2):850
[86]
Cao X, Li X, Yu W, et al. Bipolar resistive switching properties of microcrystalline TiO2 thin films deposited by pulsed laser deposition. Mater Sci Eng B, 2009, 157(1-3):36
[87]
Blaschke D, Zahn P, Skorupa I, et al. Resistive switching in thermally oxidized titanium films. International Semiconductor Conference Dresden-Grenoble (ISCDG), 2013:1
[88]
Lee S R, Kim Y, Chang M, et al. Multi-level switching of triple-layered TaO_x RRAM with excellent reliability for storage class memory. Symposium on VLSI Technology, 2012:71
[89]
Hsiung C P, Gan J Y, Tseng S H, et al. Resistance switching characteristics of TiO2 thin films prepared with reactive sputtering. Electrochem. Solid-State Lett, 2009, 12(7):G31
[90]
Niset M, Kuhn P. Typical data retention for nonvolatile memory. Freescale Semiconductor Engineering Bulletin, 2005:EB618/D
[91]
Jo S H, Kim K H, Lu W. Programmable resistance switching in nanoscale two-terminal devices. nano Lett, 2009, 9:496
[92]
Chen A, Haddad S, Wu Y C. A temperature-accelerated method to evaluate data retention of resistive switching nonvolatile memory. IEEE Electron Device Lett, 2008, 29(1):38
[93]
Young H Do, June S K, Yoon C B, et al. Hysteretic bipolar resistive switching characteristics in TiO2-TiO2-x multilayer homojunctions. Appl Phys Lett, 2009, 95(9):093507
[94]
Jung S, Kong J, Song S, et al. Resistive switching characteristics of solution-processed transparent TiO_x for nonvolatile memory application. J Electrochem Soc, 2010, 157(11):1042
[95]
Chang W Y, Cheng K J, Tsai J M, et al. Improvement of resistive switching characteristics in TiO2 thin films with embedded Pt nanocrystals. Appl Phys Lett, 2009, 95(4):042104
[96]
Xia Q, Pickett M D, Yang J J, et al. Impact of geometry on the performance of memristive nanodevices. Nanotechnology, 2011, 22(25):254026
[97]
Okamoto H. O-Ti (oxygen-titanium). J Phase Equilibria Diffus, 2011, 32(5):473
[98]
Garg S P, Krishnamurthy N, Awasthi A, et al. The O-Ta (oxygen-Tantalum) system. J Phase Equilibria, 1996, 17(1):63
[99]
Kihlborg L, Olovsson I. Arne Magnéli, 1914-1996. Acta Cryst A, 1997, 53(2):103
[100]
Yang J J, Strachan J P, Xia Q, et al. Diffusion of adhesion layer metals controls nanoscale memristive switching. Adv Mater, 2010, 22(36):4034
[101]
Park J, Jung S, Lee J, et al. Resistive switching characteristics of ultra-thin TiO_x. Microelectron Eng, 2011, 88(7):1136
[102]
Lee T W, Nickel J H. Memristor resistance modulation for analog applications. IEEE Electron Device Lett, 2012, 33(10):1456
[103]
Tedesco J L, Stephey L, Hernández-Mora M, et al. Switching mechanisms in flexible solution-processed TiO2 memristors. Nanotechnology, 2012, 23(30):305206
[104]
Lv H B, Yin M, Fu X F, et al. Resistive memory switching of CuXO films for a nonvolatile memory application. IEEE Electron Device Lett, 2008, 29(4):309
[105]
Gu T. Role of oxygen vacancies in TiO2-based resistive switches. J Appl Phys, 2013, 113(3):033707
[106]
Ko D S, Kim S I, Ahn T Y, et al. Effect of the electrode materials on the resistive switching of Ti4O7. Appl Phys Lett, 2012, 101(5):053502
[107]
Lin K L, Hou T H, Shieh J, et al. Electrode dependence of filament formation in HfO2 resistive-switching memory. J Appl Phys, 2011, 109(8):084104
[108]
Jeong H Y, Kim Y I, Lee J Y, et al. Choi, A low-temperature-grown TiO2-based device for the flexible stacked RRAM application. Nanotechnology, 2010, 21(11):115203
[109]
Biju K P, Liu X, Bourim E M, et al. Asymmetric bipolar resistive switching in solution-processed Pt/TiO2/W devices. J Phys D Appl Phys, 2010, 43(49):495104
[110]
Chen C, Gao S, Zeng F, et al. Migration of interfacial oxygen ions modulated resistive switching in oxide-based memory devices. J Appl Phys, 2013, 114(1):014502
[111]
Yoon M J, Lee S B, Yoo H K, et al. Dielectric-breakdown-like forming process in the unipolar resistance switching of Ta2O5-X thin films. Curr Appl Phys, 2012, 1293:846
[112]
Prakash A, Maikap S, Chen G, et al. Improvement of high-�� Ta2O5-based resistive switching memory using Ti interfacial layer. International Symposium on VLSI Technology, Systems and Applications (VLSI-TSA), 2011:1
[113]
Hernández-Rodríguez E, Márquez-Herrera A, Zaleta-Alejandre E, et al. Effect of electrode type in the resistive switching behaviour of TiO2 thin films. J Phys D Appl Phys, 2013, 46(4):045103
[114]
Schindler C, Thermadam S C P, Waser R, et al. Bipolar and unipolar resistive switching in Cu-doped SiO2. IEEE Trans Electron Devices, 2007, 54(10):2762
[115]
Saura X, Lian X, Jiménez D, et al. Field-effect control of breakdown paths in HfO2 based MIM structures. Microelectron Reliab, 2013, 53(9-11):1346
Fig. 1.  Typical construction of memristive devices.

(a) Titanium oxide memristive device with TiO2 bulk layer. (b) Tantalum oxide memristive device with TaO2 bulk layer and Ta2O5 insulator layer.


Fig. 2.  (Color online) Switching mechanism of TaO2 devices.

(a) A negative bias is applied on the top electrode to switch the device ON,repelling negatively charged oxygen ions away from the Ta2O5 layer towards the TaO2 layer. (b) The Ta2O5 phase is reduced to TaO and Ta. After low-resistance TaO and Ta phases align,it forms conducting channels in the Ta2O5 layer. (c) Positive bias is applied on the top electrode to switch the device OFF,attracting oxygen ions towards the Ta2O5 layer. (d) The oxygen ions oxidize TaO and Ta phases to become Ta2O5 phases,and cause the collapse of conducting channels.


Fig. 3.  (Color online) Switching mechanism of TaO2 devices.

(a) A negative bias is applied on the top electrode to switch the device ON, repelling negatively charged oxygen ions away from the Ta2O5 layer towards the TaO2 layer. (b) The Ta2O5 phase is reduced to TaO and Ta. After low-resistance TaO and Ta phases align, it forms conducting channels in the Ta2O5 layer. (c) Positive bias is applied on the top electrode to switch the device OFF, attracting oxygen ions towards the Ta2O5 layer. (d) The oxygen ions oxidize TaO and Ta phases to become Ta2O5 phases, and cause the collapse of conducting channels.


Fig. 4.  Resistance switching of memristive devicesŒ.[87]

Fig. 5.  Temperature-accelerated degradation plot at temperature T.

Fig. 6.  Arrhenius plot to find the rate constant at room temperature.

40.0×103-1 corresponds to a room temperature of 25 °C,while 4.0,5.0 and 6.7 (×10-3) °C-1 corresponds to the temperature-accelerated degradation performed at 250,200,and 150 °C respectively.


Fig. 7.  Ti-O phase diagram[97].

Fig. 8.  Ta-O phase diagram[98].

Table 1.   Gibbs free energy of the formation of metal oxides used in MIM resistive switching devices.

Metal oxideGibbs free energy of formation,Δf G° (kJ/mol)
Hafnium oxide (HfO2)1088:2
Zirconium oxide (ZrO2)1042:8
Titanium oxide (TiO2)888:8
Silicon oxide (SiO2)856:4
Vanadium oxide (VO2)446:4
Nickel oxide (NiO)211:7
Tantalum oxide (TaO2)209:0
Copper oxide (Cu2O)149:0
DownLoad: CSV

Table 2.   Window functions proposed for the behavior of charge carriers approaching electrode boundaries.

Author(s) Window function,f(x)
Strukov et al.[3]f(x)=x(L-w)/L2
Joglekar et al.[71]f(x)=1-(2x-1)2p
Biolek et al.[73]f(x)=1-(x-stp(-i))2p
Prodromakis et al.[74]f(x)=1-[(x-0.5)2+0.75]p
Kvatinsky et al.[76] fon(x)=exp[exp(xaonp)]
foff(x)=exp[exp(xaoffp)]
DownLoad: CSV
[1]
Chua L. Memristor-the missing circuit element. IEEE Trans Circuit Theory, 1971, 18(5):507
[2]
Schroeder H, Jeong D S. Resistive switching in a Pt/TiO2/Pt thin film stack-a candidate for a non-volatile ReRAM. Microelectron Eng, 2007, 84(9-10):1982
[3]
Strukov D B, Snider G S, Stewart D R, et al. The missing memristor found. Nature, 2008, 453(1):80
[4]
Yang J J, Strukov D B, Stewart D R. Memristive devices for computing. Nat Nanotechnol, 2013, 8:13
[5]
Bez R, Pirovano A. Non-volatile memory technologies:emerging concepts and new materials. Mater Sci Semicond Process, 2004, 7(4-6):349
[6]
Duan S, Hu X, Wang L, et al. Memristor-based RRAM with applications. Sci China Inf Sci, 2012, 55(6):1446
[7]
Batas D, Fiedler H. A memristor SPICE implementation and a new approach for magnetic flux-controlled memristor modeling. IEEE Trans Nanotechnol, 2011, 10(2):250
[8]
Jo S H, Chang T, Ebong I, et al. Nanoscale memristor device as synapse in neuromorphic systems. Nano Lett, 2010, 10(4):1297
[9]
Vontobel P O, Robinett W, Kuekes P J, et al. Writing to and reading from a nano-scale crossbar memory based on memristors. Nanotechnology, 2009, 20(42):1
[10]
Eshraghian K, Cho K, Kavehei O, et al. Memristor MOS content addressable memory (MCAM):hybrid architecture for future high performance search engines. IEEE Trans VLSI, 2011, 19(8):1407
[11]
Yoshida C, Tsunoda K, Noshiro H, et al. High speed resistive switching in Pt/TiO2/TiN film for nonvolatile memory application. Appl Phys Lett, 2007, 91(22):223510
[12]
Yang J J, Miao F, Pickett M D, et al. The mechanism of electroforming of metal oxide memristive switches. Nanotechnology, 2009, 20(21):215201
[13]
Boboila S, Desnoyers P. Write endurance in flash drives:measurements and analysis. USENIX Conference on File and Storage Technologies, 2010:115
[14]
Rajachidambaram J S, Murali S, Conley J F, et al. Bipolar resistive switching in an amorphous zinc tin oxide memristive device. J Vac Sci Technol B, 2013, 31(1):01A104
[15]
Driscoll T, Kim H T, Chae B G, et al. Phase-transition driven memristive system. Appl Phys Lett, 2009, 95(4):043503
[16]
Yao J, Sun Z, Zhong L, et al. Resistive switches and memories from silicon oxide. Nano Lett, 2010, 10(10):4105
[17]
Chen A, Haddad S, Wu Y J, et al. Non-volatile resistive switching for advanced memory applications. IEEE International Electron Devices Meeting, 2005:746
[18]
Zhang X. Resistive switching characteristics of Ni/HfO2/Pt ReRAM. J Semicond, 2012, 33(5):054011
[19]
Sakamoto T, Banno N, Iguchi N, et al. A Ta2O5 solid-electrolyte switch with improved reliability. IEEE Symposium on VLSI Technology, 2007:38
[20]
Prakash A, Jana D, Maikap S. TaO_x-based resistive switching memories:prospective and challenges. Nanoscale Res Lett, 2013, 8(1):418
[21]
Yang J J, Zhang M X, Strachan J P, et al. High switching endurance in TaOX memristive devices. Appl Phys Lett, 2010, 97(23):232102
[22]
Akinaga H, Shima H. Resistive random access memory (ReRAM):a metal oxide memory cell. Proc IEEE, 2010:98(12):2237
[23]
Fujisaki Y. Overview of emerging semiconductor non-volatile memories. IEICE Electron Express, 2012, 9(10):908
[24]
Mazumder P, Kang S M, Waser R. Memristors:devices, models, and applications. Proc IEEE, 2012, 100(6):1911
[25]
Ho P W C, Hatem F O, Almurib H A F, et al. Comparison on TiO2 and TaO2 based bipolar resistive switching devices. IEEE International Conference on Electronic Design (ICED), 2014:249
[26]
Lee M J, Lee C B, Lee D, et al. A fast, high-endurance and scalable non-volatile memory device made from asymmetric Ta2O5-x/TaO2-x bilayer structures. Nat Mater, 2011, 10(8):625
[27]
Yoo H K, Kang B S, Lee S B. Forming time of conducting channels in double-layer Pt/Ta2O5/TaOx/Pt and single-layer Pt/TaOx/Pt resistance memories. Thin Solid Films, 2013, 540:190
[28]
Yoo H K, Lee S B, Lee J S, et al. Conversion from unipolar to bipolar resistance switching by inserting Ta2O5 layer in Pt/TaOX/Pt cells. Appl Phys Lett, 2011, 98(18):183507
[29]
Lee S B, Yoo H K, Kim K, et al. Forming mechanism of the bipolar resistance switching in double-layer memristive nanodevices. Nanotechnology, 2012, 23(31):315202
[30]
Wei Z, Kanzawa Y, Arita K, et al. Highly reliable TaOx ReRAM and direct evidence of redox reaction mechanism. IEEE International Electron Devices Meeting, 2008:1
[31]
Fukuda N, Fukuju K, Yogosawa I, et al. Forming-free resistance random access memory using Ta2O5/TaOx bi-layer prepared by magnetron sputtering method. MRS Proc, 2012, 1430
[32]
Pagnia H, Sotnik N. Bistable switching in electroformed metal-insulator-metal devices. Phys Status Solidi, 1988, 108(1):11
[33]
Collins R A, Bowman G, Sutherland R R. Electroforming of MIM thin-film devices in air at atmospheric pressure. J Phys D Appl Phys, 1971, 4(11):49
[34]
Dearnaley G, Stoneham A M, Morgan D V. Electrical phenomena in amorphous oxide films. Reports Prog Phys, 1970, 33(3):1129
[35]
Thurstans R E, Oxley D P. The electroformed metal-insulator-metal structure:a comprehensive model. J Phys D Appl Phys, 2002, 35(8):802
[36]
Strachan J P, Strukov D B, Borghetti J, et al. The switching location of a bipolar memristor:chemical, thermal and structural mapping. Nanotechnology, 2011, 22(25):254015
[37]
Münstermann R, Yang J J, Strachan J P, et al. Morphological and electrical changes in TiO2 memristive devices induced by electroforming and switching. Phys Status Solidi-Rapid Res Lett, 2010, 4(1/2):16
[38]
Dean J A. Lange's handbook of chemistry. 15th ed. New York, USA:McGraw-Hill, Inc, 1999
[39]
Lv H, Wang M, Wan H, et al. Endurance enhancement of Cu-oxide based resistive switching memory with Al top electrode. Appl Phys Lett, 2009, 94(21):213502
[40]
Nauenheim C, Kuegeler C, Ruediger A, et al. Investigation of the electroforming process in resistively switching TiO2 nanocrosspoint junctions. Appl Phys Lett, 2010, 96(12):122902
[41]
Kwon D H, Kim K M, Jang J H, et al. Atomic structure of conducting nanofilaments in TiO2 resistive switching memory. Nat Nanotechnol, 2010, 5:148
[42]
Miao F, Strachan J P, Yang J J, et al. Anatomy of a nanoscale conduction channel reveals the mechanism of a high-performance memristor. Adv Mater, 2011, 23(47):5633
[43]
Hosotani K, Park S G, Nishi Y. Electric field dependent switching and degradation of resistance random access memory. IEEE International Integrated Reliability Workshop Final Report, 2009:11
[44]
Dearnaley G, Morgan D V, Stoneham A M. A model for filament growth and switching in amorphous oxide films. J Non Cryst Solids, 1970, 4:593
[45]
Kim K M, Choi B J, Hwang C S. Localized switching mechanism in resistive switching of atomic-layer-deposited TiO2 thin films. Appl Phys Lett, 2007, 90(24):242906
[46]
Sire C, Blonkowski S, Gordon M J, et al. Statistics of electrical breakdown field in HfO2 and SiO2 films from millimeter to nanometer length scales. Appl Phys Lett, 2007, 91(24):242905
[47]
Murali S, Rajachidambaram J S, Han S Y, et al. Resistive switching in zinc-tin-oxide. Solid State Electron, 2013, 79:248
[48]
Yang M K, Park J W, Ko T K, et al. Resistive switching characteristics of TiN/MnO2/Pt memory devices. Phys Status Solidi-Rapid Res Lett, 2010, 4(8/9):233
[49]
Lu Y M, Noman M, Chen W, et al. Elimination of high transient currents and electrode damage during electroformation of TiO2-based resistive switching devices. J Phys D Appl Phys, 2012, 45(39):395101
[50]
Salaoru I, Prodromakis T, Khiat A, et al. Resistive switching of oxygen enhanced TiO2 thin-film devices. Appl Phys Lett, 2013, 102(1):013506
[51]
Huang J J, Kuo C W, Chang W C, et al. Transition of stable rectification to resistive-switching in Ti/TiO2/Pt oxide diode. Appl Phys Lett, 2010, 96(26):262901
[52]
Hudec B, Hranai M, Hušeková K, et al. Resistive switching in RuO2/TiO2/RuO2 MIM structures for non-volatile memory application. International Conference on Advanced Semiconductor Devices & Microsystems, 2010:255
[53]
Choi B J, Jeong D S, Kim S K, et al. Resistive switching mechanism of TiO2 thin films grown by atomic-layer deposition. J Appl Phys, 2005, 98(3):033715
[54]
Do Y H, Kwak J S, Hong J P, et al. Al electrode dependent transition to bipolar resistive switching characteristics in pure TiO2 films. J Appl Phys, 2008, 104(11):114512
[55]
Strachan J P, Pickett M D, Yang J J, et al. Direct identification of the conducting channels in a functioning memristive device. Adv Mater, 2010, 22(32):3573
[56]
Yang J J, Pickett M D, Li X, et al. Memristive switching mechanism for metal/oxide/metal nanodevices. Nat Nanotechnol, 2008, 3(7):429
[57]
Sawa A. Resistive switching in transition metal oxides. Mater today, 2008, 11(6):28
[58]
Yasuhara R, Fujiwara K, Horiba K, et al. Inhomogeneous chemical states in resistance-switching devices with a planar-type Pt/CuO/Pt structure. Appl Phys Lett, 2009, 95(1):012110
[59]
Sun P, Su L, Ling L, et al. Simulation study of conductive filament growth dynamics in oxide-electrolyte-based ReRAM. J Semicond, 2014, 35(10):104007
[60]
Park G S, Kim Y B, Park S Y, et al. In situ observation of filamentary conducting channels in an asymmetric Ta2O5-x/TaO2-x bilayer structure. Nat Commun, 2013, 4:2382
[61]
Strachan J P, Medeiros-Ribeiro G, Yang J J, et al. Spectromicroscopy of tantalum oxide memristors. Appl Phys Lett, 2011, 98(24):242114
[62]
Bi J, Han Z. Characteristics of HfO2/Hf-based bipolar resistive memories. J Semicond, 2015, 36(6):064010
[63]
Kim K M, Choi B J, Shin Y C, et al. Anode-interface localized filamentary mechanism in resistive switching of TiO2 thin films. Appl Phys Lett, 2007, 91(1):012907
[64]
Lin C Y, Wu C Y, Wu C Y, et al. Modified resistive switching behavior of ZrO2 memory films based on the interface layer formed by using Ti top electrode. J Appl Phys, 2007, 102(9):094101
[65]
Prakash A, Maikap S, Chiu H C, et al. Enhanced resistive switching memory characteristics and mechanism using a Ti nanolayer at the W/TaOX interface. Nanoscale Res Lett, 2014, 9(1):125
[66]
Zhuo V Y, Jiang Y, Zhao R, et al. Improved switching uniformity and low-voltage operation in TaO-based RRAM using Ge reactive layer. IEEE Electron Device Lett, 2013,34(9):1130
[67]
Seong D, Jo M, Lee D, et al. HPHA effect on reversible resistive switching of Pt/Nb-doped SrTiO3 Schottky junction for nonvolatile memory application. Electrochem Solid-State Lett, 2007, 10(6):168
[68]
Hatem F O, Ho P W C, Kumar T N, et al. Modeling of bipolar resistive switching of a nonlinear MISM memristor. Semicond Sci Technol, 2015, 30(11):115009
[69]
Kim S, Choi S, Lu W. Comprehensive physical model of dynamic resistive switching in an oxide memristor. ACS Nano, 2014
[70]
Mattsson M S, Niklasson G A, Forsgren K, et al. A frequency response and transient current study of ��-Ta2O5:methods of estimating the dielectric constant, direct current conductivity, and ion mobility. J Appl Phys, 1999, 85(4):2185
[71]
Joglekar Y N, Wolf S J. The elusive memristor:properties of basic electrical circuits. Eur J Phys, 2009, 30(4):661
[72]
Strukov D B, Williams R S. Exponential ionic drift:fast switching and low volatility of thin-film memristors. Appl Phys A, 2009, 94(3):515
[73]
Biolek Z, Biolek D, Biolková V. SPICE model of memristor with nonlinear dopant drift. Radio Engineering, 2009, 18(2):210
[74]
Prodromakis T, Peh B P, Papavassiliou C, et al. A versatile memristor model with non-linear dopant kinetics. IEEE Trans Electron Devices, 2011, 58(9):3099
[75]
Rossel C, Meijer G I, Breìmaud D, et al. Electrical current distribution across a metal-insulator-metal structure during bistable switching. J Appl Phys, 2001, 90(6):2892
[76]
Kvatinsky S, Friedman E G, Kolodny A, et al. TEAM:threshold adaptive memristor model. IEEE Trans Circuits Syst I, 2013, 60(1):211
[77]
Zhang L, Chen Z, Yang J J, et al. A compact modeling of TiO2-TiO2-x memristor. Appl Phys Lett, 2013, 102(15):153503
[78]
Sze S M, Ng K K. Physics of semiconductor devices. 3rd ed. New Jersey, USA:John Wiley & Sons, 2007
[79]
Ji H H, Lee M J, Chang L B, et al. Modeling for bipolar resistive memory switching in transition-metal oxides. Phys Rev B, 2010, 82(15):155321
[80]
Pickett M D, Strukov D B, Borghetti J L, et al. Switching dynamics in titanium dioxide memristive devices. J Appl Phys, 2009, 106(7):074508
[81]
Zhang F, Gan X, Li X, et al. Realization of rectifying and resistive switching behaviors of TiO2 nanorod arrays for nonvolatile memory. Electrochem Solid-State Lett, 2011, 14(10):H422
[82]
Acharyya D, Hazra A, Bhattacharyya P. A journey towards reliability improvement of TiO2 based resistive random access memory:a review. Microelectron Reliab, 2014, 54(3):541
[83]
Franke E, Trimble C L, DeVries M J, et al. Dielectric function of amorphous tantalum oxide from the far infrared to the deep ultraviolet spectral region measured by spectroscopic ellipsometry. J Appl Phys, 2000, 88(9):5166
[84]
Kukli K, Aarik J, Aidla A, et al. Properties of tantalum oxide thin films grown by atomic layer deposition. Thin Solid Films, 1995, 260(2):135
[85]
Fleming R M, Lang D V, Jones C D W, et al. Defect dominated charge transport in amorphous Ta2O5 thin films. J Appl Phys, 2000, 88(2):850
[86]
Cao X, Li X, Yu W, et al. Bipolar resistive switching properties of microcrystalline TiO2 thin films deposited by pulsed laser deposition. Mater Sci Eng B, 2009, 157(1-3):36
[87]
Blaschke D, Zahn P, Skorupa I, et al. Resistive switching in thermally oxidized titanium films. International Semiconductor Conference Dresden-Grenoble (ISCDG), 2013:1
[88]
Lee S R, Kim Y, Chang M, et al. Multi-level switching of triple-layered TaO_x RRAM with excellent reliability for storage class memory. Symposium on VLSI Technology, 2012:71
[89]
Hsiung C P, Gan J Y, Tseng S H, et al. Resistance switching characteristics of TiO2 thin films prepared with reactive sputtering. Electrochem. Solid-State Lett, 2009, 12(7):G31
[90]
Niset M, Kuhn P. Typical data retention for nonvolatile memory. Freescale Semiconductor Engineering Bulletin, 2005:EB618/D
[91]
Jo S H, Kim K H, Lu W. Programmable resistance switching in nanoscale two-terminal devices. nano Lett, 2009, 9:496
[92]
Chen A, Haddad S, Wu Y C. A temperature-accelerated method to evaluate data retention of resistive switching nonvolatile memory. IEEE Electron Device Lett, 2008, 29(1):38
[93]
Young H Do, June S K, Yoon C B, et al. Hysteretic bipolar resistive switching characteristics in TiO2-TiO2-x multilayer homojunctions. Appl Phys Lett, 2009, 95(9):093507
[94]
Jung S, Kong J, Song S, et al. Resistive switching characteristics of solution-processed transparent TiO_x for nonvolatile memory application. J Electrochem Soc, 2010, 157(11):1042
[95]
Chang W Y, Cheng K J, Tsai J M, et al. Improvement of resistive switching characteristics in TiO2 thin films with embedded Pt nanocrystals. Appl Phys Lett, 2009, 95(4):042104
[96]
Xia Q, Pickett M D, Yang J J, et al. Impact of geometry on the performance of memristive nanodevices. Nanotechnology, 2011, 22(25):254026
[97]
Okamoto H. O-Ti (oxygen-titanium). J Phase Equilibria Diffus, 2011, 32(5):473
[98]
Garg S P, Krishnamurthy N, Awasthi A, et al. The O-Ta (oxygen-Tantalum) system. J Phase Equilibria, 1996, 17(1):63
[99]
Kihlborg L, Olovsson I. Arne Magnéli, 1914-1996. Acta Cryst A, 1997, 53(2):103
[100]
Yang J J, Strachan J P, Xia Q, et al. Diffusion of adhesion layer metals controls nanoscale memristive switching. Adv Mater, 2010, 22(36):4034
[101]
Park J, Jung S, Lee J, et al. Resistive switching characteristics of ultra-thin TiO_x. Microelectron Eng, 2011, 88(7):1136
[102]
Lee T W, Nickel J H. Memristor resistance modulation for analog applications. IEEE Electron Device Lett, 2012, 33(10):1456
[103]
Tedesco J L, Stephey L, Hernández-Mora M, et al. Switching mechanisms in flexible solution-processed TiO2 memristors. Nanotechnology, 2012, 23(30):305206
[104]
Lv H B, Yin M, Fu X F, et al. Resistive memory switching of CuXO films for a nonvolatile memory application. IEEE Electron Device Lett, 2008, 29(4):309
[105]
Gu T. Role of oxygen vacancies in TiO2-based resistive switches. J Appl Phys, 2013, 113(3):033707
[106]
Ko D S, Kim S I, Ahn T Y, et al. Effect of the electrode materials on the resistive switching of Ti4O7. Appl Phys Lett, 2012, 101(5):053502
[107]
Lin K L, Hou T H, Shieh J, et al. Electrode dependence of filament formation in HfO2 resistive-switching memory. J Appl Phys, 2011, 109(8):084104
[108]
Jeong H Y, Kim Y I, Lee J Y, et al. Choi, A low-temperature-grown TiO2-based device for the flexible stacked RRAM application. Nanotechnology, 2010, 21(11):115203
[109]
Biju K P, Liu X, Bourim E M, et al. Asymmetric bipolar resistive switching in solution-processed Pt/TiO2/W devices. J Phys D Appl Phys, 2010, 43(49):495104
[110]
Chen C, Gao S, Zeng F, et al. Migration of interfacial oxygen ions modulated resistive switching in oxide-based memory devices. J Appl Phys, 2013, 114(1):014502
[111]
Yoon M J, Lee S B, Yoo H K, et al. Dielectric-breakdown-like forming process in the unipolar resistance switching of Ta2O5-X thin films. Curr Appl Phys, 2012, 1293:846
[112]
Prakash A, Maikap S, Chen G, et al. Improvement of high-�� Ta2O5-based resistive switching memory using Ti interfacial layer. International Symposium on VLSI Technology, Systems and Applications (VLSI-TSA), 2011:1
[113]
Hernández-Rodríguez E, Márquez-Herrera A, Zaleta-Alejandre E, et al. Effect of electrode type in the resistive switching behaviour of TiO2 thin films. J Phys D Appl Phys, 2013, 46(4):045103
[114]
Schindler C, Thermadam S C P, Waser R, et al. Bipolar and unipolar resistive switching in Cu-doped SiO2. IEEE Trans Electron Devices, 2007, 54(10):2762
[115]
Saura X, Lian X, Jiménez D, et al. Field-effect control of breakdown paths in HfO2 based MIM structures. Microelectron Reliab, 2013, 53(9-11):1346
1

Quantum light source devices of In(Ga)As semiconductor self-assembled quantum dots

Xiaowu He, Yifeng Song, Ying Yu, Ben Ma, Zesheng Chen, et al.

Journal of Semiconductors, 2019, 40(7): 071902. doi: 10.1088/1674-4926/40/7/071902

2

Recent advances in flexible and wearable organic optoelectronic devices

Hong Zhu, Yang Shen, Yanqing Li, Jianxin Tang

Journal of Semiconductors, 2018, 39(1): 011011. doi: 10.1088/1674-4926/39/1/011011

3

Flexible devices: from materials, architectures to applications

Mingzhi Zou, Yue Ma, Xin Yuan, Yi Hu, Jie Liu, et al.

Journal of Semiconductors, 2018, 39(1): 011010. doi: 10.1088/1674-4926/39/1/011010

4

Flexible magnetic thin films and devices

Ping Sheng, Baomin Wang, Runwei Li

Journal of Semiconductors, 2018, 39(1): 011006. doi: 10.1088/1674-4926/39/1/011006

5

Preface to the Special Topic on Devices and Circuits for Wearable and IoT Systems

Zhihua Wang, Yong Hei, Zhangming Zhu

Journal of Semiconductors, 2017, 38(10): 101001. doi: 10.1088/1674-4926/38/10/101001

6

Research progress of Si-based germanium materials and devices

Buwen Cheng, Cheng Li, Zhi Liu, Chunlai Xue

Journal of Semiconductors, 2016, 37(8): 081001. doi: 10.1088/1674-4926/37/8/081001

7

Surface acoustic wave devices for sensor applications

Bo Liu, Xiao Chen, Hualin Cai, Mohammad Ali Mohammad, Xiangguang Tian, et al.

Journal of Semiconductors, 2016, 37(2): 021001. doi: 10.1088/1674-4926/37/2/021001

8

Packaging investigation of optoelectronic devices

Zhike Zhang, Yu Liu, Jianguo Liu, Ninghua Zhu

Journal of Semiconductors, 2015, 36(10): 101001. doi: 10.1088/1674-4926/36/10/101001

9

A 2DEG charge density based drain current model for various Al and In molefraction mobility dependent nano-scale AlInGaN/AlN/GaN HEMT devices

Godwin Raj, Hemant Pardeshi, Sudhansu Kumar Pati, N Mohankumar, Chandan Kumar Sarkar, et al.

Journal of Semiconductors, 2013, 34(4): 044002. doi: 10.1088/1674-4926/34/4/044002

10

Process optimization of a deep trench isolation structure for high voltage SOI devices

Zhu Kuiying, Qian Qinsong, Zhu Jing, Sun Weifeng

Journal of Semiconductors, 2010, 31(12): 124009. doi: 10.1088/1674-4926/31/12/124009

11

CuPc/C60 heterojunction thin film optoelectronic devices

Imran Murtaza, Ibrahim Qazi, Khasan S. Karimov

Journal of Semiconductors, 2010, 31(6): 064005. doi: 10.1088/1674-4926/31/6/064005

12

Energy Transfer Probability Between Host and Guest in Doped Organic Electrophosphorescent Devices

Li Hongjian, Ouyang Jun, Dai Guozhang, Dai Xiaoyu, Pan Yanzhi, et al.

Chinese Journal of Semiconductors , 2006, 27(4): 674-678.

13

An All-E-Beam Lithography Process for the Patterningof 2D Photonic Crystal Waveguide Devices

Yu Hejun, Yu Jinzhong, Chen Shaowu

Chinese Journal of Semiconductors , 2006, 27(11): 1894-1899.

14

Modified Reynolds Equation for Squeeze-Film Air Damping of Slotted Plates in MEMS Devices

Sun Yuancheng, Bao Minhang, Yang Heng, Huang Yiping

Chinese Journal of Semiconductors , 2006, 27(3): 473-477.

15

An Analytical Model for the Surface Electrical Field Distribution and Optimization of Bulk-Silicon Double RESURF Devices

Li Qi, Li Zhaoji, Zhang Bo

Chinese Journal of Semiconductors , 2006, 27(7): 1177-1182.

16

Electron Injection Enhancement by Diamond-Like Carbon Film in Polymer Electroluminescence Devices

Li Hongjian, Yan Lingling, Huang Baiyun, Yi Danqing, Hu Jin, et al.

Chinese Journal of Semiconductors , 2006, 27(1): 30-34.

17

Status and Trends in Advanced SOI Devices and Materials

Balestra Francis

Chinese Journal of Semiconductors , 2006, 27(4): 573-582.

18

Status and Trends in Advanced SOI Devices and Materials

Wang Tianxing, Wei Hongxiang, Ren Cong, Han Xiufeng, Clifford E, et al.

Chinese Journal of Semiconductors , 2006, 27(4): 591-597.

19

Terahertz Semiconductor Quantum Well Devices

Liu H C, Luo H, Ban D, Wachter M, Song C Y, et al.

Chinese Journal of Semiconductors , 2006, 27(4): 627-634.

20

W/TiN Gate Thin-Film Fully-Depleted SOI CMOS Devices

Chinese Journal of Semiconductors , 2005, 26(1): 6-10.

  • Search

    Advanced Search >>

    GET CITATION

    Patrick W. C. Ho, Firas Odai Hatem, Haider Abbas F. Almurib, T. Nandha Kumar. Comparison between Pt/TiO2/Pt and Pt/TaOX/TaOY/Pt based bipolar resistive switching devices[J]. Journal of Semiconductors, 2016, 37(6): 064001. doi: 10.1088/1674-4926/37/6/064001
    P. W. C. Ho, F O Hatem, H. A. F. Almurib, T. N. Kumar. Comparison between Pt/TiO2/Pt and Pt/TaOX/TaOY/Pt based bipolar resistive switching devices[J]. J. Semicond., 2016, 37(6): 064001. doi:  10.1088/1674-4926/37/6/064001.
    shu

    Export: BibTex EndNote

    Article Metrics

    Article views: 4118 Times PDF downloads: 69 Times Cited by: 0 Times

    History

    Received: 15 September 2015 Revised: 06 December 2015 Online: Published: 01 June 2016

    Catalog

      Email This Article

      User name:
      Email:*请输入正确邮箱
      Code:*验证码错误
      Patrick W. C. Ho, Firas Odai Hatem, Haider Abbas F. Almurib, T. Nandha Kumar. Comparison between Pt/TiO2/Pt and Pt/TaOX/TaOY/Pt based bipolar resistive switching devices[J]. Journal of Semiconductors, 2016, 37(6): 064001. doi: 10.1088/1674-4926/37/6/064001 ****P. W. C. Ho, F O Hatem, H. A. F. Almurib, T. N. Kumar. Comparison between Pt/TiO2/Pt and Pt/TaOX/TaOY/Pt based bipolar resistive switching devices[J]. J. Semicond., 2016, 37(6): 064001. doi:  10.1088/1674-4926/37/6/064001.
      Citation:
      Patrick W. C. Ho, Firas Odai Hatem, Haider Abbas F. Almurib, T. Nandha Kumar. Comparison between Pt/TiO2/Pt and Pt/TaOX/TaOY/Pt based bipolar resistive switching devices[J]. Journal of Semiconductors, 2016, 37(6): 064001. doi: 10.1088/1674-4926/37/6/064001 ****
      P. W. C. Ho, F O Hatem, H. A. F. Almurib, T. N. Kumar. Comparison between Pt/TiO2/Pt and Pt/TaOX/TaOY/Pt based bipolar resistive switching devices[J]. J. Semicond., 2016, 37(6): 064001. doi:  10.1088/1674-4926/37/6/064001.

      Comparison between Pt/TiO2/Pt and Pt/TaOX/TaOY/Pt based bipolar resistive switching devices

      DOI: 10.1088/1674-4926/37/6/064001
      More Information
      • Corresponding author: Patrick W. C. Ho Email: kecx2pha@nottingham.edu.my
      • Received Date: 2015-09-15
      • Revised Date: 2015-12-06
      • Published Date: 2016-06-01

      Catalog

        /

        DownLoad:  Full-Size Img  PowerPoint
        Return
        Return