Ch13-Yu

advertisement
Metal Oxide Resistive Switching Memory
Shimeng Yu, Byoungil Lee, H.-S. Philip Wong
Center for Integrated Systems and Department of Electrical Engineering, Stanford University
13.1 Introduction
Information storage device is a key component of nanoelectronic
systems. Conventional memories such as SRAM, DRAM, and
FLASH are facing formidable device scaling challenges. Emerging
memory device concepts have been actively pursued both in industry
and academia in hopes of finding solutions for future information
storage needs [1-2]. The ideal characteristics for a memory device
include fast programming speed (~ns), long retention time (>10
years), low power consumption, good reliability, high integrated
density, and continued scalability. Several candidates have been proposed to achieve the above goals, such as magnetoresistive random
access memory (MRAM) [3], ferroelectric random access memory
(FeRAM) [4]. In recent years, memory devices based on the electrically switchable resistance phenomenon have been extensively studied. The basic principle of this kind of memory is to use a high re-
2
sistance state (HRS) or low resistance state (LRS) to store the information data “0” or “1”, and the transition between the two states can
be triggered by electrical inputs. Roughly speaking, resistive switching memories can be classified into three groups [5]: (i) phasechange memory (PCM) based on chalcogenides [6-9], which relies
on the temperature induced change between the crystalline phase
(corresponding to LRS) and the amorphous phase (corresponding to
HRS) (ii) programmable-metallization-cell (PMC) memory based on
solid electrolytes or polymers [10-13], which relies on the formation
(corresponding to LRS) or the rupture (corresponding to HRS) of a
metallic conducting bridge between two electrodes, and (iii) resistance random access memory (RRAM) based on metal oxides.
Among these resistive switching memories, metal oxide memory is
promising for practical applications due to its compatibility with silicon CMOS fabrication technology. Here, we review the research
progress of metal oxide memory including the physical mechanism
of switching, state of the art device performances, as well as device
cell structure for integration into a memory array.
The negative differential resistances phenomenon in oxides was
firstly reported in the 1960s [14-15], then it was reviewed by G.
Dearnaley et al. [16] in 1970. Recent work on the resistive switching
metal oxide memory can be traced back to the discovery of hysteresis I-V characteristics in perovskite oxides [17-19] such as
Pr0.7Ca0.3MnO3, SrTiO3, SrZrO3, etc. in the late 1990s and the early
2000s. Since 2004, the research activities have focused on binary
metal oxides [20-26] such as NiO, TiO2, ZrO2, ZnO, Cu2O, Al2O3,
3
HfO2, etc. because of the simplicity of the material and good compatibility with silicon CMOS fabrication process. Recently, the performance of perovskite oxide memories and related physical mechanism of resistive switching were reviewed by R. Waser et al. [27-28]
and A. Sawa [29]. Therefore, here we focus this review on binary
metal oxide memory.
13.1.1 Device Operation
It is necessary to first introduce some basic concepts and terminologies about metal oxide memory. The typical metal oxide memory
cell is a simple metal-insulator-metal (MIM) structure, as shown in
Fig. 1 (a). The switching event from high resistance state (HRS) to
low resistance state (LRS) is referred to as the “set” process. Conversely, the switching event from LRS to HRS is referred to as the
“reset” process. In some cases, for the fresh samples in its initial resistance state (IRS), a larger voltage is needed to trigger on the resistive switching behaviors for the subsequent cycles. This is called the
“electroforming” or “forming” process. The switching modes of
metal oxide memory can be broadly classified into two switching
modes: unipolar and bipolar. Unipolar switching means the switching direction depends on the amplitude of the applied voltage but not
on the polarity, thus set/reset can occur at the same polarity, and
usually it can symmetrically occur at both forward and reversed
voltages, as illustrated in Fig. 1(b). Bipolar switching means the
switching direction depends on the polarity of the applied voltage,
thus set can only occur at one polarity and reset can only occur at the
4
reverse polarity, as illustrated in Fig. 1 (c). To avoid a hard dielectric
breakdown in the set process, it is recommended to enforce a set
compliance current, which is usually provided by the semiconductor
parameter analyzer, or more practically, by a memory cell selection
(or access) transistor/diode or a series resistor. To read the data from
the cell, a small voltage is applied that does not affect the state of the
memory cell. For non-volatile application, the cell should retain its
state at standby mode and free from disturb from reading/writing of
neighboring cells.
Fig. 1 (a) Schematic of metal-insulator-metal (MIM) structure for metal oxide
memory cell, and schematic of metal oxide memory’s I-V curves, showing two
modes of operation: (b) unipolar and (c) bipolar
13.1.2 Device Characteristics
To further understand how the metal oxide memory works, the device characteristics of HfO2 based memory cell [30-31] from Industrial Technology Research Institute (ITRI) are presented as an example. Fig. 2 (a) shows the transmission electron microscopy (TEM)
image of the concave structure device of TiN/Ti/HfOx/TiN stack
with 30 nm cell size; here 5 nm HfOx layer is used for the main
switching layer, TiN is used for the electrode material, and a thin Ti
buffer layer is used for improving the switching stability. Fig. 2 (b)
shows the typical I-V curve of such memory cell. A 200 µA set
compliance current is enforced, and the device exhibits bipolar
switching. Fig. 2 (c) shows the switching endurance testing result.
The set/reset programming condition is +1.5 V/-1.4 V pulse with
5
500 µs width, and it is seen that after 106 switching cycles, the
HRS/LRS resistance ratio is still larger than 100, although there is
some degradation. Fig. 2 (d) shows the data retention testing result,
the measurement is performed at 150 °C, and it is seen that 10 years
lifetime is expected using a simple linear extrapolation. Besides, fast
switching speed (~5 ns), and multi-bit storage potential is also
demonstrated in the HfO2 memory.
Besides HfO2, other binary metal oxides such as NiO, TiO2, ZrO2,
ZnO, Cu2O, Al2O3 have been extensively explored. Up to now, dozens of binary metal oxides have been found to exhibits electric-fieldinduced bistable resistance switching behavior. Most of the metals
are transition metals, and some are lanthanide series metals. The materials for switching layer and electrode layer are summarized in Table 13.1. Note that some nitrides, e.g. TiN are also used for the electrode layer.
Fig. 2 (a) Transmission electron microscopy (TEM) image of the concave structure device of TiN/Ti/HfOx/TiN stack; (b) Typical I-V curve of the device with the
cell size of 30 nm; (c) Endurance testing by 500 µs pulse: successive 10 6 switching
cycles is achieved; (d) Retention testing at 150 °C: 10 years lifetime is expected.
Reprinted from Ref [30-31]
Table 13.1 Summary of the materials that has been used for binary metal oxide
memory. Yellow metals corresponding binary oxides are used for switching layer;
blue metals are used for electrode layer.
13.2 Possible Physical Mechanism for Resistive Switch-
6
ing in Metal Oxides
The physical mechanism for resistive switching in metal oxide
memory is still under a heated debate. According to ITRS 2007 [32],
resistive switching mechanism can be roughly classified into three
groups: thermal effect, electronic effect, or ionic effect. Thermal effect refers to the thermal dissolution of conductive filaments triggered by local Joule-heating and works like a traditional household
fuse albeit at the nanoscale [33-34]. It is also referred to as fuse/antifuse switching type. The electronic effect mechanism conjectures
that injected charges are trapped by interface defects. These trapped
charges modify the Schottky barrier height between electrodes and
oxides and thereby changing the conductance through the MIM
structure [35-36]. Another case of electronic effect is Mott metalinsulator-transition due to the strong electron-electron correlation in
some transition metal oxides [37-38]. Ionic effect refers to the migration of ions with related electrochemical reactions to form a conducting bridge between electrodes [39-40]. It is also referred to as
reduction/oxidation type. Actually, this ionic effect is just the operating principle of programmable-metallization-cell memory in solid
electrolytes or polymers as mentioned before. Metal oxides can also
serve as fast ions conductors for some particular metal ions like Ag+
and Cu2+ [41-42]. Here we plan to focus on the intrinsic properties
of metal oxides, thus we would not include those metal oxides
memory with Ag or Cu electrodes in the following discussions.
So far, no single model mentioned above can explain all the ex-
7
perimental results obtained in various materials. It seems that the
thermal dissolution model can address parts of the unipolar switching characteristics, while the ionic migration model can explain most
of the bipolar switching characteristics. However, the distinction between the two switching behaviors is ambiguous. Here we discuss
several key issues such as the conducting mechanism, the nature of
electroforming/set/reset process, the effect of electrode materials on
switching modes, with the aim of seeking a rational mechanism for
both unipolar and bipolar switching characteristics.
13.2.1 Conduction Mechanism
A lot of conductive atomic force microscopy (C-AFM) measurements [43-47] have been reported in the literature, revealing that
conductive filaments (CFs) with nanoscale diameters are formed in
the oxide layer during the set process, and large conducting current
passes through these filaments in LRS. Fig. 3 shows one of these CAFM measurements results [45]. Typically, CFs are sparsely and
non-homogeneously distributed under the electrodes. This means
that the conducting area in LRS is an extremely small portion of the
entire electrode area, which is typically quite large in most experiments. Therefore, the conductance in LRS would not decrease as
much as the conductance in HRS does when the electrode area is decreased, resulting a larger HRS/LRS resistance ratio, which is a benefit of scaling down memory cell size. It has been experimentally
demonstrated that these CFs can be formed by C-AFM tip individually [48], but in most cases multiple CFs exist in the oxide layer,
8
which is believed to be responsible for the multi-level switching behaviors exhibited in some metal oxide memory devices [49-50]. Although the filamentary conduction mechanism is widely recognized,
some researchers argued that a bulk conduction mechanism dominates the LRS resistance in their TiO2 devices [51]. They proposed
that charged dopants (mainly oxygen vacancies) migrate under a
voltage bias to the TiO2/electrode interface. The interfacial changes
modify the Schottky barrier height at the interface and lead to the resistive switching in their devices. In essence, such arguments do not
contradict with the filamentary conduction mechanism, provided
that the so-called bulk conduction is regarded as many parallel CFs.
It should be noted that if CFs are dense enough in ultra-scaled devices, the bulk conduction behavior may arise. A usual experiment
that aims at distinguishing filamentary or bulk conduction is to
measure the trend of the HRS/LRS resistance ratio versus the cell
area. If the ratio goes up when the cell area is scaled down, filamentary conduction prevails. If the ratio remains almost constant when
the cell is scaled down, bulk conduction prevails. However, so far,
these experiments have not been performed for devices with truly
nanometer scale size (< 100 nm) memory cells and therefore the results are not yet conclusive.
Fig. 3 Conductive atomic force microscopy (C-AFM) of LRS (a) and HRS (b) conductance in Al2O3 memory, showing filamentary conduction dominates in LRS.
(scanned area: 2 µm×2 µm; maximum y scale: (a) 105 nA, (b) 5 nA). Reprinted
from [45]
9
The CFs are conjectured to be made up of oxygen vacancies (Vo)
in most metal oxide memory devices, and this assumption was confirmed by micro X-ray fluorescence (XRF) and X-ray absorption
near-edge spectroscopy (XANES) [52]. It is well known that Vo can
act as an effective donor in n-type metal oxides. Ab-initio calculation of the rutile TiO2 electronic structure [53] reveals that Vo can
produce a defect state within the band gap. This state is occupied by
two electrons that are localized on Ti 3d orbital of the nearest Ti atoms to Vo. If a chain of such Vo forms, it is expected that electrons
delocalization occurs and the hopping probability can increase remarkably along these CFs and ultimately lead to the LRS. Although
Vo plays an important role in the switching behaviors of most metal
oxide memory devices, the composition of CFs is not limited to Vo.
There are quite a few reports that the CFs in NiO memory are composed of excess metallic Ni in NiO. This suggestion is confirmed by
measurement of the dependence of the LRS resistance on temperature [54] and XPS composition analysis [55]. It was further verified
by electron energy loss spectroscopy (EELS) that there exists a Nirich phase at the grain boundaries, which implies that the CFs of Ni
are formed as precipitates at the grain boundaries [56]. A simple way
to distinguish whether the CFs are metallic or semiconducting is to
measure the LRS resistance temperature dependency. If the LRS resistance goes up with the increase of temperature, the CFs are metallic and may consist of metal precipitates. On the contrary, if the LRS
resistance drops with the increase of temperature, CFs are semiconducting and may consist of Vo.
10
There are many efforts to fit the I-V characteristics of HRS and
LRS to current conduction models in the literature. Most of the reports show an Ohmic relationship in the LRS. And in the semiconducting CFs, Mott variable hopping conduction [57] is proposed to
dominate in LRS, and this assumption was supported by some temperature varying measurements [58-59] and AC conductance measurement [60]. But, the conduction fitting results in HRS are quite diverse, Poole-Frenkel emission (I~V*exp(V1/2)) [61-62], Schottky
emission (I~exp(V1/2)) [63-64], the space charge limited current
(SCLC) characteristic (the Ohmic region I~V, followed by the
child’s square law region I~V2) [65-66], were observed in various
metal oxide memories. The diverse I-V characteristics in HRS involving different leakage mechanism may be associated with the different dielectric properties or different fabrication processes conditions, e.g. annealing temperature, annealing ambient, and the
properties of the interface between the oxides and the electrodes.
Nevertheless, the key issue of metal oxides memory modeling is not
the leakage current mechanism in HRS, but to investigate the mechanism that triggers the resistive switching. So in the next section, we
will discuss the forming/set and reset process, respectively.
13.2.2 Electroforming/Set/Reset Process with Oxygen Migration
Dielectric breakdown is a process of local materials transitioning
from insulating to conductive under a high external electric field.
This phenomenon is well-known for gate dielectrics in MOSFETs
and has been studied for a long time. Recently, by site-specific struc-
11
tural analysis using high resolution transmission electron microscopy (HR-TEM) with EELS, X. Li et al. [67] revealed that after dielectric breakdown, oxygen atoms in gate dielectric SiO2 were missing.
Substoichiometric silicon oxide SiOx (with x<2) was formed, and
the local energy gap was lowered with intermediate bonding state of
silicon atoms Si1+, Si2+, and Si3+ in the percolation leakage path. Also, chemical bond breakage and the local Joule heating due to large
current flowing through the percolation path are believed to be the
main driving forces leading to the oxygen dissociation.
Similarly, in metal oxides, the electroforming/set process is interpreted to be some kind of soft dielectric breakdown [60, 68-69]. J. J.
Yang et al. [70] claimed that they observed oxygen gas bubbles at
the electrode surface of their bipolar TiO2 memory devices in the
electroforming process. They regarded the electroforming process in
metal oxide memory as an electro-reduction and Vo creation process
caused by high electric field, which is then enhanced by local Joule
heating. During electroforming, Vo are created and drift towards the
cathode, forming localized CFs in the oxide. Simultaneously, oxygen
ions drift towards the anode where they discharge to evolve oxygen
gas. M.-J. Lee et al. [71] investigated the composition change in
their unipolar NiO memory devices during the set process by secondary ion mass spectroscopy (SIMS), as shown in Fig. 4. The authors suggested that the oxygen atoms within the NiO layer migrated
toward the Pt electrode after the switching, thus leaving the Ni rich
filaments in the NiO layer.
12
Fig. 4 (a) X-ray photoelectron spectroscopy (XPS) analysis of NiO memory; the
sample deposited at an oxygen partial pressure of 5% shows the coexistence of Ni
(852.8 keV) and NiO (854.3 keV) peaks, which exhibited bistable resistive switching; the sample deposited at an oxygen partial pressure of 30% did not exhibit any
stable resistive switching phenomenon, suggesting CFs may consist of metallic Ni
precipitates. (b) Secondary-ion mass spectroscopy (SIMS) data in NiO memory,
showing the migration of oxygen atoms towards the electrodes after electrical
switching. Reprinted from [71]
Therefore, in both unipolar and bipolar switching metal oxide
memory devices, the electroforming/set process is conjectured to be
associated with the generation of Vo and the migration of the oxygen
atoms, generating CFs consist of either Vo or metal precipitates. In
fresh samples, usually the amount of Vo is small, thus a high voltage
is needed to trigger the electroforming process. After electroforming,
the devices have switched to LRS, and a sufficient amount of Vo is
present. In subsequent switching cycles, only a portion of the Vo, the
ones near the anode, can be recombined during the reset process.
This is why the set voltage would be smaller than forming voltage
and the resistance in HRS would be much smaller than the resistance
in IRS. Through fitting of electrical parameters, the formation and
rupture of the CFs is estimated to occur in a localized region (3–10
nm) thick near the anode [72]. The remaining Vo rich region is referred to as the “virtual cathode” [28].
Obviously, it is not desirable to have a large electroforming voltage
in practical applications. Thus significant efforts have been made to
achieve the so-called “forming-free” devices. The forming voltage is
13
linearly dependent on the thickness of the oxide film [30, 73-75]. So
a thinner oxide film is effective in reducing the forming voltage. H.
Y. Lee et al. [30] claimed that the as-fabricated atomic layer deposition (ALD) HfO2 device is free from the electroforming process as
the thickness of the film is thinned to be 3 nm. Besides, controlling
the annealing ambient during deposition to obtain oxygen deficient
films is also helpful in reducing the forming voltage [76-78].
So far, we have stated that the nature of dielectric breakdown is
the formation of oxygen deficient percolation leakage paths. Note
that gate dielectric breakdown is a permanent and irreversible process, while the switching process in metal oxide memory is a reversible process. So the memory device should have a location to
store the missing oxygen atoms during the set process and then drive
them back during the reset process. The concept of an “oxygen reservoir” is thus envisioned. Usually, the anode electrode materials act
as the oxygen reservoir, and it also prevents the oxygen from escaping from the device to the ambient. The anode electrode material
then provides a source of oxygen for reset process. Fig. 5 illustrates
how the oxygen reservoir works in a Ti/MnO2/Pt memory device
[79]. X-ray photoelectron spectroscopy (XPS) reveals that amount of
non-lattice oxygen became less in LRS than in HRS, and the corresponding shift of Ti 2p peak indicates that an interfacial TiOx layer
was formed. This observation suggests that during the set process,
oxygen ions moved toward the Ti anode when a positive bias was
applied, where they reduced the Ti and formed TiOx. This interfacial
TiOx layer is believed to act as an oxygen reservoir that stores oxy-
14
gen atoms. When a negative bias was applied to the Ti anode, oxygen ions moved away from the interfacial layer toward the oxide
bulk layer and recombine with Vo. Therefore, the bipolar switching
characteristics are assumed to be the formation and rupture of the
CFs consisting of Vo associated with oxygen ion migration.
Fig. 5 XPS spectra of Ti/MnO2 interface. (a) The O 1s core levels spectra of the
lattice peak (529.45 eV) and an additional non-lattice peak (531.33 eV), the
change of non-lattice oxygen indicates oxygen ions migration during switching;
(b) The spectra of Ti 2p, the shift of Ti 2p peak indicates an formation of interfacial TiOx layer, which can be regarded as an oxygen reservoir. Reprinted from
[79].
In the bipolar switching case, it is straightforward to conceive of
oxygen ions migrating away from the interface, since oxygen ions
are negative charged. Under a negative bias, oxygen ions can be
pushed toward the oxide bulk layer and then recombine with Vo and
rupture the CFs if the CFs are made up of Vo. C. Yoshida et al. [80]
performed C-AFM writing experiments in
18
O tracer gas atmos-
phere and related composition analysis by SIMS on the NiO films.
Their results suggest that external oxygen can penetrate into the oxide layer, and the composition change of the surface region may be
responsible for the resistance change. H. Y. Jeong et al. [81] performed EELS oxygen element mapping experiments in HRS and
LRS of Al/TiO2/Al structure. The results suggest that the drift of oxygen ions by the applied bias is the microscopic origin of the bistable resistivity switching. So the reset mechanism for bipolar switch-
15
ing is quite unambiguous in the literature [82-84]. Recently, the nonlinear dynamical properties of bipolar switching TiO2 were modeled
by the drift and diffusion of ionized dopants (oxygen vacancies) in
the oxide thin film [85-87]. In essence, the description using oxygen
vacancies is equivalent to the description using oxygen ions here.
In the unipolar switching case, the mainstream viewpoint of reset is
due to a thermal dissolution of CFs by local Joule heating [88-90].
Electro-thermal calculation [88] suggests that the local temperature
can rise to around 600 K during the switching process. But this
model is only phenomenological and does not reveal the microscopic nature of the rupture of CFs. Firstly, it does not address whether
the loss of oxygen during set process can be recaptured in subsequent switching cycles or not, and the model does not address the
question of whether the set process is a reversible process. Secondly,
if the CFs are assumed to be simply “dissolved” by local Joule heating, then the CFs should rupture in the middle of the filament where
the temperature is highest [91], since the metal electrodes work as a
large heat sink. The local Joule heating conjecture contradicts with
the many experimental observations that switching occurs in the region near the anode as mentioned before. Alternatively, M.-J. Lee et
al. [71] regarded the reset process of their unipolar NiO memory as
an oxidation process of the metallic CFs, which is a consequence of
the thermally activated oxygen atoms diffusing from anode side followed by oxidation of the Ni-rich CFs. This argument acknowledges
that the reset in unipolar switching is a Joule heating-assisted process. At the same time, this argument is essentially consistent with
16
the electrochemical ionic migration model for bipolar switching devices. The model provides an inherent link between unipolar and bipolar switching. In the next section, we discuss another important
question of why some materials exhibit unipolar characteristics,
while some other materials exhibit bipolar characteristics before we
make the final conclusion.
13.2.3 The Effect of Electrode Materials on Switching Modes
There are many reports [92-94] on the effect of electrode materials
on the resistive switching characteristics of metal oxide memory. In
most cases, the switching mode is not an inherent property of the oxide film but an effect of the interaction between the oxide and electrode materials. R. Waser et al. pointed out that for bipolar switching
the MIM structure should have some asymmetry, such as different
electrode materials, while for uniploar switching the MIM structure
may be symmetric [27]. Fig. 6 shows two examples: the Pt/ZnO/Pt
[61] and Pt/NiO/Pt [20] structure show uniplolar switching, while
the TiN/ZnO/Pt [95] and Pt/NiO/SrRuO3 [96] structure show bipolar
switching. Besides, many other metal oxides can show either unipolar or bipolar behavior depending on what electrode materials are
used. Pt/ZrO2/Pt [97], Pt/TiO2/Pt [21], Pt/HfO2/Pt [62] all show unipolar switching, and Ti/ZrO2/Pt [97], Pt/TiO2/TiN [98], TiN/HfO2/Pt
[99] all show bipolar switching. It should be noted that the unipolar
switching I-V curves are usually symmetric, since there is no distinction of either of the electrodes. As a result, for unipolar switching
devices, bipolar switching behaviors can also be demonstrated [100],
17
which means the set can be triggered at one polarity and reset can be
triggered at the reversed polarity. Such case is also referred to “nonpolar” switching [74]. But in bipolar switching devices, there is always one polarity that reset cannot occur when the same polarity bias as set process is applied.
Fig. 6 (a) I-V characteristics of Pt/ZnO/Pt and TiN/ZnO/Pt devices, (b) Pt/NiO/Pt
and Pt/NiO/SrRuO3 devices, showing the electrode materials effect on switching
modes.
Data are collected from:
Ref [1] W.-Y. Chang, Y.-C. Lai, T.-B. Wu, S.-F. Wang, F. Chen, and M.-J. Tsai,
Appl. Phys. Lett. 92, 022110 (2008).
Ref [2] N. Xu, L. F. Liu, X. Sun, C. Chen, Y. Wang, D. D. Han, X. Y. Liu, R. Q. Han,
J. F. Kang, and B. Yu, Semicond. Sci. Technol. 23, 075019 (2008).
Ref [3] S. Seo, M. J. Lee, D. H. Seo, E. J. Jeoung, D.-S. Suh, Y. S. Joung, and I. K.
Yoo, I. R. Hwang, S. H. Kim, I. S. Byun, J.-S. Kim, J. S. Choi, and B. H. Park,
Appl. Phys. Lett. 85, p. 5655 (2004).
Ref [4] J. S. Choi, J.-S. Kim, I. R. Hwang, S. H. Hong, S. H. Jeon, S.-O. Kang, B.
H. Park, D. C. Kim, M. J. Lee, and S. Seo, Appl. Phys. Lett. 95, 022109 (2009).
The electrode materials can be roughly classified into two groups:
one is noble metals, such as Pt, Au, Ru, etc, which are resistant to
oxidation; the other one is non-noble metals, such as Ti, Al, Cu, Ni,
W, TiN, TaN, etc, which may suffer oxidation during the switching
process. As observed above, devices using noble metals always
show unipolar behaviors. Even with the Pt/TiO2/TiN [98] structure,
unipolar switching can occur at the Pt side. However, the unipolar
switching cannot occur at TiN side, and only bipolar switching is obtained at the TiN side. As mentioned before, during the set process,
18
the oxygen ions migrate toward the anode. Thus, the anode materials
may react with the oxygen and form an interfacial layer if the anode
materials are oxidizable. P. Zhou et al. [101] compared the different
roles played by the Pt electrode and the TaN electrode on the switching modes of CuxO memory. Fig. 7 (a) and (b) shows depth profile
by Auger Electron Spectroscopy (AES) for Pt/CuxO/Cu structure
and TaN/CuxO/Cu structure, we can see that the oxygen concentration has an obvious shift towards the TaN layer, which indicates that
a reaction occurs between the TaN electrode and oxide film. This interfacial TaON layer was further confirmed by XPS binding energy
depth spectra. The bright gray areas under the TaN electrode in the
TEM picture of Fig. 7 (c) are assumed to be the TaON ultra-thin layer. And Fig. 7 (d) illustrates that with Pt as top electrode, reset can
occur at both bias polarities, showing a unipolar switching behavior,
while with TaN as top electrode, reset can occur only under the reversed bias, showing a bipolar switching behavior. Therefore, we
can infer that if noble metals like Pt are used for the electrode, it
hardly forms such interfacial oxide layer. And without the diffusion
barrier, the thermally activated oxygen ions may diffuse back during
the reset process to the oxide bulk layer since there is a large concentration gradient of oxygen across the interface region. This may
account for the unipolar reset mechanism. If oxidizable materials are
used for the electrode, it may form an interfacial oxide layer between the electrode and oxide films, which may act as an oxygen
diffusion barrier. Thus, it is difficult for the oxygen ions to diffuse
back through the thermal activation. Then only by the acceleration
19
of a reversed electric field can the oxygen ions drift back to rupture
the CFs. This may account for the bipolar reset mechanism. Numerical simulation [102] based on the oxygen ions non-linear transport
model is employed to support the above assumptions.
Fig. 7 (a) Auger Electron Spectroscopy (AES) depth profile of (a) Pt/Cu xO/Cu
structure and (b) TaN/CuxO/Cu structure, indicating the existence of an interfacial
TaON layer; (c) TEM picture of the TaN/CuxO/Cu structure, showing the interfacial TaON layer; (d) I-V characteristics of Pt/CuxO/Cu structure with unipolar
switching behavior and TaN/CuxO/Cu structure with bipolar switching behavior.
Reprinted from [101]
13.2.4 Summary of the Physical Mechanism for Resistive Switching in Metal Oxide Memory
Here, we would like to summarize the physical resistive switching mechanism in metal oxide memory. It should be noted, however,
while a self-consistent physical picture can be constructed through
the discussions above, the resistive switching phenomenon in metal
oxides involves a large variety of materials, so the model described
here may not apply to all combinations of materials.
Basically, resistive switching is conjectured to be due to the formation and dissolution of conductive filaments localized at the anode interface, and is a reversible switching process. The conductive
filaments may consist of oxygen vacancies or excess metallic precipitates. Multiple, parallel filaments may be formed. The filaments
may preferentially be located at the grain boundaries. Experimental
evidences suggest that the switching process is an electrochemical
process associated with oxygen migration, oxidation and reduction.
20
As shown in Fig. 8, during the electroforming process, soft dielectric
breakdown occurs and oxygen ions drift to the anode interface by
the high electric field, where they are discharged as neutral nonlattice oxygen if the anode materials are noble metals or react with
the oxidizable anode materials to form an interfacial oxide layer.
Meanwhile in the bulk, the resultant oxygen vacancies or metal precipitates form the highly conducting paths. During the reset process,
oxygen ions migrate back to the bulk either to recombine with the
oxygen vacancies or to oxidize the metal precipitates. In the unipolar
switching case, the Joule heating thermally activates the diffusion of
oxygen ions. Oxygen ions diffuse against the electric field from the
anode due to the concentration gradient. In the bipolar switching
case, usually interfacial layer between the anode and oxide acts as a
diffusion barrier, and oxygen ions can only drift back aided by the
electric field. Usually only a part of the conducting paths is ruptured,
leaving the bottom part of the conducting paths to be a virtual cathode that extend partially into the film. Then during the subsequent
set process, a process similar to the electroforming occurs but only
in the region near anode. The physical picture presented here does
not contradict with the previous thermal model for unipolar switching or ionic model for bipolar switching. And it can at best be
viewed as a phenomenological description of experimental observations. Details of the physics of switching remain an area of active research.
Fig. 8 Schematic of the switching process in metal oxide memory
21
13.3 Performances of Metal Oxide Memory Devices
Firstly, we would like to discuss the fabrication techniques briefly.
Currently, the fabrication of metal oxide memory devices uses mainly conventional semiconductor fabrication processes. Sputtering of
metals followed by annealing in oxygen ambient or reactive sputtering in oxygen ambient are the most common deposition techniques
for metal oxide layers. Other methods used are atomic layer deposition (ALD) [30, 43], pulsed laser deposition (PLD) [103], metal organic chemical vapor deposition (MOCVD) [26] and the sol-gel
method [104]. Several papers report resistive switching behavior of
self-assembly grown nanostructure, e.g. NiO nanowires [105-106].
The thickness of the metal oxide layer is usually around 10-50 nm.
The substrate of the devices is mainly silicon; however, there are
several efforts to fabricate metal oxide memory on transparent substrate, e.g. glass [107] or flexible substrate, e.g. polyethersulfone
(PES) [108-109]. In the rest of this section, we will discuss the characteristics of metal oxide memory devices including noise margin,
scalability, power consumption, speed, reliability and uniformity.
The noise margin for the read operation is characterized by the
HRS/LRS resistance ratio. Generally, a ratio >10 is desired for easier
sense amplifier design. Almost all the metal oxide memory devices
in the literature exceed this requirement.
Scalability is one of the key concerns for any kind of memory. Because of the filamentary conduction nature, metal oxide memory devices may be scaled down to the nanoscale. Sub-100nm feature size
22
cells have been fabricated by 193-nm lithography with a contact
hole structure [78], as well as by nanoimprint lithography or e-beam
lithography with a crossbar arrays structure [51, 110-111]. Recently,
aggressively scaled 30 nm×30 nm HfO2 memory devices in 1Kb array have been demonstrated with excellent yield [31]. Resistive
switching behavior has been successfully triggered by C-AFM tip on
a 10 nm × 10 nm region of NiO thin films [71], suggesting the potential to scale the cell size even down to the sub-10 nm regime. Fig.
9 plots the general scaling trend of HRS and LRS resistance. The
leakage current in HRS is mainly due to bulk leakage current. Thus,
the resistance of HRS increases as the inverse of the cell area, roughly following the Ohm’s law. The conduction current in LRS is mainly filamentary conduction current, as discussed before. So the resistance of LRS has only a slight dependency on the cell area. This
trend of increasing HRS/LRS resistance ratio as cell area decreases
is a benefit of device scaling.
Fig. 9 HRS and LRS resistance versus cell area of metal oxide memory devices.
Data are collected from:
Ref [1] I. G. Baek, M. S. Lee, S. Seo, M. J. Lee, D. H. Seo, D.-S. Suh, J. C. Park, S.
O. Park, H. S. Kim, I. K. Yoo, U.-I. Chung, and I. T. Moon, Tech. Dig. IEDM, p.
587 (2004).
Ref [2] H. Y. Lee, P. S. Chen, T. Y. Wu, Y. S. Chen, C. C. Wang, P. J. Tzeng, C. H.
Lin, F. Chen, C. H. Lien, and M.-J. Tsai, Tech. Dig. IEDM, p. 297 (2008).
Ref [3] N. Xu, B. Gao, L. F. Liu, B. Sun, X. Y. Liu, R. Q. Han, J. F. Kang, and B.
Yu, Tech. Dig. Symp. VLSI Technol., p. 100 (2008).
Ref [4] S. Kim, H. Moon, D. Gupta, S. Yoo, and Y.-K. Choi, IEEE Trans. Electron
Devices 56, p. 696 (2009).
23
Ref [5] M. K. Yang, J.-W. Park, T. K. Ko, and J.-K. Lee, Appl. Phys. Lett. 95,
042105 (2009).
Ref [6] H. B. Lv, M. Yin, X. F. Fu, Y. L. Song, L. Tang, P. Zhou, C. H. Zhao, T. A.
Tang, B. A. Chen, and Y. Y. Lin, IEEE Electron Device Lett. 29, p.309 (2008).
Ref [7] X. Sun, B. Sun, L. F. Liu, N. Xu, X. Y. Liu, R. Q. Han, J. F. Kang, G. Xiong,
and T. P. Ma, IEEE Electron Device Lett. 30, p. 334 (2009).
Although the read noise margin is improved in the scaled memory
cell, the filamentary conduction nature of the memory cell results in
increasing power density as devices scale down. The typical switching voltage in literature is around 1-5 V, while the switching current
can be many orders different, depending on the material and the device structure. The key to reducing power consumption is to reduce
the reset switching current. The peak value of current in metal oxide
memory is the LRS current at the point when the reset process occurs, which is usually referred to as the reset current. Recently, Y.
Wu et al. [112] have demonstrated extremely low reset current (~1
μA) for ALD grown Al2O3 memory cell. Yet, most reports in the literature show a typical reset current for a single device on the order
of mA or hundreds of μA. Fig. 10 plots the general scaling trend of
reset current and corresponding reset current density. It is seen that
the reset current reduce only slightly when the devices are scaled
down, thus leading to a remarkable increase of the current density
required for reset. This presents a significant challenge for the
memory cell selection devices, which need to provide such a large
current density for ultra-scaled cells, e.g. ~ MA/cm2 even for a relatively large 100 nm×100 nm cell area. Although there is contrary
24
viewpoint that the reset current would decrease with reduced cell area [113], it should be noted that it is not a fair comparison because in
those experiments smaller set compliance was applied for the smaller memory cell. We will see in the following discussions that the reset current value can be modulated by controlling of the set compliance value. Fig. 11 plots the relationship between reset current and
set compliance current. It is seen that reset current is usually only a
little larger than the set compliance current, and more importantly,
the reset current decreases almost linearly with the decrease of set
compliance current. A possible reason for this relationship is that
with a smaller set compliance current, less oxygen atoms migrate to
the anode and weaker CFs are formed during the set process. Thus a
smaller reset current is needed to rupture the CFs. In the literature
the reset current of single memory cell is usually difficult to be reduced down to sub mA regime, even if sub mA set compliance current is applied. The deviation from the linear relationship between
reset current and set compliance is most probably due to the parasitic
capacitance in the measurement setup as discussed in [114]. Thus the
reset current tends to be elevated at the mA level even if the set
compliance of the parameter analyzer used in the measurement is
reduced down to the μA level. By using an on-chip 1 transistor-1 resistor (1T-1R) structure rather than the single resistor (1R) structure,
the parasitic capacitance is greatly reduced, thus the reset current
tends to follow the linear relationship with set compliance current.
Sub-100 μA reset current has been successfully demonstrated in several 1T-1R cell structures by adjusting the selection transistor’s gate
25
voltage to deliver a small set compliance current [30, 115-117]. As a
consequence, however, the smaller set compliance may result in a
smaller HRS/LRS resistance ratio. Fig. 12 plots the dependence of
the LRS resistance on the set compliance current. The LRS resistance increases almost inversely with the decreasing set compliance current. Smaller set compliance forms weaker CFs, thus leading to a larger LRS resistance. Given certain HRS resistance, the
requirement for a sufficient ratio limits the smallest set compliance
that can be used. Remember that HRS resistance rises up rapidly
with the decrease of cell area, as mentioned before. This means for
the scaled memory cell, a larger LRS resistance can be tolerated, and
a smaller set compliance can be utilized to achieve lower reset current. With this methodology, by using 5 μA set compliance, a greatly
reduced reset current (~50 μA) is demonstrated in a 100 nm×100 nm
NiO memory cell with sufficient HRS/LRS ratio (~104) [71, 113].
Fig. 10 The peak value of reset current and corresponding current density versus
cell area
Data are collected from:
Ref [1] I. G. Baek, M. S. Lee, S. Seo, M. J. Lee, D. H. Seo, D.-S. Suh, J. C. Park, S.
O. Park, H. S. Kim, I. K. Yoo, U.-I. Chung, and I. T. Moon, Tech. Dig. IEDM, p.
587 (2004).
Ref [2] H. Y. Lee, P. S. Chen, T. Y. Wu, Y. S. Chen, C. C. Wang, P. J. Tzeng, C. H.
Lin, F. Chen, C. H. Lien, and M.-J. Tsai, Tech. Dig. IEDM, p. 297 (2008).
Ref [3] C. Rohde, B. J. Choi, D. S. Jeong, S. Choi, J.-S. Zhao, and C. S. Hwang,
Appl. Phys. Lett. 86, 262907 (2005).
Ref [4] K. M. Kim, B. J. Choi, B. W. Koo, S. Choi, D. S. Jeong, and C. S. Hwang,
Electrochem. Solid-State Lett., 9, G343 (2006).
26
Ref [5] K. M. Kim, B. J. Choi, and C. S. Hwang, Appl. Phys. Lett. 91, 012907
(2007).
Ref [6] C. Yoshida, K. Tsunoda, H. Noshiro, and Y. Sugiyama, Appl. Phys. Lett. 91,
223510 (2007).
Ref [7] Y. C. Shin, J. Song, K. M. Kim, B. J. Choi, S. Choi, H. J. Lee, G. H. Kim, T.
Eom, and C. S. Hwang, Appl. Phys. Lett. 92, 162904 (2008).
Ref [8] W.-Y. Chang, Y.-T. Ho, T.-C. Hsu, F. Chen, M.-J. Tsai, and T.-B. Wu, Electrochem. Solid-State Lett., 12, H135 (2009).
Ref [9] X. Sun, B. Sun, L. F. Liu, N. Xu, X. Y. Liu, R. Q. Han, J. F. Kang, G. Xiong,
and T. P. Ma, IEEE Electron Device Lett. 30, p. 334 (2009).
Fig. 11 The reset current versus set compliance current
Data are collected from:
Ref [1] S. Seo, M. J. Lee, D. H. Seo, E. J. Jeoung, D.-S. Suh, Y. S. Joung, and I. K.
Yoo, I. R. Hwang, S. H. Kim, I. S. Byun, J.-S. Kim, J. S. Choi, and B. H. Park,
Appl. Phys. Lett. 85, p. 5655 (2004).
Ref [2] K. Kinoshita, K. Tsunoda, Y. Sato, H. Noshiro, S. Yagaki, M. Aoki, and Y.
Sugiyama, Appl. Phys. Lett. 93, 033506 (2008).
Ref [3] K. Tsunoda, K. Kinoshita, H. Noshiro, Y. Yamazaki, T. Iizuka, Y. Ito, A.
Takahashi, A. Okano, Y. Sato, T. Fukano, M. Aoki, and Y. Sugiyama, Tech. Dig.
IEDM, p. 767 (2007).
Ref [4] C. Rohde, B. J. Choi, D. S. Jeong, S. Choi, J.-S. Zhao, and C. S. Hwang,
Appl. Phys. Lett. 86, 262907 (2005).
Ref [5] H. Shima, F. Takano, H. Muramatsu, H. Akinaga, Y. Tamai, I. H. Inque,
and H. Takagi, Appl. Phys. Lett. 93, 113504 (2008)
Fig. 12 LRS resistance versus set compliance current
Data are collected from:
Ref [1] K. Tsunoda, K. Kinoshita, H. Noshiro, Y. Yamazaki, T. Iizuka, Y. Ito, A.
Takahashi, A. Okano, Y. Sato, T. Fukano, M. Aoki, and Y. Sugiyama, Tech. Dig.
27
IEDM, p. 767 (2007).
Ref [2] B. Lee, H.-S. P. Wong, Tech. Dig. Symp. VLSI Technol., p. 28 (2009).
Ref [3] C.-Y. Lin, C.-Y. Wu, C.-Y. Wu, C. Hu, and T.-Y. Tseng, J. Electrochem. Soc.
154, G189 (2007).
Ref [4] D. Lee, D.-J. Seong, H. J. Choi, I. Jo, R. Dong ,W. Xiang, ,S. Oh, M. Pyun,
S.-O Seo, S. Heo, M. Jo, D.-K. Hwang, H. K Park, M. Chang, M. Hasan, and H.
Hwang, Tech. Dig. IEDM, p. 1 (2006).
An attractive feature of metal oxide memory is the nanosecond
switching speed, which is comparable to DRAM and is substantially
faster than Flash on a single bit programming basis. It is at least
three to ten times faster than the emerging phase change memory.
Fig. 13 plots the HRS and LRS resistance versus the programming
pulse width. It is seen that with smaller pulse width, HRS resistance
drops while LRS resistance rises, leading to a smaller HRS/LRS resistance ratio. The shorter pulses result in a less complete switching
process. Thus, the requirement to maintain a sufficient noise margin
limits the switching speed. The measurement of switching speed involves a delicate procedure [118]: a load resistor is connected in series with the memory cell; the programming pulse (Vp) is applied to
the stacked resistor and cell structure through a pulse generator,
while an oscilloscope is used to monitor the voltage (Vcell) across the
memory cell. Note that Vcell = Vp×Rcell / (Rcell+RL), thus the time
when Vcell suddenly rises or drops can be estimated as the switching
time of reset or set process for the applied cell voltage Vcell. The
magnitude of the current can be calculated from the measured cell
voltage at the oscilloscope divided by load resistor, Icell = Vcell/Rcell.
28
With this methodology, the relationship between the switching time
and the pulse amplitude can be investigated. There are reports indicating that in order to achieve faster switching, a larger pulse amplitude is required [119]. Therefore, a trade-off exists between fast
switching speed and low power consumption. Bipolar switching devices have shown ultra-fast switching speed using reasonable voltage pulse amplitudes, e.g. +2 V/10 ns for set and -1.5 V/10 ns for reset in TaOx memory [64], +3.2 V/5 ns for set and -2.7 V/5 ns for
reset in HfO2 memory [30]. Some earlier works showed that for unipolar switching devices, the reset process needs a much longer pulse
width than set process, e.g. +1 V/10 ns for set and +0.5 V/5 μs for
reset in NiO memory [73], +3 V/10ns for set and +2.5 V/5 μs for reset in TiO2 memory [120]. It is suggested that a certain amount of
time (~μs) is necessary to sufficiently heat up the CFs to trigger the
rupture [120], since the unipolar reset mechanism involves a thermally assisted oxygen diffusion process. But recently, several works
demonstrated that ultra-fast reset switching speed that is comparable
to the set switching speed is also achievable in unipolar switching
devices, e.g. +2.8 V/10 ns for set and +1.8 V/5 ns for reset in Tidoped NiO memory [121], +1.5 V/10 ns for set and +1 V/10 ns for
reset in scaled NiO memory [113]. It is shown that the reset switching time decreases significantly with decreasing cell area, from 1 μs
(30 μm×30 μm cell) to 10 ns (100 nm×100 nm cell) for unipolar
NiO memory [71]. If individual CFs are similar in size or resistivity
and only the number of CFs varies between different cell areas, then
the switching time should be insensitive to the cell area, because
29
CFs are connected in parallel and ruptured independently. Therefore,
it is inferred the CFs should be weaker or have larger resistivity for
the smaller cells [71]. This trend that smaller cells have faster
switching speed can also be due to the smaller parasitic capacitance
of the smaller cell.
Fig. 13 HRS and LRS resistance versus the programming pulse width
Data are collected from:
Ref [1] S. Muraoka, K. Osano, Y. Kanzawa, S. Mitani, S. Fujii, K.Katayama, Y.
Katoh, Z. Wei, T. Mikawa, K. Arita, Y. Kawashima, R. Azuma, K. Kawai, K.
Shimakawa, A. Odagawa, and T. Takagi, Tech. Dig. IEDM, p. 779 (2007).
Ref [2] K. Tsunoda, K. Kinoshita, H. Noshiro, Y. Yamazaki, T. Iizuka, Y. Ito, A.
Takahashi, A. Okano, Y. Sato, T. Fukano, M. Aoki, and Y. Sugiyama, Tech. Dig.
IEDM, p. 767 (2007).
Ref [3] A. Chen, S. Haddad, Y. C. Wu, Z. Lan, T. N. Fang, and S. Kaza, Appl. Phys.
Lett. 91, 123517 (2007).
Ref [4] B. J. Choi, S. Choi, K. M. Kim, Y. C. Shin, C. S. Hwang, S.-Y. Hwang, S.-S.
Cho, S. Park, and S.-K. Hong, Appl. Phys. Lett. 89, 012906 (2006).
Ref [5] Y. H. Do, J. S. Kwak, Y. C. Bae, K. Jung, H. Im, and J. P. Hong, Appl.
Phys. Lett. 95, 093507 (2009).
The reliability characteristics include the endurance and retention.
Current Flash technology shows a maximum number of programming cycles between 103 and 107, depending on the device type.
Metal oxide memory should provide at least the same endurance,
preferably a better one. Excellent endurance records in the literature
are shown for unipolar NiO memory (>106) [73], bipolar HfO2
memory (>106) [30], and bipolar TaOx memory (>109) [64] at the
30
single cell level. The failure of switching may be due to grain structure change near the anode region [56] and insufficient non-lattice
oxygen ions to rupture the CFs [83], since during the repeated
switching cycles, oxygen may escape from the device. Most nonvolatile memory applications specify at least 10 years data retention
time at typical operating temperatures (e.g. 85 0C). Promising retention records in the literature are shown for unipolar Ti-doped NiO
memory (>1000 hours @ 150 0C) [121], and bipolar TaOx memory
(>3000 hours @ 150 0C) [64]. A practical method to estimate the retention capability is to draw the Arrhenius plot of retention time versus the reciprocal of temperature, and extrapolate the data obtained
in high temperature regime to the operating temperature (e.g. 85 0C)
[64, 122]. The loss of data in LRS or HRS may be associated with
the diffusion of oxygen ions or oxygen vacancies due to their concentration gradient, and this process can be accelerated at elevated
temperature because of the temperature dependent diffusion coefficient. This is the basis of any temperature varying experiments to estimate the retention time. A more complicated statistical method to
estimate the retention capability by temperature–accelerated measurement is developed by A. Chen et al. [123].
Uniformity of device characteristics is a key problem that hinders
the metal oxide memory from entering large-scale manufacturing.
Significant parameter fluctuations exist in terms of variations of the
switching voltages as well as the resistance in HRS and LRS. These
parameters fluctuate from programming cycle to programming cycle
and from device to device. Many efforts have been expended to im-
31
prove the uniformity. D. C. Kim et al. [124] proposed to use IrO2 as
the top electrode for the NiO memory to help stabilize the local oxygen migrations for the formation and rupture of CFs. W.-Y. Chang et
al. [125] proposed to embed Pt nanocrystals into the TiO2 memory
to provide an easy path for the formation and rupture of CFs by local
enhancement of the electric field adjacent to the Pt nanocrystals. S.
Yu et al. [126] proposed to stack a thin Al buffer layer on HfO2
memory to diffuse Al impurity atoms into the switching layer, aiming to stabilize the formation and rupture of CFs along the impurities
due to a lower Vo formation energy caused by the impurities. Q. Liu
et al. [127] proposed to implant Ti into ZrO2 memory to provide a
source of Vo and play the role of a seed for forming CFs. The key
idea of these techniques is to induce the formation and rupture of
CFs in specific locations during the switching instead of allowing
them to occur randomly. Besides materials solutions mentioned
above, novel programming methods are also helpful in reducing parameter fluctuations. Using multiple pulses rather than a single pulse
[128] or using a ramped series of pulses [129] can remarkably improve the cycle to cycle uniformity.
To summarize, the scalability of metal oxide memory devices is the
most competitive characteristics compared with other emerging
memories. It is foreseeable to achieve sufficient noise margin
(>10X), low energy consumption (<pJ per switching cycle), fast
switching speed (~ns) for nanoscale (10 nm×10 nm) cells in the near
future. However, the reliability and uniformity of metal oxide
memory needs to be further improved before this technology can be
32
commercialized.
13.4 Cell Structure of Metal Oxide Memory Arrays
For random access application, each metal oxide memory cell is
usually connected with a cell selection transistor). This cell structure
is referred to as 1 transistor-1 resistor (1T-1R). By controlling the selection transistor’s gate voltage, varying current can be provided to
control the switching characteristics of memory cell. As mentioned
in Section 3, a smaller set compliance results in a larger LRS resistance, which is preferred to reduce reset current. At the same time,
the set compliance should be larger enough to guarantee a sufficient
HRS/LRS resistance ratio. By adjusting the selection transistor’s
gate voltage, it is easy to optimize the programming current for the
set and reset processes [115-116, 119]. However, the area penalty of
using a selection transistor is apparent, even with a delicate layout
design of integrating the memory cell into the via contact of the
transistor. The 1T-1R cell area can range from 6F2 (F: feature size
used for patterning the cell) using aggressive DRAM-like design
rules with borderless contacts and zero gate to source/drain spacing
to 8F2 (with F/2 gate to source/drain spacing) and 18F2 (contacts
with borders).
For high-density integration, a crossbar matrix with 4F2 area is
preferred. The simplest way is to connect the word and bit lines at
33
each node by a memory cell. The crossbar matrix can be a set of orthogonal nanoscale metal wires with the metal oxide as the memory
element at the crossbar junction. TiO2 memory with junction area
(50 nm×50 nm) [51], NiO memory with junction area (70 nm×70
nm) [110] have been reported. B. Lee et al. [111] demonstrated a
novel structure for NiO memory which utilizes the sidewalls between the two nanoscale metal wires as the active switching region.
Thus, the active memory cell area is even smaller than the layout
cell area (48 nm×48 nm).
A cross-talk problem with the above simple crossbar matrix may
arise in read operation [130]. As shown in Fig. 14 (a), if the cell to
be read out happens to be in HRS with surrounding cells in LRS, the
reading current can easily flow through the surrounding cells in the
LRS and thus a LRS data will be mistakenly read out. In order to
avoid these parasitic leakage paths, a cell selection element with
large I-V non-linearity should be added at each node. The cell selection element is usually envisioned to be a diode (with an exponential
I-V characteristics). Therefore, this cell structure refers to 1 diode-1
resistor (1D-1R). The reverse biased diodes effectively cuts off the
leakage current paths, thus the interference between neighboring
cells is prevented, as shown in Fig. 14 (b). The diode on/off ratio required ranges from 104 to 106 depending on the memory array size,
the on/off ratio of the memory cell, and the resistance of the interconnect wire, etc. It should be noted for the 1D-1R structure, the resistive switching element must be unipolar switching devices because the diode limits the reversed current that a bipolar switching
34
device requires.
Fig. 14 The crossbar matrix (a) without cell selection elements and (b) with cell
selection elements
With such crossbar structure, the cell area can in principle be
scaled down to 4F2. Furthermore, by stacking the crossbar structure
in the third dimension, the effective cell size can be further reduced
to 2F2, 1F2, and so on [131]. The 1D-1R structure can be stacked for
3-D integration. Thus there are successive attempts to fabricate
stacked nanowire arrays for NiO memory with different cell selection elements [132-135].
Here we would like to discuss the cell selection element material.
A p–n diode is the most common device for the cell selection element. Although high performance p–n diode is easily fabricated with
current epitaxial silicon technology for the planar device structures,
it is not feasible to implement epitaxial silicon-based p–n diode into
the stacked crossbar metal oxide memory structures because it is difficult to grow epitaxial silicon on a metal layer and high processing
temperatures are required. On the other hand, amorphous silicon allows for lower processing temperatures. But it does not meet the requirement for current density for memory cell programming. Therefore, new materials need to be explored for the cell selection
element, which should both allow for low processing temperatures
and also provide high current drive. Compared with silicon diode,
oxide diode is attractive because it offers better flexibility in processing technology because it can be fabricated over any substrate
35
even at room temperature [131]. If the oxide material is oxygen deficient with sufficient amount of oxygen vacancies, such as TiO2,
ZrO2, ZnO and indium tin oxide (ITO), it is n-type; while if the oxide material is metal deficient with sufficient amount of metal vacancies, such as NiO, it is p-type. Thus a combination of p-type oxide and n-type oxide essentially forms a p-n diode. The research on
oxide diode started earlier than that on the metal oxide memory even
in the 1990s, e.g. p-NiO/n-Al: ZnO [136], p-NiO/n-ITO [137]. Recently, several kinds of oxide diodes [133], such as p-NiO/n-TiO2, pNiO/n-ZnO, p-NiO/n-InZnO, p-CuO/n-InZnO, have been stacked
with Pt/NiO/Pt structure in series, among which p-CuO/n-InZnO is
regarded as the best candidate. Besides the p-n oxide diode, single
oxide materials can also be applied to rectify the current. Y. C. Shin
et al. [138] demonstrated a Schottky-type cell structure consisting of
Pt/ITO/TiO2/Pt stacked layers. The low and high potential barrier at
the ITO/TiO2 and TiO2/Pt junctions, respectively, constitute the rectifying properties of the stacked structure, where TiO2 acts as the resistive switching layer. VO2 is found to exhibit electric field-induced
metal–insulator transition [139]. Unlike other metal oxide memory
devices, the resistive switching behavior in VO2 is not bistable, so it
is sometimes referred to threshold switching. The sharp transition of
current in VO2 when turned on provides an ideal behavior for
switching element. M.-J. Lee et al. [140] demonstrated a prototype
cell structure consisting of Pt/VO2/Pt/NiO/Pt stacked layers. Table
13.2 compares several switching elements mentioned here for the
metal oxide memory cell in aspects of forward current density, for-
36
ward/reverse current ratio, and turn on voltage. Generally, high forward current density, high forward/reverse current ratio, and low
turn on voltage are preferred. Programming current density of the
order of 10 MA/cm2 is required for metal oxide memory today.
Thus, the oxide based diodes reported to date are several orders of
magnitude too low in current drive for the crossbar memory application.
Forward Current
Density (A/cm2)
Forward/Reverse
Current Ratio
Turn on Voltage
(V)
p-NiO/n-TiO2
Ref [1]
p-NiO/n-ZnO
Ref [1]
p-NiO/n-IZO
Ref [1]
p-CuO/n-IZO
Ref [1]
ITO
Ref [2]
VO2
Ref [3]
~20
~300
~10,000
~30,000
~1,400
~4,000*
4.7×105
2.6×105
3×104
3×104
1×104
N/A
2.72
1.55
0.73
0.66
0.62
0.6
Table 13.2 Comparison of several oxide switching elements for the metal oxide
memory cell. Forward current density is extracted at +2V bias, forward/reverse
current ratio is defined at the current ratio at +2V and -2V, and turn on voltage is
defined as J = 100 A/cm2. *estimated.
Data are collected from:
Ref [1] M.-J. Lee, Y. Park, B.-S. Kang, S.-E. Ahn, C. Lee, K. Kim, W. Xianyu, G.
Stefanovich, J.-H. Lee, S.-J. Chung, Y.-H. Kim, C.-S. Lee, J.-B. Park, I.-G. Baek,
and I.-K.Yoo, Tech. Dig. IEDM, p. 771 (2007).
Ref [2] Y. C. Shin, J. Song, K. M. Kim, B. J. Choi, S. Choi, H. J. Lee, G. H. Kim, T.
Eom, and C. S. Hwang, Appl. Phys. Lett. 92, 162904 (2008).
Ref [3] M.-J. Lee, Y. Park, D.-S. Suh, E.-H. Lee, S. Seo, D.-C. Kim, R. Jung, B.-S.
Kang, S.-E. Ahn, C. B. Lee, D. H. Seo, Y.-K. Cha, I.-K. Yoo, J.-S. Kim, and B. H.
Park, Adv. Mater. 19, p. 3919 (2007).
Recently, M.-J. Lee et al. [141] demonstrated a 8×8 1D-1R array
with word/bit line selection transistors, as shown in Fig. 14. The
37
memory cell is aPt/NiO/Pt/p-CuO/n-InZnO/Pt stacked structure, and
the word line selection transistor is amorphous gallium indium zinc
oxide (GaInZnO) thin film transistor (TFT). Oxide TFT rather than
silicon transistor is utilized for the low processing temperature required for 3-D integration.
Fig. 15 (a) Scanning electron microscopy (SEM) image of an 8×8 1D–1R arrays
integrated with GaInZnO selection TFTs. Top image is the cross-section view of a
row in the 8×8 1D-1R arrays. (b) Image of the integrated devices shown over a
glass substrate. (c) I–V characteristics of a NiO cell, a 1D–1R cell, and a 1D–1R
cell with TFT selection transistor. (d) Schematic diagram of a 1D–1R cell with selection transistor. Reprinted from [141]
13.5 Summary
A review of recent research progress of metal oxide memory is presented here. Possible physical mechanism of resistive switching in
metal oxides is discussed. Filamentary conductive paths dominate
the conduction in LRS. Electrochemical/electrothermal oxygen migration, oxidation and reduction may be the microscopic origin of
resistive switching behavior. Electrode materials play an important
role in determining the switching modes: bipolar or unipolar. The
device characteristics of metal oxide memory cells reported in the
literature are summarized. Sufficient noise margin, excellent scalability, low power consumption, ultra-fast speed, good endurance and
retention make metal oxide memory a competitive candidate for fu-
38
ture non-volatile memory. Device uniformity must be further improved in order to meet the large-scale manufacturing requirements.
Cell structure for high density memory array is discussed. The 1D1R crossbar array structure is promises ultra-high integration density
with the potential for 3-D integration. Low process temperature
memory cell selection element with the required current drive and
on/off ratio is the key to realizing 3-D stackable crossbar memory
arrays.
39
Reference
1. Burr, G.W., Kurdi, B.N., Scott, J.C., Lam, C. H., Gopalakrishnan, K., Shenoy, R.S.:
Overview of candidate device technologies for storage-class memory. IBM J. Res.
Dev. 52, 449 (2008)
2. Kryder, M.H., Chang, S.K.: After Hard Drives – What Comes Next? IEEE Trans.
Magn. 45, 3406 (2009)
3. Zhu, J.-G.: Magnetoresistie Random Access Memory: The Path to Competitiveness and
Scalability. Proc. IEEE 96, 1786 (2008)
4. Arimoto, Y., Ishiwara, H.: Current Status of Ferroelectric Random-Access Memory.
MRS Bul. 29, 823 (2004)
5. Meijer, G.I.: Who Wins the Nonvolatile Memory Race? Sci. 319, 1625 (2008)
6. Wuttig, M., Yamada, N.: Phase-change materials for rewriteable data storage. Nat.
Mater. 6, 824 (2007)
7. Pirovano, A., Lacaita, A.L., Benvenuti, A., Pellizzer, F., Bez, R: Electronic switching
in phase-change memories. IEEE Trans. Electron Devices 51,452 (2004)
8. Lee, S.H., Jung, Y., Agarwal, R.: Highly scalable non-volatile and ultra-low-power
phase-change nanowire memory. Nat. Nanotechnol. 2, 626 (2007)
9. Kim, S., Zhang, Y., McVittie, J.P., Jagannathan, H., Nishi, Y., Wong, H.-S.P.: Integrating Phase-Change Memory Cell with Ge Nanowire Diode for Crosspoint Memory –
Experimental Demonstration and Analysis. IEEE Trans. Electron Devices 55, 2307
(2008)
10. Kozicki, M.N., Park, M., Mitkova, M.: Nanoscale memory elements based on solidstate electrolytes. IEEE Trans. Nanotechnol. 4, 331 (2005)
11. Wang, Z., Griffin, B., McVittie, J., Wong, S., McIntyre, C., Nishi, Y.: Resistive Switching Mechanism in ZnxCd1-xS Nonvolatile Memory Devices. IEEE Electron Device Lett.
28, 14 (2007)
12. Scott, J.C., Bozano, L.D.: Nonvolatile Memory Elements Based on Organic Materials.
Adv. Mater. 19, 1452 (2007)
13. Hong, S.H., Kim, O., Choi, S., Ree, M.: Bipolar resistive switching in a single layer
memory device based on a conjugated copolymer. Appl. Phys. Lett. 91, 093517 (2007)
14. Hickmott, T.W.: Low-Frequency Negative Resistance in Thin Anodic Oxide Films. J.
Appl. Phys. 33, 2669 (1962)
15. Gibbons, J.F., Beadle, W.E.: Switching propertiesl of thin NiO Films. Solid-State Electron. 7, 785 (1964)
16. Dearnaley, G., Stoneham, A.M., Morgan, D.V.: Electrical phenomena in amorphous
oxide films. Rep. Prog. Phys. 33, 1129 (1970)
17. Asamitsu, A., Tomioka, Y., Kuwahara, H., Tokura, Y.: Current switching of resistive
states in magnetoresistive manganites. Nat. 388, 50 (1997)
18. Beck, A., Bednorz, J.G., Gerber, C., Rossel, C., Widmer, D.: Reproducible switching
effect in thin oxide films for memory applications. Appl. Phys. Lett. 77, 139 (2000)
19. Zhuang, W.W., Pan, W., Ulrich, B.D., Lee, J.J., Stecker, L., Burmaster, A., Evans,
D.R., Hsu, S.T., Tajiri, M., Shimaoka, A., Inoue, K., Naka, T., Awaya, N., Sakiyama,
K., Wang, Y., Liu, S.Q., Wu, N.J., Ignatiev, A.: Novel colossal magnetoresistive thin
film nonvolatile resistance random access memory (RRAM). Electron Devices Meet.,
2002. IEDM ’02. Digest. Int., 193 (2002)
20. Seo, S., Lee, M.J., Seo, D.H., E. J. Jeoung, Suh, D.-S., Y. S. Joung, Yoo, I.K., I. R.
Hwang, S. H. Kim, I. S. Byun, J.-Kim, S., Choi, J.S., Park, B.H.: Reproducible resistance switching in polycrystalline NiO films. Appl. Phys. Lett. 85, 5655 (2004)
21. Rohde, C., Choi, B.J., Jeong, D.S., Choi, S., Zhao, J.-S., Hwang, C.S.: Identification of
a determining parameter for resistive switching of TiO 2 thin films. Appl. Phys. Lett.
86, 262907 (2005)
22. Lee, D., Choi, H., Sim, H., Choi, D., Hwang, H., Lee, M.-J., Seo, S.-A., Yoo, I.K.: Resistance switching of the nonstoichiometric zirconium oxide for nonvolatile memory
40
applications. IEEE Electron Device Lett. 26, 719 (2005)
23. Villafuerte, M., Heluani, S.P., Juárez, G., Simonelli, G., Braunstein, G., Duhalde, S.:
Electric-pulse-induced reversible resistance in doped zinc oxide thin films. Appl. Phys.
Lett. 90, 052105 (2007)
24. Dong, R., Lee, D.S., Xiang, W.F., Oh, S.J., Seong, D.J., Heo, S.H., Choi, H.J., Kwon,
M.J., Seo, S.N., Pyun, M.B., Hasan, M., Hwang, H.: Reproducible hysteresis and resistive switching in metal-CuxO-metal heterostructures. Appl. Phys. Lett. 90, 042107
(2007)
25. Lin, C.-Y., Wu, C.-Y., Wu, C.-Y., Hu, C., Tseng, T.-Y.: Bistable Resistive Switching in
AL2O3 Memory Thin Films. J. Electrochem. Soc. 154, G189 (2007)
26. Lee, S., Kim, W.-G., Rhee, S.-W., Yong, K.: Resistance Switching Behaviors of Hafnium Oxide Films Grown by MOCVD for Nonvolatile Memory Applications. J. Electrochem. Soc. 155, H92 (2008)
27. Waser, R., Aono, M.: Nanoionics-based resistive switching memories. Nat. Mater. 6,
833 (2007)
28. Waser, R., Dittmann, R., Staikov, G., Szot, K.: Redox-Based Resistive Switching Memories – Nanoionic Mechanisms, Prospects, and Challenges. Adv. Mater. 21, 2632
(2009)
29. Sawa, A.: Resistive switching in transition metal oxides. Mater. Today 11, 28 (2008)
30. Lee, H.Y., Chen, S., Wu, T.Y., Chen, Y.S., Wang, C.C., Tzeng, J., Lin, C.H., Chen, F.,
Lien, C.H., Tsai, M.-J.: Low power and high speed bipolar switching with a thin reactive Ti buffer layer in robust HfO2 based RRAM. Electron Devices Meet., 2008. IEDM
2008. IEEE Int., 297 (2008)
31. Chen, Y.S., Lee, H.Y., Chen, S., Gu, Y., Chen, C.W., Lin, W.P., Liu, W.H., Hsu, Y.Y.,
Sheu, S.S., Chiang, C., Chen, W.S., Chen, F.T., Lien, C.H., Tsai, M.-J.: Highly scalable hafnium oxide memory with improvements of resistive distribution and read disturb immunity. Electron Devices Meet. (IEDM), 2009 IEEE Int., 105 (2009)
32. International Technology Roadmap for Semiconductors, 2007 edition.
http://www.itrs.net/ (2007).
33. Sato, Y., Kinoshita, K., Aoki, M., Sugiyama, Y.: Consideration of switching mechanism of binary metal oxide resistive unctions using a thermal reaction model. Appl.
Phys. Lett. 90, 033503 (2007)
34. Russo, U., Ielmini, D., Cagli, C., Lacaita, A.L., Spiga, S., Wiemer, C., Perego, M.,
Fanciulli, M.: Conductive-filament switching analysis and self-accelerated thermal
dissolution model for reset in NiO-based RRAM. Electron Devices Meet., 2007. IEDM
2007. IEEE Int., 775 (2007)
35. Fujii, T., Kawasaki, M., Sawa, A., Akoh, H., Kawazoe, Y., Tokura, Y.: Hysteretic current-voltage characteristics and resistance switching at an epitaxial oxide Schottky
junction SrRuO3/SrTi0.99Nb0.01O3. Appl. Phys. Lett. 86, 012107 (2005)
36. Xia, Y., He, W., Chen, L., Meng, X., Liu, Z.: Filed-induced resistive switching based
on space-charge-limited current. Appl. Phys. Lett. 90, 022907 (2007)
37. Rozenberg, M.J., Inoue, I.H., Sanchez, M.J.: Strong electron correlation effects in
nonvolatile electronic memory devices. Appl. Phys. Lett. 88, 033510 (2006)
38. Sanchez, M.J., Inoue, I.H., Rozenberg, M.J.: A mechanism for unipolar resistance
switching in oxide nonvolatile memory devices. Appl. Phys. Lett. 91, 252101 (2007)
39. Nian, Y.B., Strozier, J., Wu, N.J., Chen, X., Ignatiev, A.: Evidence for an Oxgen Diffusion Model for the Electric Pulse Induced Resistance Change Effect in TransitionMetal Oxides. Phys. Rev. Lett. 98, 146403 (2007)
40. Jeong, D.S., Schroeder, H., Waser, R.: Mechanism for bipolar switching in a
Pt/TiO2/Pt resistive switching cell. Phys. Rev. B 79, 195317 (2009)
41. Tsunoda, K., Fukuzumi, Y., Jameson, J.R., Wang, Z., Griffin, B., Nishi, Y.: Bipolar resistive switching in polycrystalline TiO2. Appl. Phys. Lett. 90, 113501 (2007)
42. Sakamoto, T., Banno, N., Iguchi, N., Kawaura, H., Sunamura, H., Fujieda, S., Terabe,
41
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
K., Hasegawa, T., Aono, M.: A Ta2O5 solid-electrolyte switch with improved reliability. 2007 IEEE Symp. VLSI Technol., 38 (2007)
Choi, B.J., Jeong, D.S., Kim, S.K., Rohde, C., Choi, S., Oh, J.H., Kim, H.J., Hwang,
C.S., Szot, K., Waser, R., Reichenberg, B., Tiedke, S.: Resistive switching mechanism
of TiO2 thin films grown by atomic-layer deposition. J. Appl. Phys. 98, 33715 (2005)
Szot, K., Speier, W., Bihlmayer, G., Waser, R.: Switching the electrical resistance of
indivicual dislocations in single-cyrstalline SrTiO3. Nat. Mater. 5, 312 (2006)
Kim, K.M., Choi, B.J., Koo, B.W., Choi, S., Jeong, D.S., Hwang, C.S.: Resistive
Switching in Pt/Al2O3/TiO2/Ru Stacked Structures. Electrochem. Solid-State Lett. 9,
G343 (2006)
Lee, D., Seong, D.-J., Jo, I., Xiang, F., Dong, R., Oh, S., Hwang, H.: Resistance
switching of copper doped MoOx films for nonvolatile memory applications. Appl.
Phys. Lett. 90, 122104 (2007)
Son, J.Y., Shin, Y.-H.: Direct observation of conduction filaments on resistive switching of NiO thin films. Appl. Phys. Lett. 92, 222106 (2008)
Muenstermann, R., Dittmann, R., Szot, K., Mi, S., Jia, C.-L., Meuffels, P., Waser, R.:
Realization of regular arrays of nanoscale resistive switching blocks in thin films of
Nb-doped SrTiO3. Appl. Phys. Lett. 93, 023110 (2008)
Liu, Q., Dou, C., Wang, Y., Long, S., Wang, W., Liu, M., Zhang, M., Chen, J.: Formation of multiple conductive filaments in the Cu/ZrO2:Cu/Pt device. Appl. Phys. Lett.
95, 023501 (2009)
Liu, M., Abid, Z., Wang, W., He, X., Liu, Q., Guan, W.: Multilevel resistive switching
with ionic and metallic filaments. Appl. Phys. Lett. 94, 233106 (2009)
Yang, J.J., Pickett, M.D., Li, X., Ohlberg, D.A.A., Stewart, D.R., Williams, R.S.:
Memristive switching mechanism for metal/oxide/metal nanodevices. Nat. Nanotechnol. 3, 429 (2008)
Janousch, M., Meijer, G.I., Staub, U., Delley, B., Karg, S.F., Andreasson, B.P.: Role of
Oxygen Vacancies in Cr-Doped SrTiO3 for Resistance-Change Memory. Adv. Mater.
19, 2232 (2007)
Park, S.-G., Kope, B.M., Nishi, Y.: First-principles study of resistance switching in rutile TiO2 with Oxygen vacancy. NVMTS 2008. 9th Annual, 1 (2008)
Jung, K., Seo, H., Kim, Y., Im, H., Hong, J., Park, J.-W., Lee, J.-K.: Temperature dependence of high- and low-resistance bistable states in polycrystalline NiO films.
Appl. Phys. Lett. 90, 052104 (2007)
Lee, M.J., Park, Y., Ahn, S.E., Kang, B.S., Lee, C.B., Kim, K.H., Xianyu, W.X., Lee,
J.H., Chung, S.J., Kim, Y.H., Lee, C.S., Choi, K.N., Chung, K.S.: Comparative tructural and electrical analysis of NiO and Ti doped NiO as materials for resistance random access memory. J. Appl. Phys. 103, 013706 (2008)
Park, G.-S., Li, X.-S., Kim, D.-C., Jung, R.-J., Lee, M.-J., Se, S.: Observation of electric-field induced Ni filament channels in polycrystalline NiOx film. Appl. Phys. Lett.
91, 222103 (2007)
Mott, N.F., Davis, E.A.: Electronic Processes in Non-Crystalline Materials. Clarendon
Press, Oxford (1979)
Ho, C.H., Lai, E.K., Lee, M.D., Pan, C.L., Yao, Y.D., Hsieh, K.Y., Liu, R., Lu, C.Y.: A
Highly Reliable Self-Aligned Graded Oxide WOx Resistance Memory: Conduction
Mechanisms and Reliability. 2007 IEEE Symp. VLSI Technol., 228 (2007)
Liu, Q., Guan, W., Long, S., Liu, M., Zhang, S., Wang, Q., Chen, J.: Resistance
switching of Au-implanted-ZrO2 film for nonvolatile memory application. J. Appl.
Phys. 104, 114514 (2008)
Xu, N., Liu, L.F., Sun, X., Liu, X.Y., Han, D.D., Wang, Y., Han, R.Q., Kang, J.F., Yu,
B.: Characteristics and mechanism of conduction.set process in TiN/ZnO/Pt resistance
switching random-access memories. Appl. Phys. Lett. 92, 232112 (2008)
Chang, W.Y., Lai, Y.-C., Wu, T.-B., Wang, S.-F., Chen, F., Tsai, M.-J.: Unipolar resis-
42
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
74.
75.
76.
77.
78.
tiveswitchin characteristics of ZnO thin films for nonvolatile memory applications.
Appl. Phys. Lett. 92, 022110 (2008)
Kim, Y.-M., Lee, J.-S.: Reproducible resistance switching characteristics of hafnium
oxide-based nonvolatile memory devices. J. Appl. Phys. 104, 114115 (2008)
Lin, C.-Y., Wang, S.-Y., Lee, D.-Y., Tseng, T.Y.: Electrical Properties and Fatigue Behaviors of ZrO2 Resistive Switching Thin Films. J. Electrochem. Soc. 155, H615
(2008)
Wei, Z., Kanzawa, Y., Arita, K., Katoh, Y., Kawai, K., Muraoka, S., Mitani, S., Fujii,
S., Katayama, K., Iijima, M., Mikawa, T., Ninomiya, T., Miyanaga, R., Kawashima,
Y., Tsuji, K., Himeno, A., Okada, T., Azuma, R., Shimakawa, K., Sugaya, H., Takagi,
T., Yasuhara, R., Horiba, K., Kumigashira, H., Oshima, M.: Highly reliable TaOx ReRAM and direct evidence of redox reaction mechanism. IEEE Int. Electron Devices
Meeting, 2008. IEDM 2008. IEEE Int., 293 (2008)
Lee, H.Y., Chen, P.-S., Wu, T.-Y., Chen, Y.S., Chen, F., Wang, C.-C., Tzeng, P.-J., Lin,
C.H., Tsai, M.-J., Lien, C.: HfOx Bipolar Resistive Memory with Robust Endurance
Using AlCu as Buffer Electrode. IEEE Electron Device Lett. 30, 703 (2009)
Liu, Q., Guan, W., Long, S., Jia, R., Liu, M.,Chen, J.: Resistive switching memory effect of ZrO2 films with Zr+ implanted. Appl. Phys. Lett. 92, 012117 (2008)
Li, X., Tung, C.H., Pey, K.L.: The nature of dielectric breakdown. Appl. Phys. Lett.
93, 072903 (2008)
Fujiwara, K.K, Nemoto, T., Rozenberg, M.J., Nakamura, Y., Takagi, H.: Resistance
Switching and Formation of a Conductive Bridge in Metal/Binary Oxide/Metal Structure for Memory Devices. Jpn. J. Appl. Phys. 47, 6266 (2008)
Buh, G.-H., Hwang, I., Park, B.H.: Time-dependent electroforming in NiO resistive
switching devices. Appl. Phys. Lett. 95, 142101 (2009)
Yang, J.J., Miao, F., Pickett, M.D., Ohlberg, D.A.A., Stewart, D.R., Lau, C.N., Williams, R.S.: The mechanism of electroforming of metal oxide memristive switches.
Nanotechnol. 20, 215201(2009)
Lee, M.-J., Han, S., Jeon, S.H., Park, B.H., Kang, B.S., Ahn, S.-E., Kim, K.H., Lee,
C.B., Kim, C.J., Yoo, I.-K., Seo, D.H., Li, X.-S., Park, J.-B., Lee, J.-H., Park, Y.: Electrical Manipulation of Nanofilaments in Transition-Metal Oxides for Resistance-Based
Memory. Nano Lett. 9,1476 (2009)
Kim, K.M., Choi, B.J., Shin, Y.C., Choi, S., Hwang, C.S.: Anode-interface localized
filamentary mechanism in resistive switching of TiO2 thin films. Appl. Phys. Lett. 91,
012907 (2007)
Baek, I.G., Lee, M.S., Seo, S., Lee, M.J., Seo, D.H., Suh, D.-S., Park, J.C., Park, S.O.,
Kim, H.S., Yoo, I.K., Chung, U.-I., Moon, I.T.: Highly scalable nonvolatile resistive
memory using simple binary oxide driven by asymmetric unipolar voltage pulses.
Electron Devices Meet., 2004. IEDM Tech. Dig. IEEE Int., 587 (2004)
Inoue, I.H., Yasuda, S., Akinaga, H., Takagi, H.: Nonpolar resistance switching of
metal/binary-transition-metal
oxides/
metal
sandwiches:
Homogeneous/inhomogeneous transition of current distribution. Phys. Rev. B 77, 035105 (2008)
Walcyzk, C., Wenger, C., Sohal, R., Lukosius, M., Fox, A., DÄ…browski, J., Wolansky,
D., Tillack, B., Mussig, H.-J., Schroeder, T.: Pulse-induced low-power resistive
switching in HfO2 metal0insulator-metal diodes for nonvolatile memory applications.
J. Appl. Phys. 105, 114103 (2009)
Hsiung, C.-P., Gan, J.-Y., Tseng, S.-H., Tai, N.-H., Tzeng, P.-J., Lin, C.-H., Chen, F.,
Tsai, M.-J.: Resistance Switching Characteristics of TiO2 Thin Films Prepared with
Reactive Sputtering. Electrochem. Solid-State Lett. 12, G31 (2009)
Cao, X., Li, X., Gao, X., Yu, W., Liu, X., Zhang, Y., Chen, L., Cheng, X.: Formingfree colossal resistive switching effect in rare-earth-oxide Gd2O3 films for memristor
applications. J. Appl. Phys. 106, 073723 (2009)
Goux, L., Lisoni, J.G., Wang, X.P., Jurczak, M., Wouters, D.J.: Optimized Ni Oxidation
43
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
89.
90.
91.
92.
93.
94.
95.
96.
in 80-nm Contact Holes for Integration of Forming-Free and Low-Power Ni/NiO/Ni
Memory Cells. IEEE Trans. Electron Devices 56,2363 (2009)
Yang, M.K., Park, J.-W., Ko, T.K., Lee, J.-K.: Bipolar resistive switching behavior in
TiMnO2/Pt structure for nonvolatile memory devices. Appl. Phys. Lett. 95, 042105
(2009)
Yoshida, C., Kinoshita, K., Yamasaki, T., Sugiyama, Y.: Direct observation of oxygen
movement during resistance switching in NiO/Pt film. Appl. Phys. Lett. 93, 042106
(2008)
Jeong, H.Y., Lee, J.Y., Choi, S.-Y., Kim, J.W.: Microscopic origin of bipolar resistive
switching of nanoscale titanium oxide thin films. Appl. Phys. Lett. 95, 162108 (2009)
Fujimoto, M., Koyama, H., Konagai, M., Hosoi, Y., Ishihara, K., Ohnishi, S., Awaya,
N.: TiO2 anatase nanolayer on TiN thin film exhibiting high-speed bipolar resistive
switching. Appl. Phys. Lett. 89, 223509 (2006)
Xu, N., Gao, B., Liu, L.F., Sun, B., Liu, X.Y., Han, R.Q., Kang, J.F., Yu, B.: A unified
physical model of switching behavior in oxide-based RRAM. 2008 Symp. VLSI Technol., 100 (2008)
Bruchhaus, R., Waser, R.: Bipolar Resistive Switching in Oxides for Memory. In: Ramanathan, S. (ed.) Thin Film Metal-Oxides: Fundamentals and Applications in Electronics and Energy, pp. 131-168, Springer, Heidelberg (2009)
Strukov, D.B., Snider, G.S., Stewart, D.R., Williams, R.S.: The missing memristor
found. Nat. 453, 80 (2008)
Strukov, D.B., Borghetti, J.L., Williams, R.S.: Coupled Ionic and Electronic Transport
Model of Thin-Film Semiconductor Memristive Behavior. Small 5, 1058 (2009)
Pickett, M.D., Strukov, D.B., Borghetti, J.L., Yang, J.J., Snider, G.S., Stewart, D.R.,
Williams, R.S.: Switching dynamics in titanium dioxide memristive devices. J. Appl.
Phys. 106, 074508 (2009)
Russo, U., Ielmini, D., Cagli, C., Lacaita, A.L.: Filament Conduction and Reset Mechanism in NiO-Based Resistive-Switching Memory (RRAM) Devices. IEEE Trans. Electron Devices 56, 186 (2009)
Russo, U., Ielmini, D., Cagli, C., Lacaita, A.L.: Self-Accelerated Thermal Dissolution
Model for Reset Programming in Unipolar Resistive-Switching Memory (RRAM) Devices. IEEE Trans. Electron Devices 56, 193 (2009)
Cagli, C., Nardi, F., Ielmini, D.: Modeling of Set/Reset Operations in NiO-Based Resistive-Switching Memory Devices. IEEE Trans. Electron Devices 56, 1712 (2009)
Jeong, D.S., Choi, B.J., Hwang, C.S.: Study of the negative resistance phenomenon in
transition metal oxide films from a statistical mechanics point of view. J. Appl. Phys.
100, 113724 (2006)
Seo, S., Lee, M.J., Kim, D.C., Ahn, S.E., Park, B.-H., Kim, Y.S., Yoo, I.K., Byun, I.S.,
Hwang, I.R., Kim, S.H., Kim, J.-S., Choi, J.S., Lee, J.H., Jeon, S.H., Hong, S.H., Park,
B.H.: Electrode dependence of resistance switching in polycrystalline NiO films. Appl.
Phys. Lett. 87, 263507 (2005)
Lee, C.B., Kang, B.S., Lee, M.J., Ahn, S.E., Stefanovich, G., Xianyu, W.X., Kim,
K.H., Hur, J.H., Yin, H.X., Park, Y., Yoo, I.K., Park, J.-B., Park, B.H.: Electromigration effect of Ni Electrodes on the resistive switching characteristics of NiO thin films.
Appl. Phys. Lett. 91, 082104 (2007)
Lee, C.B., Kang, B.S., Benayad, A., Lee, M.J., Ahn, S.-E., Kim, K.H., Stefanovich, G.,
Park, Y., Yoo, I.K.: Effects of metal electrodes on the resistive memory switching property of NiO thin films. Appl. Phys. Lett. 93, 042115 (2008)
Xu, N., Liu, L.F., Sun, X., Chen, C., Wang, Y., Han, D.D., Liu, X.Y., Han, R.Q., Kang,
J.F., Yu, B.: Bipolar switching behavior in TiN/ZnO/Pt resistive nonvolatile memory
with fast switching and long retention. Semicond. Sci. Technol. 23, 075019 (2008)
Choi, J.S., Kim, J.-S., Hwang, I.R., Hong, S.H., Jeon, S.H., Kang, S.-O., Park, B.H.,
Kim, D.C., Lee, M.J., Seo, S.: Different resistance switching behaviors of NiO thin
44
films deposited on Pt and SrRuO3 electrodes. Appl. Phys. Lett. 95, 022109 (2009)
97. Lin, C.-Y., Wu, C.-Y., Wu, C.-Y., Lee, T.-C., Yang, F.-L., Hu, C., Tseng, T.-Y.: Effect of
Top Electrode Material on Resistive Switching Properties of ZrO 2 Film Memory Devices. IEEE Electron Device Lett. 28, 366 (2007)
98. Yoshida, C., Tsunoda, K., Noshiro, H., Sugiyama, Y.: High speed resistive switching in
Pt/TiO2/TiN film for nonvolatile memory application. Appl. Phys. Lett. 91, 223510
(2007)
99. Gao, B., Zhang, H.W., Yu, S., Sun, B., Liu, L.F., Liu, X.Y., Wang, Y., Han, R.Q.,
Kang, J.F., Yu, B., Wang, Y.Y.: Oxide-based RRAM: Uniformity improvement using a
new material-oriented methodology. 2009 Symp. VLSI Technol., 30 (2009)
100. Lin, C.-Y., Wu, C.-Y., Wu, C.-Y., Tseng, T.Y., Hu, C.: Modified resistive switching behavior of ZrO2 memory films based on the interface layer formed by using Ti top electrode. J. Appl. Phys. 102, 094101 (2007)
101. Zhou, P., Yin, M., Wan, H.J., Lu, H.B., Tang, T.A., Lin, Y.Y.: Role of TaON interface
for CuxO resistive switching memory based on a combined model. Appl. Phys. Lett.
94, 053510 (2009)
102. Yu, S., Wong, H.-S.P.: A phenomenological model of oxygen ion transport for metal
oxide resistive switching memory. Memory Workshop (IMW) 2010 IEEE Int. (2010)
103. Sun, X., Sun, B., Liu, L.F., Xu, N., Liu, X.Y., Han, R.Q., Kang, J.F., Xiong, G., Ma,
T.P.: Wavelength tuning of GaAs LED’s through surface effects. IEEE Electron Device
Lett. 30, 334 (2009)
104. Sun, B., Liu, Y.X., Liu, L.F., Xu, N., Wang, Y., Liu, X.Y., Han, R.Q., Kang, J.F.: Highly uniform resistive switching characteristics of TiN/ZrO2/Pt memory devices. J. Appl.
Phys. 105, 061630 (2009)
105. Kim, S.I., Lee, J.H., Chang, Y.W., Hwang, S.S., Yoo, K.-H.: Reversible resistive
switching behaviors in NiO nanowires. Appl. Phys. Lett. 93, 033503 (2008)
106. Oka, K., Yanagida, T., Nagashima, K., Tanaka, H., Kawai, T.: Nonvolatile Bipolar Resistive Memory Switching in Single Crystalline NiO Heterostructured Nanowires. J.
Am. Chem. Soc. 131, 3434 (2009)
107. Seo, J.W., Park, J.-W., Lim, K.S., Yang, J.-H., Kang, S.J.: Transparent resistive random access memory and its characteristics for nonvolatile resistive switching. Appl.
Phys. Lett. 93, 223505 (2008)
108. Kim, S., Choi, Y.-K: Resistive switching of aluminum oxide for flexible memory. Appl.
Phys. Lett. 92, 223508 (2008)
109. Kim, S., Moon, H., Gupta, D., Yoo, S., Choi, Y.-K.: Resistive Switching Characteristics of Sol-Gel Zinc Oxide Films for Flexible Memory Applications. IEEE Trans. Electron Devices 56, 696 (2009)
110. Yun, D.K., Kim, K.-D., Kim, S., Lee, J.-H., Park, H.-H., Jeong, J.-H., Choi Y.-K.,
Choi, D.-G.: Mass fabrication of resistive random access crossbar arrays by step and
flash impring lithography. Nanotechnol. 20, 445305 (2009)
111. Lee, B., Wong, H.-S.P.: NiO resistance change memory with a novel structure for 3D
integration and improved confinement of conduction path. VLSI Technol., 209 Symp.,
28 (2009)
112. Wu, Y., Lee, B., Wong, H.-S.P.: Ultra-low power Al2O3-based RRAM with 1µA reset
current. 2010 Int. Symp. VLSI Technol. Syst. App. (VLSI-TSA), 136 (2010)
113. Ahn, S.-E., Lee, M.-J., Park, Y., Kang, B.S., Lee, C.B., Kim, K.H., Seo, S., Suh, D.-S.,
Kim, D.-C., Hur, J., Xianyu, W., Stefanovich, G., Yin, H., Yoo, I.-K., Lee, J.-H., Park,
J.-B., Baek, I.-G., Park, B.H.: Write Current Reduction in Transition Metal Oxide
Based Resistance Change Memory. Adv. Mater. 20, 924 (2008)
114. Kinoshita, K., Tsunoda, K., Sato, Y., Noshiro, H., Yagaki, S., Aoki, M., Sugiyama, Y.:
Reduction in the reset current in a resistive random access memory consisting of NiO x
brought about by reducing a parasitic capacitance. Appl. Phys. Lett. 93, 033506
(2008)
45
115. Chen, A., Haddad, S., Wu, Y.-C., Fang, T.-N., Lan, Z., Avanzino, S., Pangrle, S., Buynoski, M., Rathor, M., Cai, W., Tripsas, N., Bill, C., VanBuskirk, M., Taguchi, M.:
Tech. Dig. IEDM, 765 (2005)
116. Chen, A., Haddad, S., Wu, Y.C., Fang, T.N., Kaza, S., Z. Lan: Erasing characteristics
of Cu2O metal-insulator-metal resistive switching memory. Appl. Phys. Lett. 92,
013503 (2008)
117. Sato, Y., Tsunoda, K., Kinoshita, K., Noshiro, H., Aoki, M., Sugiyama, Y.: Sub-100µA
Reset Current of Nickel Oxide Resistive Memory through Control of Filamentary Conductance by Current Limit of MOSFET. IEEE Trans. Electron Devices 55, 1185 (2008)
118. Ielmini, D., Cagli, C., Nardi, F.: Resistance transition in metal oxides induced by electronic threshold switching. Appl. Phys. Lett. 94, 063511 (2009)
119. Chen, A., Haddad, S., Wu, Y.C., Z. Lan, Fang, T.N., Kaza, S.: Switching characteristics of Cu2O metal-insulator-metal resistive memory. Appl. Phys. Lett. 91, 123517
(2007)
120. Choi, B., J., Choi, S., Kim, K.M., Shin, Y.C., Hwang, C.S., Hwang, S.-Y., Cho, S.-S.,
Park, S., Hong, S.-K.: Study on the resistive switching time of TiO2 thin films. Appl.
Phys. Lett. 89, 012906 (2006)
121. Tsunoda, K., Kinoshita, K., Noshiro, H., Yamazaki, Y., Iizuka, T., Ito, Y., Takahashi,
A., Okano, A., Sato, Y., Fukano, T., Aoki, M., Sugiyama, Y.: Low Power and High
Speed Switching of Ti-doped NiO ReRAM under the Unipolar Voltage Source of less
than 3 V. Electron Devices Meet., 2007. IEDM 2007. IEEE Int., 767 (2007)
122. Lv, H.B., Yin, M., Fu, X.F., Song, Y.L., Tang, L., Zhou, P., Zhao, C.H., Tang, T.A.,
Chen, B.A., Lin, Y.Y.: Resistive Memory Switching of CuxO Films for a Nonvolatile
Memory Application. IEEE Electron Device Lett. 29, 309 (2008)
123. Chen, A., Haddad, S., Wu, Y.-C.: A Temperature-Accelerated Method to Evaluate Data
Retention of Resistive Switching Nonvolatile Memory. IEEE Electron Device Lett. 29,
38 (2008)
124. Kim, D.C., Lee, M.J., Ahn, S.E., Seo, S., Park, J.C., Yoo, I.K. Baek, I.G., Kim, H.J.,
Yim, E.K., Lee, J.E., Park, S.O., Kim, H.S., Chung, U-I., Moon, J.T., Ryu, B.I.: Improvement of resistive memory switching in NiO using IrO 2. Appl. Phys. Lett. 88,
232106 (2006)
125. Chang, W.-Y., Cheng, K.-J., Tsai, J.-M., Chen, H.-J., Chen, F., Tsai, M.-J., Wu, T.-B.:
Improvement of resistive switching characteristics in TiO2 thin films with embedded Pt
nanocrystals. Appl. Phys. Lett. 95, 042104 (2009)
126. Yu, S., Gao, B., Dai, H.B., Sun, B., Liu, L.F., Liu, X.Y., Han, R.Q., Kang, J.F., Yu, B.:
Improved Uniformity of Resistive Switching Behaviors in HfO2 Thin Films with Embedded Al Layers. Electrochem. Solid-State Lett., 13, H36 (2010)
127. Liu, Q., Long, S., Wang, W., Zuo, Q., Zhang, S., Chen, J., Liu, M.: Improvement of
Resistive Switching Properties in ZrO2-Based ReRAM with Implanted Ti Ions. IEEE
Electron Device Lett. 30,1335 (2009)
128. Yoo, I.K., Kang, B.S., Park, Y.D., Lee, M.J., Park, Y.: Interpretation of nanoscale conducting paths and their control in nickel oxide (NiO) thin films. Appl. Phys. Lett. 92,
202112 (2008)
129. Yin, M., Zhou, Lv, H.B., Xu, J., Song, Y.L., Fu, X.F., Tang, T.A., B. Chen, A., Lin,
Y.Y.: Improvement of Resistive Switching in CuxO Using New RESET Mode. IEEE
Electron Device Lett. 29, 681 (2008)
130. Liang, J., Wong, H.-S.P.: Size limitation of cross-point memory array and its dependence on data storage pattern and device parameters. Interconnect Technol. Conf.
(IITC), 2010 Int. (2010)
131. Lee, M.-J., Seo, S., Kim, D.-C., Ahn, S.-E., Seo, D.H., Yoo, I.-K., Baek, I.-G., Kim,
D.-S., Byun, I.-S., Kim, S.-H., Hwang, I.-R., Kim, J.-S., Jeon, S.-H., Park, B.H.: A
Low-Temperature-Grown Oxide Diode as a New Switch Element for High-Density
Nonvolatile Memories. Adv. Mater. 19, 73 (2007)
46
132. Baek, I.G., Kim, D.C., Lee, M.J., Kim, H.-J., Yim, E.K., Lee, M.S., Lee, J.E., Ahn,
S.E., Seo, S., Lee, J.H., Park, J.C., Cha, Y.K., Park, S.O., Kim, H.S., Yoo, I.K., Chung,
U-I., Moon J.T., Ryu, B.I.: Multi-layer cross-point binary oxide resistive memory (OxRRAM) for post-NAND storage application. Electron Devices Meet., 2005. IEDM
Tech. Dig. IEEE Int., 750 (2005)
133. Lee, M.-J., Park, Y., Kang, B.-S., Ahn, S.-E., Lee, C., Kim, K., Xianyu, W., Stefanovich, G., Lee, J.-H., Chung, S.-J., Kim, Y.-H., Lee, C.-S., Park, J.-B., Baek, I.-G.,
Yoo, I.-K.: 2-stack 1D-1R Cross-point Structure with Oxide Diodes as Switch Elements for High Density Resistance RAM Applications. Electron Devices Meet., 2007.
IEDM 207. IEEE Int., 771 (2007)
134. Lee, M.-J., Lee, C.B., Kim, S., Yin, H., Park, J., Ahn, S.E., Kang, B.S., Kim, K.H.,
Stefanovich, G., Song, I., Kim, S.-W., Lee, J.H., Chung, S.J., Kim, Y.H., Lee, C.S.,
Park, J.B., Baek, I.G., Kim, C.J., Park, Y.: Stack friendly all-oxide 3D RRAM using
GaInZnO peripheral TFT realized over glass substrates. Electron Devices Meet.,
2008, IEDM 2008, IEEE Int., 85 (2008)
135. Yoon, H.S., Baek, I.-G., Zhao, J., Sim, H., Park, M.Y., Lee, H., Oh, G.-H., Shin, J.C.,
Yeo, I.-S., Chung, U.-I.: Vertical cross-point resistance change memory for ultra-high
density non-volatile memory applications. 2009 Symp. VLSI Technol., 26 (2009)
136. Ohya, Y., Uedo, M., Takahasi, Y.: Oxide Thin Film Diode Fabricated by Liquid-Phase
Method. Jpn. J. Appl. Phys., 35, 4738 (1996)
137. Lee, W.Y., Mauri D., Hwang, C.: High-current-density ITOx/NiOx thin-film diodes.
Appl. Phys. Lett., 72, 1584 (1998)
138. Shin, Y.C., Song, J., Kim, K.M., Choi, B.J., Choi, S., Lee, H.J., Kim, G.H., Eom, T.,
Hwang, C.S.: (In,Sn)2O3/TiO2/Pt Schottky-type diode switch for the TiO2 resistive
switching memory array. Appl. Phys. Lett. 92, 162904 (2008)
139. Okimura, K.: J. Sakai: Time-dependent Characteristics of Electric Field-induced Metal-Insulator Transtion of Planer VO2/c-Al2O3 Structure. Jpn. J. Appl. Phys., 46, L813
(2007)
140. Lee, M.-J., Park, Y., Suh, D.-S., Lee, E.-H., Seo, S., Kim, D.-C., Jung, R., Kang, B.-S.,
Ahn, S.-E., Lee, C.B., Seo, D.H., Cha, Y.-K., Yoo, I.-K., Kim, J.-S., Park, B.H.: Two
Series Oxide Resistors Applicable to High Speed and High Density Nonvolatile
Memory. Adv. Mater. 19, 3919 (2007)
141. Lee, M.-J., Kim, S.I., Lee, C.B., Yin, H., Ahn, S.-E., Kang, B.S., Kim, K.H., Park,
J.C., Kim, C.J., Song, I., Kim, S.W., Stefanovich, G., Lee, J.H., Chung, S.J., Kim,
Y.H., Park, Y.: Low-Temperature-Grown Transition Metal Oxide Based Storage Materials and Oxide Transitstors for High-Density Non-volatile Memory. Adv. Funct. Mater. 19, 1587 (2009)
Download