Memristor

A memristor (/ˈmɛmrɪstər/; a portmanteau of memory resistor) is a non-linear two-terminal electrical component relating electric charge and magnetic flux linkage. It was described and named in 1971 by Leon Chua, completing a theoretical quartet of fundamental electrical components which comprises also the resistor, capacitor and inductor.[1] No physical memristor component has yet been demonstrated.

Memristor
InventedLeon Chua (1971)
Electronic symbol

Chua and Kang later generalized the concept to memristive systems.[2] Such a system comprises a circuit, of multiple conventional components, which mimics key properties of the ideal memristor component and is also commonly referred to as a memristor. Several such memristor system technologies have been developed, notably ReRAM.

The identification of genuine memristive properties in both theoretical and practical devices has attracted controversy.

The memristor as a fundamental electrical component

Conceptual symmetries of resistor, capacitor, inductor, and memristor.

Chua in his 1971 paper identified a theoretical symmetry between the non-linear resistor (voltage vs. current), non-linear capacitor (voltage vs. charge), and non-linear inductor (magnetic flux linkage vs. current). From this symmetry he inferred the characteristics of a fourth fundamental non-linear circuit element, linking magnetic flux and charge, which he called the memristor. In contrast to a linear (or non-linear) resistor the memristor has a dynamic relationship between current and voltage including a memory of past voltages or currents. Other scientists had proposed dynamic memory resistors such as the memistor of Bernard Widrow, but Chua introduced a mathematical generality.

Derivation and characteristics

The memristor was originally defined in terms of a non-linear functional relationship between magnetic flux linkage Φm(t) and the amount of electric charge that has flowed, q(t):[1]

The magnetic flux linkage, Φm, is generalized from the circuit characteristic of an inductor. It does not represent a magnetic field here. Its physical meaning is discussed below. The symbol Φm may be regarded as the integral of voltage over time.[3]

In the relationship between Φm and q, the derivative of one with respect to the other depends on the value of one or the other, and so each memristor is characterized by its memristance function describing the charge-dependent rate of change of flux with charge.

Substituting the flux as the time integral of the voltage, and charge as the time integral of current, the more convenient forms are;

To relate the memristor to the resistor, capacitor, and inductor, it is helpful to isolate the term M(q), which characterizes the device, and write it as a differential equation.

DeviceCharacteristic property (units)Differential equation
Resistor (R)Resistance (V / A, or ohm, Ω)R = dV / dI
Capacitor (C)Capacitance (C / V, or farad)C = dq / dV
Inductor (L)Inductance (Wb / A, or henry)L = dΦm / dI
Memristor (M)Memristance (Wb / C, or ohm)M = dΦm / dq

The above table covers all meaningful ratios of differentials of I, q, Φm, and V. No device can relate dI to dq, or m to dV, because I is the derivative of q and Φm is the integral of V.

It can be inferred from this that memristance is charge-dependent resistance. If M(q(t)) is a constant, then we obtain Ohm's Law R(t) = V(t)/I(t). If M(q(t)) is nontrivial, however, the equation is not equivalent because q(t) and M(q(t)) can vary with time. Solving for voltage as a function of time produces

This equation reveals that memristance defines a linear relationship between current and voltage, as long as M does not vary with charge. Nonzero current implies time varying charge. Alternating current, however, may reveal the linear dependence in circuit operation by inducing a measurable voltage without net charge movement—as long as the maximum change in q does not cause much change in M.

Furthermore, the memristor is static if no current is applied. If I(t) = 0, we find V(t) = 0 and M(t) is constant. This is the essence of the memory effect.

Analogously, we can define a as menductance.[1]

The power consumption characteristic recalls that of a resistor, I2R.

As long as M(q(t)) varies little, such as under alternating current, the memristor will appear as a constant resistor. If M(q(t)) increases rapidly, however, current and power consumption will quickly stop.

M(q) is physically restricted to be positive for all values of q (assuming the device is passive and does not become superconductive at some q). A negative value would mean that it would perpetually supply energy when operated with alternating current.

Modelling and validation

In order to understand the nature of memristor function, some knowledge of fundamental circuit theoretic concepts is useful, starting with the concept of device modelling.[4]

Engineers and scientists seldom analyze a physical system in its original form. Instead, they construct a model which approximates the behaviour of the system. By analyzing the behaviour of the model, they hope to predict the behaviour of the actual system. The primary reason for constructing models is that physical systems are usually too complex to be amenable to a practical analysis.

In the 20th century, work was done on devices where researchers did not recognize the memristive characteristics. This has raised the suggestion that such devices should be recognised as memristors.[4] Pershin and Di Ventra[5] have proposed a test that can help to resolve some of the long-standing controversies about whether an ideal memristor does actually exist or is a purely mathematical concept.

The rest of this article primarily addresses memristors as related to ReRAM devices, since the majority of work since 2008 has been concentrated in this area.

Superconducting memristor component

Dr. Paul Penfield, in a 1974 MIT technical report[6] mentions the memristor in connection with Josephson junctions. This was an early use of the word "memristor" in the context of a circuit device.

One of the terms in the current through a Josephson junction is of the form:

where is a constant based on the physical superconducting materials, is the voltage across the junction and is the current through the junction.

Through the late 20th century, research regarding this phase-dependent conductance in Josephson junctions was carried out.[7][8][9][10] A more comprehensive approach to extracting this phase-dependent conductance appeared with Peotta and DiVentra's seminal paper in 2014.[11]

Memristor circuits

Due to the practical difficulty of studying the ideal memristor, we will discuss other electrical devices which can be modelled using memristors. For a mathematical description of a memristive device (systems), see Theory.

A discharge tube can be modelled as a memristive device, with resistance being a function of the number of conduction electrons .[2]

is the voltage across the discharge tube, is the current flowing through it and is the number of conduction electrons. A simple memristance function is . and are parameters depending on the dimensions of the tube and the gas fillings. An experimental identification of memristive behaviour is the "pinched hysteresis loop" in the plane. For an experiment that shows such a characteristic for a common discharge tube, see "A physical memristor Lissajous figure" (YouTube). The video also illustrates how to understand deviations in the pinched hysteresis characteristics of physical memristors.[12][13]

Thermistors can be modelled as memristive devices.[13]

is a material constant, is the absolute body temperature of the thermistor, is the ambient temperature (both temperatures in Kelvin), denotes the cold temperature resistance at , is the heat capacitance and is the dissipation constant for the thermistor.

A fundamental phenomenon that has hardly been studied is memristive behaviour in pn-junctions.[14] The memristor plays a crucial role in mimicking the charge storage effect in the diode base, and is also responsible for the conductivity modulation phenomenon (that is so important during forward transients).

Criticisms

In 2008, a team at HP Labs claimed to have found Chua's missing memristor based on an analysis of a thin film of titanium dioxide thus connecting the operation of ReRAM devices to the memristor concept. According to the HP labs, the memristor would operate in the following way: the memristor's electrical resistance is not constant but depends on the history of current that had previously flowed through the device, i.e., its present resistance depends on how much electric charge has flowed in what direction through it in the past; the device remembers its history—the so-called non-volatility property.[15] When the electric power supply is turned off, the memristor remembers its most recent resistance until it is turned on again.[16][17]

The HP result was published in the scientific journal Nature.[16][18] Following this claim, Leon Chua has argued that the memristor definition could be generalized to cover all forms of two-terminal non-volatile memory devices based on resistance switching effects.[15] Chua also argued that the memristor is the oldest known circuit element, with its effects predating the resistor, capacitor, and inductor.[19] There are, however, some serious doubts as to whether a genuine memristor can actually exist in physical reality.[20][21][22][23][24] Additionally, some experimental evidence contradicts Chua's generalization since a non-passive nanobattery effect is observable in resistance switching memory.[25] A simple test has been proposed by Pershin and Di Ventra[5] to analyse whether such an ideal or generic memristor does actually exist or is a purely mathematical concept. Up to now, there seems to be no experimental resistance switching device (ReRAM) which can pass the test.[5][26]

These devices are intended for applications in nanoelectronic memories, computer logic, and neuromorphic/neuromemristive computer architectures.[27][28][29] In 2013, Hewlett-Packard CTO Martin Fink suggested that memristor memory may become commercially available as early as 2018.[30] In March 2012, a team of researchers from HRL Laboratories and the University of Michigan announced the first functioning memristor array built on a CMOS chip.[31]

An array of 17 purpose-built oxygen-depleted titanium dioxide memristors built at HP Labs, imaged by an atomic force microscope. The wires are about 50 nm, or 150 atoms, wide.[32] Electric current through the memristors shifts the oxygen vacancies, causing a gradual and persistent change in electrical resistance.[33]

According to the original 1971 definition, the memristor was the fourth fundamental circuit element, forming a non-linear relationship between electric charge and magnetic flux linkage. In 2011, Chua argued for a broader definition that included all 2-terminal non-volatile memory devices based on resistance switching.[15] Williams argued that MRAM, phase-change memory and ReRAM were memristor technologies.[34] Some researchers argued that biological structures such as blood[35] and skin[36][37] fit the definition. Others argued that the memory device under development by HP Labs and other forms of ReRAM were not memristors, but rather part of a broader class of variable-resistance systems,[38] and that a broader definition of memristor is a scientifically unjustifiable land grab that favored HP's memristor patents.[39]

In 2011, Meuffels and Schroeder noted that one of the early memristor papers included a mistaken assumption regarding ionic conduction.[40] In 2012, Meuffels and Soni discussed some fundamental issues and problems in the realization of memristors.[20] They indicated inadequacies in the electrochemical modelling presented in the Nature article "The missing memristor found"[16] because the impact of concentration polarization effects on the behavior of metal−TiO2−x−metal structures under voltage or current stress was not considered. This critique was referred to by Valov et al.[25] in 2013.

In a kind of thought experiment, Meuffels and Soni[20] furthermore revealed a severe inconsistency: If a current-controlled memristor with the so-called non-volatility property[15] exists in physical reality, its behavior would violate Landauer's principle of the minimum amount of energy required to change "information" states of a system. This critique was finally adopted by Di Ventra and Pershin[21] in 2013.

Within this context, Meuffels and Soni[20] pointed to a fundamental thermodynamic principle: Non-volatile information storage requires the existence of free-energy barriers that separate the distinct internal memory states of a system from each other; otherwise, one would be faced with an "indifferent" situation, and the system would arbitrarily fluctuate from one memory state to another just under the influence of thermal fluctuations. When unprotected against thermal fluctuations, the internal memory states exhibit some diffusive dynamics, which causes state degradation.[21] The free-energy barriers must therefore be high enough to ensure a low bit-error probability of bit operation.[41] Consequently, there is always a lower limit of energy requirement – depending on the required bit-error probability – for intentionally changing a bit value in any memory device.[41][42]

In the general concept of memristive system the defining equations are (see Theory):

where u(t) is an input signal, and y(t) is an output signal. The vector x represents a set of n state variables describing the different internal memory states of the device. is the time-dependent rate of change of the state vector x with time.

When one wants to go beyond mere curve fitting and aims at a real physical modeling of non-volatile memory elements, e.g., resistive random-access memory devices, one has to keep an eye on the aforementioned physical correlations. To check the adequacy of the proposed model and its resulting state equations, the input signal u(t) can be superposed with a stochastic term ξ(t), which takes into account the existence of inevitable thermal fluctuations. The dynamic state equation in its general form then finally reads:

where ξ(t) is, e.g., white Gaussian current or voltage noise. On base of an analytical or numerical analysis of the time-dependent response of the system towards noise, a decision on the physical validity of the modeling approach can be made, e.g., would the system be able to retain its memory states in power-off mode?

Such an analysis was performed by Di Ventra and Pershin[21] with regard to the genuine current-controlled memristor. As the proposed dynamic state equation provides no physical mechanism enabling such a memristor to cope with inevitable thermal fluctuations, a current-controlled memristor would erratically change its state in course of time just under the influence of current noise.[21][43] Di Ventra and Pershin[21] thus concluded that memristors whose resistance (memory) states depend solely on the current or voltage history would be unable to protect their memory states against unavoidable Johnson–Nyquist noise and permanently suffer from information loss, a so-called "stochastic catastrophe". A current-controlled memristor can thus not exist as a solid-state device in physical reality.

The above-mentioned thermodynamic principle furthermore implies that the operation of 2-terminal non-volatile memory devices (e.g. "resistance-switching" memory devices (ReRAM)) cannot be associated with the memristor concept, i.e., such devices cannot by itself remember their current or voltage history. Transitions between distinct internal memory or resistance states are of probabilistic nature. The probability for a transition from state {i} to state {j} depends on the height of the free-energy barrier between both states. The transition probability can thus be influenced by suitably driving the memory device, i.e., by "lowering" the free-energy barrier for the transition {i} → {j} by means of, for example, an externally applied bias.

A "resistance switching" event can simply be enforced by setting the external bias to a value above a certain threshold value. This is the trivial case, i.e., the free-energy barrier for the transition {i} → {j} is reduced to zero. In case one applies biases below the threshold value, there is still a finite probability that the device will switch in course of time (triggered by a random thermal fluctuation), but – as one is dealing with probabilistic processes – it is impossible to predict when the switching event will occur. That is the basic reason for the stochastic nature of all observed resistance-switching (ReRAM) processes. If the free-energy barriers are not high enough, the memory device can even switch without having to do anything.

When a 2-terminal non-volatile memory device is found to be in a distinct resistance state {j}, there exists therefore no physical one-to-one relationship between its present state and its foregoing voltage history. The switching behavior of individual non-volatile memory devices thus cannot be described within the mathematical framework proposed for memristor/memristive systems.

An extra thermodynamic curiosity arises from the definition that memristors/memristive devices should energetically act like resistors. The instantaneous electrical power entering such a device is completely dissipated as Joule heat to the surrounding, so no extra energy remains in the system after it has been brought from one resistance state xi to another one xj. Thus, the internal energy of the memristor device in state xi, U(V, T, xi), would be the same as in state xj, U(V, T, xj), even though these different states would give rise to different device's resistances, which itself must be caused by physical alterations of the device's material.

Other researchers noted that memristor models based on the assumption of linear ionic drift do not account for asymmetry between set time (high-to-low resistance switching) and reset time (low-to-high resistance switching) and do not provide ionic mobility values consistent with experimental data. Non-linear ionic-drift models have been proposed to compensate for this deficiency.[44]

A 2014 article from researchers of ReRAM concluded that Strukov's (HP's) initial/basic memristor modelling equations do not reflect the actual device physics well, whereas subsequent (physics-based) models such as Pickett's model or Menzel's ECM model (Menzel is a co-author of that article) have adequate predictability, but are computationally prohibitive. As of 2014, the search continues for a model that balances these issues; the article identifies Chang's and Yakopcic's models as potentially good compromises.[45]

Martin Reynolds, an electrical engineering analyst with research outfit Gartner, commented that while HP was being sloppy in calling their device a memristor, critics were being pedantic in saying that it was not a memristor.[46]

In the article "The Missing Memristor has Not been Found", published in Scientific Reports in 2015 by Vongehr and Meng,[23] it was shown that the real memristor defined in 1971 is not possible without using magnetic induction. This was illustrated by constructing a mechanical analog of the memristor and then analytically showing that the mechanical memristor cannot be constructed without using an inertial mass. As it is well known that the mechanical equivalent of an electrical inductor is mass, it proves that memristors are not possible without using magnetic induction. Thus, it can be argued that the variable-resistance devices, such as the ReRAMs, and the conceptual memristors may have no equivalence at all.[23][47]

Experimental tests for memristors

Chua suggested experimental tests to determine if a device may properly be categorized as a memristor:[2]

  • The Lissajous curve in the voltage-current plane is a pinched hysteresis loop when driven by any bipolar periodic voltage or current without respect to initial conditions.
  • The area of each lobe of the pinched hysteresis loop shrinks as the frequency of the forcing signal increases.
  • As the frequency tends to infinity, the hysteresis loop degenerates to a straight line through the origin, whose slope depends on the amplitude and shape of the forcing signal.

According to Chua[48][49] all resistive switching memories including ReRAM, MRAM and phase-change memory meet these criteria and are memristors. However, the lack of data for the Lissajous curves over a range of initial conditions or over a range of frequencies complicates assessments of this claim.

Experimental evidence shows that redox-based resistance memory (ReRAM) includes a nanobattery effect that is contrary to Chua's memristor model. This indicates that the memristor theory needs to be extended or corrected to enable accurate ReRAM modeling.[25]

Theory of memristor systems

In 2008 researchers from HP Labs introduced a model for a memristance function based on thin films of titanium dioxide.[16] For RON ≪ ROFF the memristance function was determined to be

where ROFF represents the high resistance state, RON represents the low resistance state, μv represents the mobility of dopants in the thin film, and D represents the film thickness. The HP Labs group noted that "window functions" were necessary to compensate for differences between experimental measurements and their memristor model due to non-linear ionic drift and boundary effects.

Operation as a switch

For some memristors, applied current or voltage causes substantial change in resistance. Such devices may be characterized as switches by investigating the time and energy that must be spent to achieve a desired change in resistance. This assumes that the applied voltage remains constant. Solving for energy dissipation during a single switching event reveals that for a memristor to switch from Ron to Roff in time Ton to Toff, the charge must change by ΔQ = QonQoff.

Substituting V = I(q)M(q), and then ∫dq/V = ∆Q/V for constant VTo produces the final expression. This power characteristic differs fundamentally from that of a metal oxide semiconductor transistor, which is capacitor-based. Unlike the transistor, the final state of the memristor in terms of charge does not depend on bias voltage.

The type of memristor described by Williams ceases to be ideal after switching over its entire resistance range, creating hysteresis, also called the "hard-switching regime".[16] Another kind of switch would have a cyclic M(q) so that each off-on event would be followed by an on-off event under constant bias. Such a device would act as a memristor under all conditions, but would be less practical.

Memristive systems

In the more general concept of an n-th order memristive system the defining equations are

where u(t) is an input signal, y(t) is an output signal, the vector x represents a set of n state variables describing the device, and g and f are continuous functions. For a current-controlled memristive system the signal u(t) represents the current signal i(t) and the signal y(t) represents the voltage signal v(t). For a voltage-controlled memristive system the signal u(t) represents the voltage signal v(t) and the signal y(t) represents the current signal i(t).

The pure memristor is a particular case of these equations, namely when x depends only on charge (x = q) and since the charge is related to the current via the time derivative dq/dt = i(t). Thus for pure memristors f (i.e. the rate of change of the state) must be equal or proportional to the current i(t) .

Pinched hysteresis

Example of pinched hysteresis curve, V versus I

One of the resulting properties of memristors and memristive systems is the existence of a pinched hysteresis effect.[50] For a current-controlled memristive system, the input u(t) is the current i(t), the output y(t) is the voltage v(t), and the slope of the curve represents the electrical resistance. The change in slope of the pinched hysteresis curves demonstrates switching between different resistance states which is a phenomenon central to ReRAM and other forms of two-terminal resistance memory. At high frequencies, memristive theory predicts the pinched hysteresis effect will degenerate, resulting in a straight line representative of a linear resistor. It has been proven that some types of non-crossing pinched hysteresis curves (denoted Type-II) cannot be described by memristors.[51]

Extended memristive systems

Some researchers have raised the question of the scientific legitimacy of HP's memristor models in explaining the behavior of ReRAM.[38][39] and have suggested extended memristive models to remedy perceived deficiencies.[25]

One example[52] attempts to extend the memristive systems framework by including dynamic systems incorporating higher-order derivatives of the input signal u(t) as a series expansion

where m is a positive integer, u(t) is an input signal, y(t) is an output signal, the vector x represents a set of n state variables describing the device, and the functions g and f are continuous functions. This equation produces the same zero-crossing hysteresis curves as memristive systems but with a different frequency response than that predicted by memristive systems.

Another example suggests including an offset value a to account for an observed nanobattery effect which violates the predicted zero-crossing pinched hysteresis effect.[25]

Implementations

Titanium dioxide memristor

Interest in the memristor revived when an experimental solid-state version was reported by R. Stanley Williams of Hewlett Packard in 2007.[53][54][55] The article was the first to demonstrate that a solid-state device could have the characteristics of a memristor based on the behavior of nanoscale thin films. The device neither uses magnetic flux as the theoretical memristor suggested, nor stores charge as a capacitor does, but instead achieves a resistance dependent on the history of current.

Although not cited in HP's initial reports on their TiO2 memristor, the resistance switching characteristics of titanium dioxide were originally described in the 1960s.[56]

The HP device is composed of a thin (50 nm) titanium dioxide film between two 5 nm thick electrodes, one titanium, the other platinum. Initially, there are two layers to the titanium dioxide film, one of which has a slight depletion of oxygen atoms. The oxygen vacancies act as charge carriers, meaning that the depleted layer has a much lower resistance than the non-depleted layer. When an electric field is applied, the oxygen vacancies drift (see Fast ion conductor), changing the boundary between the high-resistance and low-resistance layers. Thus the resistance of the film as a whole is dependent on how much charge has been passed through it in a particular direction, which is reversible by changing the direction of current.[16] Since the HP device displays fast ion conduction at nanoscale, it is considered a nanoionic device.[57]

Memristance is displayed only when both the doped layer and depleted layer contribute to resistance. When enough charge has passed through the memristor that the ions can no longer move, the device enters hysteresis. It ceases to integrate q=∫I dt, but rather keeps q at an upper bound and M fixed, thus acting as a constant resistor until current is reversed.

Memory applications of thin-film oxides had been an area of active investigation for some time. IBM published an article in 2000 regarding structures similar to that described by Williams.[58] Samsung has a U.S. patent for oxide-vacancy based switches similar to that described by Williams.[59] Williams also has a U.S. patent application related to the memristor construction.[60]

In April 2010, HP labs announced that they had practical memristors working at 1 ns (~1 GHz) switching times and 3 nm by 3 nm sizes,[61] which bodes well for the future of the technology.[62] At these densities it could easily rival the current sub-25 nm flash memory technology.

Polymeric memristor

In 2004, Krieger and Spitzer described dynamic doping of polymer and inorganic dielectric-like materials that improved the switching characteristics and retention required to create functioning nonvolatile memory cells.[63] They used a passive layer between electrode and active thin films, which enhanced the extraction of ions from the electrode. It is possible to use fast ion conductor as this passive layer, which allows a significant reduction of the ionic extraction field.

In July 2008, Erokhin and Fontana claimed to have developed a polymeric memristor before the more recently announced titanium dioxide memristor.[64]

In 2010, Alibart, Gamrat, Vuillaume et al.[65] introduced a new hybrid organic/nanoparticle device (the NOMFET : Nanoparticle Organic Memory Field Effect Transistor), which behaves as a memristor[66] and which exhibits the main behavior of a biological spiking synapse. This device, also called a synapstor (synapse transistor), was used to demonstrate a neuro-inspired circuit (associative memory showing a pavlovian learning).[67]

In 2012, Crupi, Pradhan and Tozer described a proof of concept design to create neural synaptic memory circuits using organic ion-based memristors.[68] The synapse circuit demonstrated long-term potentiation for learning as well as inactivity based forgetting. Using a grid of circuits, a pattern of light was stored and later recalled. This mimics the behavior of the V1 neurons in the primary visual cortex that act as spatiotemporal filters that process visual signals such as edges and moving lines.

Layered memristor

In 2014, Bessonov et al. reported a flexible memristive device comprising a MoOx/MoS2 heterostructure sandwiched between silver electrodes on a plastic foil.[69] The fabrication method is entirely based on printing and solution-processing technologies using two-dimensional layered transition metal dichalcogenides (TMDs). The memristors are mechanically flexible, optically transparent and produced at low cost. The memristive behaviour of switches was found to be accompanied by a prominent memcapacitive effect. High switching performance, demonstrated synaptic plasticity and sustainability to mechanical deformations promise to emulate the appealing characteristics of biological neural systems in novel computing technologies.

Atomristor

Atomristor is defined as the electrical devices showing memristive behavior in atomically thin nanomaterials or atomic sheets. In 2018, Ge and Wu et al.[70] first reported a universal memristive effect in single-layer TMD (MX2, M = Mo, W; and X = S, Se) atomic sheets based on vertical metal-insulator-metal (MIM) device structure. These atomristors offer forming-free switching and both unipolar and bipolar operation. The switching behavior is found in single-crystalline and poly-crystalline films, with various metallic electrodes (gold, silver and graphene). Atomically thin TMD sheets are prepared via CVD/MOCVD, enabling low-cost fabrication. Afterwards, taking advantage of the low "on" resistance and large on/off ratio, a high-performance zero-power RF switch is proved based on MoS2 atomristors, indicating a new application of memristors.[71]

Ferroelectric memristor

The ferroelectric memristor[72] is based on a thin ferroelectric barrier sandwiched between two metallic electrodes. Switching the polarization of the ferroelectric material by applying a positive or negative voltage across the junction can lead to a two order of magnitude resistance variation: ROFF ≫ RON (an effect called Tunnel Electro-Resistance). In general, the polarization does not switch abruptly. The reversal occurs gradually through the nucleation and growth of ferroelectric domains with opposite polarization. During this process, the resistance is neither RON or ROFF, but in between. When the voltage is cycled, the ferroelectric domain configuration evolves, allowing a fine tuning of the resistance value. The ferroelectric memristor's main advantages are that ferroelectric domain dynamics can be tuned, offering a way to engineer the memristor response, and that the resistance variations are due to purely electronic phenomena, aiding device reliability, as no deep change to the material structure is involved.

Carbon nanotube memristor

In 2013, Ageev, Blinov et al.[73] reported observing memristor effect in structure based on vertically aligned carbon nanotubes studying bundles of CNT by scanning tunneling microscope.

Later it was found[74] that CNT memristive switching is observed when a nanotube has a non-uniform elastic strain ΔL0. It was shown that the memristive switching mechanism of strained СNT is based on the formation and subsequent redistribution of non-uniform elastic strain and piezoelectric field Edef in the nanotube under the influence of an external electric field E(x,t).

Spin memristive systems

Spintronic memristor

Chen and Wang, researchers at disk-drive manufacturer Seagate Technology described three examples of possible magnetic memristors.[75] In one device resistance occurs when the spin of electrons in one section of the device points in a different direction from those in another section, creating a "domain wall", a boundary between the two sections. Electrons flowing into the device have a certain spin, which alters the device's magnetization state. Changing the magnetization, in turn, moves the domain wall and changes the resistance. The work's significance led to an interview by IEEE Spectrum.[76] A first experimental proof of the spintronic memristor based on domain wall motion by spin currents in a magnetic tunnel junction was given in 2011.[77]

Memristance in a magnetic tunnel junction

The magnetic tunnel junction has been proposed to act as a memristor through several potentially complementary mechanisms, both extrinsic (redox reactions, charge trapping/detrapping and electromigration within the barrier) and intrinsic (spin-transfer torque).

Extrinsic mechanism

Based on research performed between 1999 and 2003, Bowen et al. published experiments in 2006 on a magnetic tunnel junction (MTJ) endowed with bi-stable spin-dependent states[78](resistive switching). The MTJ consists in a SrTiO3 (STO) tunnel barrier that separates half-metallic oxide LSMO and ferromagnetic metal CoCr electrodes. The MTJ's usual two device resistance states, characterized by a parallel or antiparallel alignment of electrode magnetization, are altered by applying an electric field. When the electric field is applied from the CoCr to the LSMO electrode, the tunnel magnetoresistance (TMR) ratio is positive. When the direction of electric field is reversed, the TMR is negative. In both cases, large amplitudes of TMR on the order of 30% are found. Since a fully spin-polarized current flows from the half-metallic LSMO electrode, within the Julliere model, this sign change suggests a sign change in the effective spin polarization of the STO/CoCr interface. The origin to this multistate effect lies with the observed migration of Cr into the barrier and its state of oxidation. The sign change of TMR can originate from modifications to the STO/CoCr interface density of states, as well as from changes to the tunneling landscape at the STO/CoCr interface induced by CrOx redox reactions.

Reports on MgO-based memristive switching within MgO-based MTJs appeared starting in 2008[79] and 2009.[80] While the drift of oxygen vacancies within the insulating MgO layer has been proposed to describe the observed memristive effects,[80] another explanation could be charge trapping/detrapping on the localized states of oxygen vacancies[81] and its impact[82] on spintronics. This highlights the importance of understanding what role oxygen vacancies play in the memristive operation of devices that deploy complex oxides with an intrinsic property such as ferroelectricity[83] or multiferroicity.[84]

Intrinsic mechanism

The magnetization state of a MTJ can be controlled by Spin-transfer torque, and can thus, through this intrinsic physical mechanism, exhibit memristive behavior. This spin torque is induced by current flowing through the junction, and leads to an efficient means of achieving a MRAM. However, the length of time the current flows through the junction determines the amount of current needed, i.e., charge is the key variable.[85]

The combination of intrinsic (spin-transfer torque) and extrinsic (resistive switching) mechanisms naturally leads to a second-order memristive system described by the state vector x = (x1,x2), where x1 describes the magnetic state of the electrodes and x2 denotes the resistive state of the MgO barrier. In this case the change of x1 is current-controlled (spin torque is due to a high current density) whereas the change of x2 is voltage-controlled (the drift of oxygen vacancies is due to high electric fields). The presence of both effects in a memristive magnetic tunnel junction led to the idea of a nanoscopic synapse-neuron system.[86]

Spin memristive system

A fundamentally different mechanism for memristive behavior has been proposed by Pershin[87] and Di Ventra.[88][89] The authors show that certain types of semiconductor spintronic structures belong to a broad class of memristive systems as defined by Chua and Kang.[2] The mechanism of memristive behavior in such structures is based entirely on the electron spin degree of freedom which allows for a more convenient control than the ionic transport in nanostructures. When an external control parameter (such as voltage) is changed, the adjustment of electron spin polarization is delayed because of the diffusion and relaxation processes causing hysteresis. This result was anticipated in the study of spin extraction at semiconductor/ferromagnet interfaces,[90] but was not described in terms of memristive behavior. On a short time scale, these structures behave almost as an ideal memristor.[1] This result broadens the possible range of applications of semiconductor spintronics and makes a step forward in future practical applications.

Self-directed channel memristor

In 2017, Dr Kris Campbell formally introduced the self-directed channel (SDC) memristor.[91] The SDC device is the first memristive device available commercially to researchers, students and electronics enthusiast worldwide.[92] The SDC device is operational immediately after fabrication. In the Ge2Se3 active layer, Ge-Ge homopolar bonds are found and switching occurs. The three layers consisting of Ge2Se3/Ag/Ge2Se3, directly below the top tungsten electrode, mix together during deposition and jointly form the silver-source layer. A layer of SnSe is between these two layers ensuring that the silver-source layer is not in direct contact with the active layer. Since silver does not migrate into the active layer at high temperatures, and the active layer maintains a high glass transition temperature of about 350 °C (662 °F), the device has significantly higher processing and operating temperatures at 250 °C (482 °F) and at least 150 °C (302 °F), respectively. These processing and operating temperatures are higher than most ion-conducting chalcogenide device types, including the S-based glasses (e.g. GeS) that need to be photodoped or thermally annealed. These factors allow the SDC device to operate over a wide range of temperatures, including long-term continuous operation at 150 °C (302 °F).

Potential applications

Memristors remain a laboratory curiosity, as yet made in insufficient numbers to gain any commercial applications. Despite this lack of mass availability, according to Allied Market Research the memristor market was worth $3.2 million in 2015 and will be worth $79.0 million by 2022.[93]

A potential application of memristors is in analog memories for superconducting quantum computers.[11]

Memristors can potentially be fashioned into non-volatile solid-state memory, which could allow greater data density than hard drives with access times similar to DRAM, replacing both components.[33] HP prototyped a crossbar latch memory that can fit 100 gigabits in a square centimeter,[94] and proposed a scalable 3D design (consisting of up to 1000 layers or 1 petabit per cm3).[95] In May 2008 HP reported that its device reaches currently about one-tenth the speed of DRAM.[96] The devices' resistance would be read with alternating current so that the stored value would not be affected.[97] In May 2012, it was reported that the access time had been improved to 90 nanoseconds, which is nearly one hundred times faster than the contemporaneous Flash memory. At the same time, the energy consumption was just one percent of that consumed by Flash memory.[98]

Memristor have applications in programmable logic,[99] signal processing,[100] Super-resolution imaging,[101] physical neural networks,[102] control systems,[103] reconfigurable computing,[104] brain-computer interfaces,[105] and RFID.[106] Memristive devices are potentially used for stateful logic implication, allowing a replacement for CMOS-based logic computation[107]. Several early works have been reported in this direction.[108][109]

In 2009, a simple electronic circuit[110] consisting of an LC network and a memristor was used to model experiments on adaptive behavior of unicellular organisms.[111] It was shown that subjected to a train of periodic pulses, the circuit learns and anticipates the next pulse similar to the behavior of slime molds Physarum polycephalum where the viscosity of channels in the cytoplasm responds to periodic environment changes.[111] Applications of such circuits may include, e.g., pattern recognition. The DARPA SyNAPSE project funded HP Labs, in collaboration with the Boston University Neuromorphics Lab, has been developing neuromorphic architectures which may be based on memristive systems. In 2010, Versace and Chandler described the MoNETA (Modular Neural Exploring Traveling Agent) model.[112] MoNETA is the first large-scale neural network model to implement whole-brain circuits to power a virtual and robotic agent using memristive hardware.[113] Application of the memristor crossbar structure in the construction of an analog soft computing system was demonstrated by Merrikh-Bayat and Shouraki.[114] In 2011, they showed[115] how memristor crossbars can be combined with fuzzy logic to create an analog memristive neuro-fuzzy computing system with fuzzy input and output terminals. Learning is based on the creation of fuzzy relations inspired from Hebbian learning rule.

In 2013 Leon Chua published a tutorial underlining the broad span of complex phenomena and applications that memristors span and how they can be used as non-volatile analog memories and can mimic classic habituation and learning phenomena.[116]

Derivative devices

Memistor and memtransistor

The memistor and memtransistor are transistor based devices which include memristor function.

Memcapacitors and meminductors

In 2009, Di Ventra, Pershin, and Chua extended[117] the notion of memristive systems to capacitive and inductive elements in the form of memcapacitors and meminductors, whose properties depend on the state and history of the system, further extended in 2013 by Di Ventra and Pershin.[21]

Memfractance and memfractor, 2nd- and 3rd-order memristor, memcapacitor and meminductor

In September 2014, Mohamed-Salah Abdelouahab, Rene Lozi, and Leon Chua published a general theory of 1st-, 2nd-, 3rd-, and nth-order memristive elements using fractional derivatives.[118]

History

Precursors

Sir Humphry Davy is said by some to have performed the first experiments which can be explained by memristor effects as long ago as 1808.[19][119] However the first device of a related nature to be constructed was the memistor (i.e. memory resistor), a term coined in 1960 by Bernard Widrow to describe a circuit element of an early artificial neural network called ADALINE. A few years later, in 1968, Argall published an article showing the resistance switching effects of TiO2 which was later claimed by researchers from Hewlett Packard to be evidence of a memristor.[56]

Theoretical description

Leon Chua postulated his new two-terminal circuit element in 1971. It was characterized by a relationship between charge and flux linkage as a fourth fundamental circuit element.[1] Five years later he and his student Sung Mo Kang generalized the theory of memristors and memristive systems including a property of zero crossing in the Lissajous curve characterizing current vs. voltage behavior.[2]

Twenty-first century

On May 1, 2008, Strukov, Snider, Stewart, and Williams published an article in Nature identifying a link between the 2-terminal resistance switching behavior found in nanoscale systems and memristors.[16]

On January 23, 2009, Di Ventra, Pershin, and Chua extended the notion of memristive systems to capacitive and inductive elements, namely capacitors and inductors, whose properties depend on the state and history of the system.[117]

In July 2014, the MeMOSat/LabOSat group[120] (composed of researchers from Universidad Nacional de General San Martín (Argentina), INTI, CNEA, and CONICET) put memory devices into orbit for their study at LEO.[121] Since then, seven missions with different devices[122] are performing experiments in low orbit, onboard Satellogic's Ñu-Sat satellites.[123][124]

On July 7, 2015 Knowm Inc announced Self Directed Channel (SDC) memristors commercially.[125] These devices remain available in small numbers.

On July 13, 2018 MemSat (Memristor Satellite) was launched to fly a memristor evaluation payload.[126]

See also

References

  1. Chua, L. (1971). "Memristor-The missing circuit element". IEEE Transactions on Circuit Theory. 18 (5): 507–519. CiteSeerX 10.1.1.189.3614. doi:10.1109/TCT.1971.1083337.
  2. Chua, L. O.; Kang, S. M. (1 January 1976), "Memristive devices and systems", Proceedings of the IEEE, 64 (2): 209–223, doi:10.1109/PROC.1976.10092
  3. Knoepfel, H. (1970), Pulsed high magnetic fields, New York: North-Holland, p. 37, Eq. (2.80)
  4. Muthuswamy, Bharathwaj; Banerjee, Santo (2019). Introduction to Nonlinear Circuits and Networks. Springer International. ISBN 978-3-319-67325-7.
  5. Pershin, Y. V.; Di Ventra, M. (2019). "A simple test for ideal memristors". Journal of Physics D: Applied Physics. 52 (1): 01LT01. arXiv:1806.07360. Bibcode:2019JPhD...52aLT01P. doi:10.1088/1361-6463/aae680.
  6. Paul L. Penfield, Jr. (1974). "1. Frequency-Power Formulas for Josephson Junctions". V. Microwave and Millimeter Wave Techniques (PDF) (Report). pp. 31–32. QPR No. 113.
  7. Langenberg, D. N. (1974), "Physical Interpretation of the term and implications for detectors", Rev. Phys. Appl. (Paris), 9: 35–40, doi:10.1051/rphysap:019740090103500
  8. Pedersen, N.F.; et al. (1972), "Magnetic field dependence and Q of the Josephson plasma resonance" (PDF), Phys. Rev. B, 11 (6): 4151–4159, Bibcode:1972PhRvB...6.4151P, doi:10.1103/PhysRevB.6.4151
  9. Pedersen, N.F.; et al. (1974), "Evidence for the existence of the Josephson quasiparticle-pair interference current", Low Temp. Phys., 13: 268–271
  10. Thompson, E.D. (1973), "Power flow for Josephson Elements", IEEE Trans. Electron Devices, 20 (8): 680–683, Bibcode:1973ITED...20..680T, doi:10.1109/T-ED.1973.17728
  11. Peotta, A.; Di Ventra, M. (2014), "Superconducting Memristors", Phys. Rev. Appl., 2 (3): 034011-1–034011-10, arXiv:1311.2975, Bibcode:2014PhRvP...2c4011P, doi:10.1103/PhysRevApplied.2.034011
  12. Muthuswamy, B.; Jevtic, J.; Iu, H. H. C.; Subramaniam, C. K.; Ganesan, K.; Sankaranarayanan, V.; Sethupathi, K.; Kim, H.; Shah, M. Pd.; Chua, L. O. (2014). "Memristor modelling". 2014 IEEE International Symposium on Circuits and Systems (ISCAS). pp. 490–493. doi:10.1109/ISCAS.2014.6865179. ISBN 978-1-4799-3432-4.
  13. Sah, M.; et al. (2015), "A Generic Model of Memristors with Parasitic Components", IEEE TCAS I: Regular Papers, 62 (3): 891–898
  14. Chua, L. O.; Tseng, C. (1974), "A Memristive Circuit Model for P-N Junction Diodes", International Journal of Circuit Theory and Applications, 2 (4): 367–389, doi:10.1002/cta.4490020406
  15. Chua, Leon (28 January 2011). "Resistance switching memories are memristors". Applied Physics A. 102 (4): 765–783. Bibcode:2011ApPhA.102..765C. doi:10.1007/s00339-011-6264-9.
  16. Strukov, Dmitri B.; Snider, Gregory S.; Stewart, Duncan R.; Williams, R. Stanley (2008). "The missing memristor found" (PDF). Nature. 453 (7191): 80–83. Bibcode:2008Natur.453...80S. doi:10.1038/nature06932. ISSN 1476-4687. PMID 18451858.
  17. Memristor FAQ, Hewlett-Packard, retrieved 2010-09-03
  18. Williams, R. S. (2008). "How We Found The Missing Memristor" (PDF). IEEE Spectrum. 45 (12): 28–35. doi:10.1109/MSPEC.2008.4687366. ISSN 0018-9235.
  19. Clarke, P. (23 May 2012), "Memristor is 200 years old, say academics", EE Times, retrieved 2012-05-25
  20. Meuffels, P.; Soni, R. (2012). "Fundamental Issues and Problems in the Realization of Memristors". arXiv:1207.7319 [cond-mat.mes-hall].
  21. Di Ventra, M.; Pershin, Y. V. (2013), "On the physical properties of memristive, memcapacitive and meminductive systems", Nanotechnology, 24 (25): 255201, arXiv:1302.7063, Bibcode:2013Nanot..24y5201D, CiteSeerX 10.1.1.745.8657, doi:10.1088/0957-4484/24/25/255201, PMID 23708238
  22. Sundqvist, Kyle M.; Ferry, David K.; Kish, Laszlo B. (21 November 2017). "Memristor Equations: Incomplete Physics and Undefined Passivity/Activity". Fluctuation and Noise Letters. 16 (4): 1771001–519. arXiv:1703.09064. Bibcode:2017FNL....1671001S. doi:10.1142/S0219477517710018.
  23. Vongehr, Sascha; Meng, Xiangkang (25 June 2015). "The Missing Memristor has Not been Found" (PDF). Scientific Reports. 5 (1): 11657. Bibcode:2015NatSR...511657V. doi:10.1038/srep11657. ISSN 2045-2322. PMC 4479989. PMID 26108376.
  24. Abraham, Isaac (2018-07-20). "The case for rejecting the memristor as a fundamental circuit element". Scientific Reports. 8 (1): 10972. Bibcode:2018NatSR...810972A. doi:10.1038/s41598-018-29394-7. ISSN 2045-2322. PMC 6054652. PMID 30030498.
  25. Valov, I.; et al. (2013), "Nanobatteries in redox-based resistive switches require extension of memristor theory", Nature Communications, 4 (4): 1771, arXiv:1303.2589, Bibcode:2013NatCo...4.1771V, doi:10.1038/ncomms2784, PMC 3644102, PMID 23612312
  26. Kim, J.; Pershin, Y. V.; Yin, M.; Datta, T.; Di Ventra, M. (2019). "An experimental proof that resistance-switching memories are not memristors". arXiv:1909.07238 [cond-mat.mes-hall].
  27. "A Survey of ReRAM-based Architectures for Processing-in-memory and Neural Networks", S. Mittal, Machine Learning and Knowledge Extraction, 2018
  28. Marks, P. (30 April 2008), "Engineers find 'missing link' of electronics", New Scientist, retrieved 2008-04-30
  29. Zidan, Mohammed A.; Strachan, John Paul; Lu, Wei D. (2018-01-08). "The future of electronics based on memristive systems". Nature Electronics. 1 (1): 22–29. doi:10.1038/s41928-017-0006-8. ISSN 2520-1131.
  30. HP 100TB Memristor drives by 2018 – if you're lucky, admits tech titan, 1 November 2013
  31. Artificial synapses could lead to advanced computer memory and machines that mimic biological brains, HRL Laboratories, March 23, 2012, retrieved March 30, 2012
  32. Bush, S. (2 May 2008), "HP nano device implements memristor", Electronics Weekly
  33. Kanellos, M. (30 April 2008), "HP makes memory from a once theoretical circuit", CNET News, retrieved 2008-04-30
  34. Mellor, C. (10 October 2011), "HP and Hynix to produce the memristor goods by 2013", The Register, retrieved 2012-03-07
  35. Courtland, R. (1 April 2011). "Memristors...Made of Blood?". IEEE Spectrum. Retrieved 2012-03-07.
  36. Johnsen, g k (24 March 2011). "Memristive model of electro-osmosis in skin". Phys. Rev. E. 83 (3): 031916. Bibcode:2011PhRvE..83c1916J. doi:10.1103/PhysRevE.83.031916. PMID 21517534.
  37. McAlpine, K. (2 March 2011), "Sweat ducts make skin a memristor", New Scientist, 209 (2802): 16, Bibcode:2011NewSc.209...16M, doi:10.1016/S0262-4079(11)60481-8, retrieved 2012-03-07
  38. Clarke, P. (16 January 2012), "Memristor brouhaha bubbles under", EETimes, retrieved 2012-03-02
  39. Marks, P. (23 February 2012), "Online spat over who joins memristor club", New Scientist, retrieved 2012-03-19
  40. Meuffels, P.; Schroeder, H. (2011), "Comment on "Exponential ionic drift: fast switching and low volatility of thin-film memristors" by D. B. Strukov and R. S. Williams in Appl. Phys. A (2009) 94: 515–519", Applied Physics A, 105 (1): 65–67, Bibcode:2011ApPhA.105...65M, doi:10.1007/s00339-011-6578-7
  41. Kish, Laszlo B.; Granqvist, Claes G.; Khatri, Sunil P.; Wen, He (16 September 2014). "Demons: Maxwell's demon, Szilard's engine and Landauer's erasure–dissipation". International Journal of Modern Physics: Conference Series. 33: 1460364. arXiv:1412.2166. Bibcode:2014IJMPS..3360364K. doi:10.1142/s2010194514603640.
  42. Kish, L. B.; Khatri, S. P.; Granqvist, C. G.; Smulko, J. M. (2015). "Critical remarks on Landauer's principle of erasure-dissipation: Including notes on Maxwell demons and Szilard engines". 2015 International Conference on Noise and Fluctuations (ICNF). pp. 1–4. doi:10.1109/ICNF.2015.7288632. ISBN 978-1-4673-8335-6.
  43. Slipko, V. A.; Pershin, Y. V.; Di Ventra, M. (2013), "Changing the state of a memristive system with white noise", Physical Review E, 87 (1): 042103, arXiv:1209.4103, Bibcode:2013PhRvE..87a2103L, doi:10.1103/PhysRevE.87.012103, PMID 23410279
  44. Hashem, N.; Das, S. (2012), "Switching-time analysis of binary-oxide memristors via a non-linear model" (PDF), Applied Physics Letters, 100 (26): 262106, Bibcode:2012ApPhL.100z2106H, doi:10.1063/1.4726421, retrieved 2012-08-09
  45. Linn, E.; Siemon, A.; Waser, R.; Menzel, S. (23 March 2014). "Applicability of Well-Established Memristive Models for Simulations of Resistive Switching Devices". IEEE Transactions on Circuits and Systems I: Regular Papers. 61 (8): 2402–2410. arXiv:1403.5801. Bibcode:2014arXiv1403.5801L. doi:10.1109/TCSI.2014.2332261.
  46. Garling, C. (25 July 2012), "Wonks question HP's claim to computer-memory missing link", Wired.com, retrieved 2012-09-23
  47. Vongehr, S. (21 Mar 2015). "Purely Mechanical Memristors and the Missing Memristor". arXiv:1504.00300 [physics.gen-ph].
  48. Chua, L. (13 June 2012), Memristors: Past, Present and future (PDF), archived from the original (PDF) on 8 March 2014, retrieved 2013-01-12
  49. Adhikari, S. P.; Sah, M. P.; Hyongsuk, K.; Chua, L. O. (2013), "Three Fingerprints of Memristor", IEEE Transactions on Circuits and Systems I, 60 (11): 3008–3021, doi:10.1109/TCSI.2013.2256171
  50. Pershin, Y. V.; Di Ventra, M. (2011), "Memory effects in complex materials and nanoscale systems", Advances in Physics, 60 (2): 145–227, arXiv:1011.3053, Bibcode:2011AdPhy..60..145P, doi:10.1080/00018732.2010.544961
  51. Biolek, D.; Biolek, Z.; Biolkova, V. (2011), "Pinched hysteresis loops of ideal memristors, memcapacitors and meminductors must be 'self-crossing'", Electronics Letters, 47 (25): 1385–1387, doi:10.1049/el.2011.2913
  52. Mouttet, B. (2012). "Memresistors and non-memristive zero-crossing hysteresis curves". arXiv:1201.2626 [cond-mat.mes-hall].
  53. Fildes, J. (13 November 2007), Getting More from Moore's Law, BBC News, retrieved 2008-04-30
  54. Taylor, A. G. (2007), "Nanotechnology in the Northwest" (PDF), Bulletin for Electrical and Electronic Engineers of Oregon, 51 (1): 1
  55. Stanley Williams, HP Labs, archived from the original on 2011-07-19, retrieved 2011-03-20
  56. Argall, F. (1968), "Switching Phenomena in Titanium Oxide Thin Films", Solid-State Electronics, 11 (5): 535–541, Bibcode:1968SSEle..11..535A, doi:10.1016/0038-1101(68)90092-0
  57. Terabe, K.; Hasegawa, T.; Liang, C.; Aono, M. (2007), "Control of local ion transport to create unique functional nanodevices based on ionic conductors", Science and Technology of Advanced Materials, 8 (6): 536–542, Bibcode:2007STAdM...8..536T, doi:10.1016/j.stam.2007.08.002
  58. Beck, A.; et al. (2000), "Reproducible switching effect in thin oxide films for memory applications", Applied Physics Letters, 77 (1): 139, Bibcode:2000ApPhL..77..139B, doi:10.1063/1.126902
  59. US Patent 7,417,271
  60. US Patent Application 11/542,986
  61. Finding the Missing Memristor - R. Stanley Williams
  62. Markoff, J. (7 April 2010), "H.P. Sees a Revolution in Memory Chip", New York Times
  63. Krieger, J. H.; Spitzer, S. M. (2004), "Non-traditional, Non-volatile Memory Based on Switching and Retention Phenomena in Polymeric Thin Films", Proceedings of the 2004 Non-Volatile Memory Technology Symposium, IEEE, p. 121, doi:10.1109/NVMT.2004.1380823, ISBN 978-0-7803-8726-3
  64. Erokhin, V.; Fontana, M. P. (2008). "Electrochemically controlled polymeric device: A memristor (and more) found two years ago". arXiv:0807.0333 [cond-mat.soft].
  65. An; Alibart, F.; Pleutin, S.; Guerin, D.; Novembre, C.; Lenfant, S.; Lmimouni, K.; Gamrat, C.; Vuillaume, D. (2010). "An Organic Nanoparticle Transistor Behaving as a Biological Spiking Synapse". Adv. Funct. Mater. 20 (2): 330–337. arXiv:0907.2540. doi:10.1002/adfm.200901335.
  66. Alibart, F.; Pleutin, S.; Bichler, O.; Gamrat, C.; Serrano-Gotarredona, T.; Linares-Barranco, B.; Vuillaume, D. (2012). "A Memristive Nanoparticle/Organic Hybrid Synapstor for Neuroinspired Computing". Adv. Funct. Mater. 22 (3): 609–616. arXiv:1112.3138. doi:10.1002/adfm.201101935. hdl:10261/83537.
  67. Pavlov's; Transistors, Organic; Bichler, O.; Zhao, W.; Alibart, F.; Pleutin, S.; Lenfant, S.; Vuillaume, D.; Gamrat, C. (2013). "Pavlov's Dog Associative Learning Demonstrated on Synaptic-Like Organic Transistors". Neural Computation. 25 (2): 549–566. arXiv:1302.3261. Bibcode:2013arXiv1302.3261B. doi:10.1162/NECO_a_00377. PMID 22970878.
  68. Crupi, M.; Pradhan, L.; Tozer, S. (2012), "Modelling Neural Plasticity with Memristors" (PDF), IEEE Canadian Review, 68: 10–14
  69. Bessonov, A. A.; et al. (2014), "Layered memristive and memcapacitive switches for printable electronics", Nature Materials, 14 (2): 199–204, Bibcode:2015NatMa..14..199B, doi:10.1038/nmat4135, PMID 25384168
  70. Ge, Ruijing; Wu, Xiaohan; Kim, Myungsoo; Shi, Jianping; Sonde, Sushant; Tao, Li; Zhang, Yanfeng; Lee, Jack C.; Akinwande, Deji (2017-12-19). "Atomristor: Nonvolatile Resistance Switching in Atomic Sheets of Transition Metal Dichalcogenides". Nano Letters. 18 (1): 434–441. Bibcode:2018NanoL..18..434G. doi:10.1021/acs.nanolett.7b04342. ISSN 1530-6984. PMID 29236504.
  71. Kim, Myungsoo; Ge, Ruijing; Wu, Xiaohan; Lan, Xing; Tice, Jesse; Lee, Jack C.; Akinwande, Deji (2018-06-28). "Zero-static power radio-frequency switches based on MoS2 atomristors". Nature Communications. 9 (1): 2524. Bibcode:2018NatCo...9.2524K. doi:10.1038/s41467-018-04934-x. ISSN 2041-1723. PMC 6023925. PMID 29955064.
  72. Chanthbouala, A.; et al. (2012), "A ferroelectric memristor", Nature Materials, 11 (10): 860–864, arXiv:1206.3397, Bibcode:2012NatMa..11..860C, doi:10.1038/nmat3415, PMID 22983431
  73. Ageev, O. A.; Blinov, Yu F.; Il’in, O. I.; Kolomiitsev, A. S.; Konoplev, B. G.; Rubashkina, M. V.; Smirnov, V. A.; Fedotov, A. A. (11 December 2013). "Memristor effect on bundles of vertically aligned carbon nanotubes tested by scanning tunnel microscopy". Technical Physics. 58 (12): 1831–1836. Bibcode:2013JTePh..58.1831A. doi:10.1134/S1063784213120025. ISSN 1063-7842.
  74. Il'ina, Marina V.; Il'in, Oleg I.; Blinov, Yuriy F.; Smirnov, Vladimir A.; Kolomiytsev, Alexey S.; Fedotov, Alexander A.; Konoplev, Boris G.; Ageev, Oleg A. (October 2017). "Memristive switching mechanism of vertically aligned carbon nanotubes". Carbon. 123: 514–524. doi:10.1016/j.carbon.2017.07.090. ISSN 0008-6223.
  75. Wang, X.; Chen, Y.; Xi, H.; Dimitrov, D. (2009), "Spintronic Memristor through Spin Torque Induced Magnetization Motion", IEEE Electron Device Letters, 30 (3): 294–297, Bibcode:2009IEDL...30..294W, doi:10.1109/LED.2008.2012270
  76. Savage, N. (16 March 2009). "Spintronic Memristor". IEEE Spectrum. Archived from the original on 2010-12-24. Retrieved 2011-03-20.
  77. Chanthbouala, A.; Matsumoto, R.; Grollier, J.; Cros, V.; Anane, A.; Fert, A.; Khvalkovskiy, A. V.; Zvezdin, K. A.; Nishimura, K.; Nagamine, Y.; Maehara, H.; Tsunekawa, K.; Fukushima, A.; Yuasa, S. (10 April 2011). "Vertical-current-induced domain-wall motion in MgO-based magnetic tunnel junctions with low current densities". Nature Physics. 7 (8): 626–630. arXiv:1102.2106. Bibcode:2011NatPh...7..626C. doi:10.1038/nphys1968.
  78. Bowen, M.; Maurice, J.-L.; Barthe´le´my, A.; Prod’homme, P.; Jacquet, E.; Contour, J.-P.; Imhoff, D.; Colliex, C. (2006). "Bias-crafted magnetic tunnel junctions with bistable spin-dependent states". Applied Physics Letters. 89 (10): 103517. Bibcode:2006ApPhL..89j3517B. doi:10.1063/1.2345592. ISSN 0003-6951.
  79. Halley, D.; Majjad, H.; Bowen, M.; Najjari, N.; Henry, Y.; Ulhaq-Bouillet, C.; Weber, W.; Bertoni, G.; Verbeeck, J.; Van Tendeloo, G. (2008). "Electrical switching in Fe/Cr/MgO/Fe magnetic tunnel junctions". Applied Physics Letters. 92 (21): 212115. Bibcode:2008ApPhL..92u2115H. doi:10.1063/1.2938696. ISSN 0003-6951.
  80. Krzysteczko, P.; Günter, R.; Thomas, A. (2009), "Memristive switching of MgO based magnetic tunnel junctions", Applied Physics Letters, 95 (11): 112508, arXiv:0907.3684, Bibcode:2009ApPhL..95k2508K, CiteSeerX 10.1.1.313.2571, doi:10.1063/1.3224193
  81. Bertin, Eric; Halley, David; Henry, Yves; Najjari, Nabil; Majjad, Hicham; Bowen, Martin; DaCosta, Victor; Arabski, Jacek; Doudin, Bernard (2011), "Random barrier double-well model for resistive switching in tunnel barriers", Journal of Applied Physics, 109 (8): 013712–013712–5, Bibcode:2011JAP...109a3712D, doi:10.1063/1.3530610, retrieved 2014-12-15
  82. Schleicher, F.; Halisdemir, U.; Lacour, D.; Gallart, M.; Boukari, S.; Schmerber, G.; Davesne, V.; Panissod, P.; Halley, D.; Majjad, H.; Henry, Y.; Leconte, B.; Boulard, A.; Spor, D.; Beyer, N.; Kieber, C.; Sternitzky, E.; Cregut, O.; Ziegler, M.; Montaigne, F.; Beaurepaire, E.; Gilliot, P.; Hehn, M.; Bowen, M. (2014-08-04), "Localized states in advanced dielectrics from the vantage of spin- and symmetry-polarized tunnelling across MgO", Nature Communications, 5: 4547, Bibcode:2014NatCo...5.4547S, doi:10.1038/ncomms5547, ISSN 2041-1723, PMID 25088937
  83. Garcia, V.; Bibes, M.; Bocher, L.; Valencia, S.; Kronast, F.; Crassous, A.; Moya, X.; Enouz-Vedrenne, S.; Gloter, A.; Imhoff, D.; Deranlot, C.; Mathur, N. D.; Fusil, S.; Bouzehouane, K.; Barthelemy, A. (2010-02-26), "Ferroelectric Control of Spin Polarization", Science, 327 (5969): 1106–1110, Bibcode:2010Sci...327.1106G, doi:10.1126/science.1184028, ISSN 0036-8075, PMID 20075211
  84. Pantel, D.; Goetze, S.; Hesse, D.; Alexe, M. (2012-02-26), "Reversible electrical switching of spin polarization in multiferroic tunnel junctions", Nature Materials, 11 (4): 289–293, Bibcode:2012NatMa..11..289P, doi:10.1038/nmat3254, ISSN 1476-1122, PMID 22367005
  85. Huai, Y. (2008), "Spin-Transfer Torque MRAM (STT-MRAM): Challenges and Prospects" (PDF), AAPPS Bulletin, 18 (6): 33, archived from the original (PDF) on 2012-03-23
  86. Krzysteczko, P.; Münchenberger, J.; Schäfers, M.; Reiss, G.; Thomas, A. (2012), "The Memristive Magnetic Tunnel Junction as a Nanoscopic Synapse-Neuron System", Advanced Materials, 24 (6): 762–766, Bibcode:2012APS..MAR.H5013T, doi:10.1002/adma.201103723, PMID 22223304
  87. "Yuriy V. Pershin".
  88. "Massimiliano Di Ventra's Homepage". physics.ucsd.edu.
  89. Pershin, Y. V.; Di Ventra, M. (2008), "Spin memristive systems: Spin memory effects in semiconductor spintronics", Physical Review B, 78 (11): 113309, arXiv:0806.2151, Bibcode:2008PhRvB..78k3309P, doi:10.1103/PhysRevB.78.113309
  90. Pershin, Y. V.; Di Ventra, M. (2008), "Current-voltage characteristics of semiconductor/ferromagnet junctions in the spin-blockade regime", Physical Review B, 77 (7): 073301, arXiv:0707.4475, Bibcode:2008PhRvB..77g3301P, doi:10.1103/PhysRevB.77.073301
  91. Campbell, K. (January 2017), "Self-directed channel memristor for high temperature operation", Microelectronics Journal, 59: 10–14, arXiv:1608.05357, doi:10.1016/j.mejo.2016.11.006
  92. Knowm Memristors, Knowm Inc
  93. "Memristor Market Expected to Reach $79.0 Million by 2020, Globally - Allied Market Research". Archived from the original on 2017-02-26. Retrieved 2017-02-25.
  94. Johnson, R. C. (30 April 2008), "'Missing link' memristor created", EE Times, retrieved 2008-04-30
  95. Finding the Missing Memristor - R. Stanley Williams
  96. Markoff, J. (1 May 2008), "H.P. Reports Big Advance in Memory Chip Design", New York Times, retrieved 2008-05-01
  97. Gutmann, E. (1 May 2008), "Maintaining Moore's law with new memristor circuits", Ars Technica, retrieved 2008-05-01
  98. Palmer, J. (18 May 2012), "Memristors in silicon promising for dense, fast memory", BBC News, retrieved 2012-05-18
  99. U.S. Patent 7,203,789
  100. U.S. Patent 7,302,513
  101. Dong, Zhekang; Sing Lai, Chun; He, Yufei; Qi, Donglian; Duan, Shukai (1 November 2019). "Hybrid dual-complementary metal–oxide–semiconductor/memristor synapse-based neural network with its applications in image super-resolution". IET Circuits, Devices & Systems. 13 (8): 1241–1248. doi:10.1049/iet-cds.2018.5062.
  102. U.S. Patent 7,359,888
  103. U.S. Patent 7,609,086
  104. U.S. Patent 7,902,857
  105. U.S. Patent 7,902,867
  106. U.S. Patent 8,113,437
  107. Luo, Li; Dong, Zhekang; Duan, Shukai; Lai, Chun Sing (20 April 2020). "Memristor-based stateful logic gates for multi-functional logic circuit". IET Circuits, Devices & Systems. doi:10.1049/iet-cds.2019.0422.
  108. Lehtonen, E.; Poikonen, J.H.; Laiho, M. (2010). "Two memristors suffice to compute all Boolean functions". Electronics Letters. 46 (3): 230. doi:10.1049/el.2010.3407.
  109. Chattopadhyay, A.; Rakosi, Z. (2011). "Combinational logic synthesis for material implication". 2011 IEEE/IFIP 19th International Conference on VLSI and System-on-Chip. p. 200. doi:10.1109/VLSISoC.2011.6081665. ISBN 978-1-4577-0170-2.
  110. Pershin, Y. V.; La Fontaine, S.; Di Ventra, M. (2009), "Memristive model of amoeba learning", Physical Review E, 80 (2): 021926, arXiv:0810.4179, Bibcode:2009PhRvE..80b1926P, doi:10.1103/PhysRevE.80.021926, PMID 19792170
  111. Saigusa, T.; Tero, A.; Nakagaki, T.; Kuramoto, Y. (2008), "Amoebae Anticipate Periodic Events" (PDF), Physical Review Letters, 100 (1): 018101, Bibcode:2008PhRvL.100a8101S, doi:10.1103/PhysRevLett.100.018101, hdl:2115/33004, PMID 18232821
  112. Versace, M.; Chandler, B. (23 November 2010). "MoNETA: A Mind Made from Memristors". IEEE Spectrum.
    Versace, M.; Chandler, B. (2010). "The brain of a new machine". IEEE Spectrum. 47 (12): 30–37. doi:10.1109/MSPEC.2010.5644776.
  113. Snider, G.; et al. (2011), "From Synapses to Circuitry: Using Memristive Memory to Explore the Electronic Brain", IEEE Computer, 44 (2): 21–28, doi:10.1109/MC.2011.48
  114. Merrikh-Bayat, F.; Bagheri-Shouraki, S.; Rohani, A. (2011), "Memristor crossbar-based hardware implementation of IDS method", IEEE Transactions on Fuzzy Systems, 19 (6): 1083–1096, arXiv:1008.5133, doi:10.1109/TFUZZ.2011.2160024
  115. Merrikh-Bayat, F.; Bagheri-Shouraki, S. (2011). "Efficient neuro-fuzzy system and its Memristor Crossbar-based Hardware Implementation". arXiv:1103.1156 [cs.AI].
  116. Chua, L. (2013). "Memristor, Hodgkin-Huxley, and Edge of Chaos". Nanotechnology. 24 (38): 383001. Bibcode:2013Nanot..24L3001C. doi:10.1088/0957-4484/24/38/383001. PMID 23999613.
  117. Di Ventra, M.; Pershin, Y. V.; Chua, L. (2009), "Circuit elements with memory: memristors, memcapacitors and meminductors", Proceedings of the IEEE, 97 (10): 1717–1724, arXiv:0901.3682, Bibcode:2009arXiv0901.3682D, doi:10.1109/JPROC.2009.2021077
  118. Abdelhouahad, M.-S.; Lozi, R.; Chua, L. (September 2014), "Memfractance: A Mathematical Paradigm for Circuit Elements with Memory" (PDF), International Journal of Bifurcation and Chaos, 24 (9): 1430023 (29 pages), Bibcode:2014IJBC...2430023A, doi:10.1142/S0218127414300237
  119. Prodromakis, T.; Toumazou, C.; Chua, L. (June 2012), "Two centuries of memristors", Nature Materials, 11 (6): 478–481, Bibcode:2012NatMa..11..478P, doi:10.1038/nmat3338, PMID 22614504
  120. Barella, M. (2016), "LabOSat: Low cost measurement platform designed for hazardous environments", 2016 Seventh Argentine Conference on Embedded Systems (CASE), pp. 1–6, doi:10.1109/SASE-CASE.2016.7968107, ISBN 978-987-46297-0-8
  121. "Probaron con éxito las memorias instaladas en el satélite argentino "Tita"". Telam. July 21, 2014.
  122. Barella, M. (2019), "Studying ReRAM devices at Low Earth Orbits using the LabOSat platform", Radiation Physics and Chemistry, 154: 85–90, Bibcode:2019RaPC..154...85B, doi:10.1016/j.radphyschem.2018.07.005
  123. "LabOSat".
  124. "Qué hace LabOSat, el laboratorio electrónico dentro de los nanosatélites Fresco y Batata". Telam. June 22, 2016.
  125. "Startup Beats HP, Hynix to Memristor Learning". EE Times. July 5, 2015.
  126. "MemSat". Gunter Space Page. May 22, 2018.

Further reading

This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.