measurements of nonlinear polarization dynamics in the

13
PHYSICAL REVIEW APPLIED 13, 044026 (2020) Editors’ Suggestion Measurements of Nonlinear Polarization Dynamics in the Tens of Gigahertz Aaron M. Hagerstrom , 1,2, * Eric J. Marksz , 1,3 Xiaohang Zhang, 3 Xifeng Lu, 1,2 Christian J. Long, 1 James C. Booth , 1 Ichiro Takeuchi, 3 and Nathan D. Orloff 1 1 Communications Technology Laboratory (CTL), National Institute of Standards and Technology (NIST), 325 Broadway, Boulder, Colorado 80305, USA 2 Department of Physics, University of Colorado, Boulder, Colorado 80309, USA 3 Department of Materials Science and Engineering, University of Maryland, College Park, Maryland 20742, USA (Received 12 August 2019; revised manuscript received 21 February 2020; accepted 4 March 2020; published 9 April 2020) Frequency-dependent linear-permittivity measurements are commonplace in the literature, providing key insights into the structure of dielectric materials. These measurements describe a material’s dynamic response to a small applied electric field. However, nonlinear dielectric materials are widely used for their responses to large applied fields, including switching in ferroelectric materials, and field tuning of the permittivity in paraelectric materials. These behaviors are described by nonlinear permittivity. Nonlinear-permittivity measurements are fraught with technical challenges because of the complex elec- trical coupling between a sample and its environment. Here, we describe a technique for measuring the complex nonlinear permittivity that circumvents many of the difficulties associated with other approaches. We validate this technique by measuring the nonlinear permittivity of a tunable Ba 0.5 Sr 0.5 TiO 3 thin film up to 40 GHz and comparing our results with a phenomenological model. These measurements provide insight into the dynamics of nonlinear dielectric materials down to picosecond timescales. DOI: 10.1103/PhysRevApplied.13.044026 I. OVERVIEW Ferroelectric and related materials have many applica- tions that rely on a nonlinear relationship between the electric field and the polarization [1]. For example, the switching of the built-in polarization leads to both non- volatile memory and alternative transistor designs that utilize switching dynamics to increase power efficiency [28]. In microwave electronics, tunability of the permit- tivity with applied electric field enables components such as filters and phase shifters to be reconfigured by applying a dc voltage [912]. These switching and tuning processes can have complex time dependence that is difficult to measure at microwave frequencies. In nonlinear dielectric materials, mesoscale structure, including domain walls and polar nanoregions (PNRs), [1320] strongly influences the dynamics of nonlin- ear dielectric materials under an applied electric field. For example, PNRs in Ba x Sr 1x TiO 3 thin films affect microwave-frequency dielectric loss and dispersion. Sev- eral reports also attribute the broad temperature depen- dence of the ferroelectric-paraelectric phase transition to PNRs [1719,19,2126]. While a great deal of previous research focused on understanding the effects of nanos- tructure on dielectric properties, in recent years researchers * [email protected] are making rapid progress in designing specific nanostruc- tures in dielectrics to induce specific properties. Examples include low-loss tunable Ruddlesden-Popper (RP) super- lattices [27], superlattices with polar vortices [28,29], and nanoparticles with vortex domains [30,31]. In all of these cases, theoretical tools (mesoscale theory and density- functional theory) could predict the static dielectric prop- erties [2730], but the dynamic properties remain more difficult to describe. Notably, even though reducing dielec- tric loss is a main motivation for the development of RP dielectrics, the current physical picture of loss mechanisms in those materials is qualitative, rather than predictive [20, 27]. In this case, the experimental characterization is lim- ited to linear permittivity and static tuning. The addition of nonlinear-permittivity measurements would provide addi- tional information that could aid in the development and validation of dynamic models. Figure 1 schematically shows two approaches to non- linear measurements. At low frequencies, the interpre- tation of time-domain measurements is straightforward [Fig. 1(a)]. One applies a steplike voltage pulse to a capacitor, and measures the current that flows as the capac- itor charges [3235]. The current is proportional to the time-derivative of the polarization. At short timescales (10 s of ps), this approach becomes more difficult because it becomes increasingly hard to account for all of the sources of signal distortion in the experiment [34,35]. 2331-7019/20/13(4)/044026(13) 044026-1 © 2020 American Physical Society

Upload: others

Post on 11-Dec-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Measurements of Nonlinear Polarization Dynamics in the

PHYSICAL REVIEW APPLIED 13, 044026 (2020)Editors’ Suggestion

Measurements of Nonlinear Polarization Dynamics in the Tens of Gigahertz

Aaron M. Hagerstrom ,1,2,* Eric J. Marksz ,1,3 Xiaohang Zhang,3 Xifeng Lu,1,2 Christian J. Long,1James C. Booth ,1 Ichiro Takeuchi,3 and Nathan D. Orloff 1

1Communications Technology Laboratory (CTL), National Institute of Standards and Technology (NIST),

325 Broadway, Boulder, Colorado 80305, USA2Department of Physics, University of Colorado, Boulder, Colorado 80309, USA

3Department of Materials Science and Engineering, University of Maryland, College Park, Maryland 20742, USA

(Received 12 August 2019; revised manuscript received 21 February 2020; accepted 4 March 2020; published 9 April 2020)

Frequency-dependent linear-permittivity measurements are commonplace in the literature, providingkey insights into the structure of dielectric materials. These measurements describe a material’s dynamicresponse to a small applied electric field. However, nonlinear dielectric materials are widely used fortheir responses to large applied fields, including switching in ferroelectric materials, and field tuningof the permittivity in paraelectric materials. These behaviors are described by nonlinear permittivity.Nonlinear-permittivity measurements are fraught with technical challenges because of the complex elec-trical coupling between a sample and its environment. Here, we describe a technique for measuring thecomplex nonlinear permittivity that circumvents many of the difficulties associated with other approaches.We validate this technique by measuring the nonlinear permittivity of a tunable Ba0.5Sr0.5TiO3 thin filmup to 40 GHz and comparing our results with a phenomenological model. These measurements provideinsight into the dynamics of nonlinear dielectric materials down to picosecond timescales.

DOI: 10.1103/PhysRevApplied.13.044026

I. OVERVIEW

Ferroelectric and related materials have many applica-tions that rely on a nonlinear relationship between theelectric field and the polarization [1]. For example, theswitching of the built-in polarization leads to both non-volatile memory and alternative transistor designs thatutilize switching dynamics to increase power efficiency[2–8]. In microwave electronics, tunability of the permit-tivity with applied electric field enables components suchas filters and phase shifters to be reconfigured by applyinga dc voltage [9–12]. These switching and tuning processescan have complex time dependence that is difficult tomeasure at microwave frequencies.

In nonlinear dielectric materials, mesoscale structure,including domain walls and polar nanoregions (PNRs),[13–20] strongly influences the dynamics of nonlin-ear dielectric materials under an applied electric field.For example, PNRs in BaxSr1−xTiO3 thin films affectmicrowave-frequency dielectric loss and dispersion. Sev-eral reports also attribute the broad temperature depen-dence of the ferroelectric-paraelectric phase transition toPNRs [17–19,19,21–26]. While a great deal of previousresearch focused on understanding the effects of nanos-tructure on dielectric properties, in recent years researchers

*[email protected]

are making rapid progress in designing specific nanostruc-tures in dielectrics to induce specific properties. Examplesinclude low-loss tunable Ruddlesden-Popper (RP) super-lattices [27], superlattices with polar vortices [28,29], andnanoparticles with vortex domains [30,31]. In all of thesecases, theoretical tools (mesoscale theory and density-functional theory) could predict the static dielectric prop-erties [27–30], but the dynamic properties remain moredifficult to describe. Notably, even though reducing dielec-tric loss is a main motivation for the development of RPdielectrics, the current physical picture of loss mechanismsin those materials is qualitative, rather than predictive [20,27]. In this case, the experimental characterization is lim-ited to linear permittivity and static tuning. The addition ofnonlinear-permittivity measurements would provide addi-tional information that could aid in the development andvalidation of dynamic models.

Figure 1 schematically shows two approaches to non-linear measurements. At low frequencies, the interpre-tation of time-domain measurements is straightforward[Fig. 1(a)]. One applies a steplike voltage pulse to acapacitor, and measures the current that flows as the capac-itor charges [32–35]. The current is proportional to thetime-derivative of the polarization. At short timescales(10 s of ps), this approach becomes more difficult becauseit becomes increasingly hard to account for all of thesources of signal distortion in the experiment [34,35].

2331-7019/20/13(4)/044026(13) 044026-1 © 2020 American Physical Society

Page 2: Measurements of Nonlinear Polarization Dynamics in the

AARON M. HAGERSTROM et al. PHYS. REV. APPLIED 13, 044026 (2020)

(a)

(b)

FIG. 1. Schematic picture of nonlinear measurements of thinfilms. (a) In time-domain measurements, tuning dynamics arecharacterized by applying a voltage pulse Vp(t) to a capacitor andmeasuring the current I(t) = V(t)/Zin that flows as the capacitorcharges. This current is related to the change of the polariza-tion with respect to time. (b) We characterize tuning dynamicsby exciting a waveguide patterned on a nonlinear material withtwo microwave frequencies ω1 and ω2, and measuring the wavesgenerated by nonlinear mixing. The complex amplitudes a and bcorrespond to the forward (towards the sample) and backward(from the sample) traveling waves, respectively. This methodcan be extended to much higher frequencies than time-domainmeasurements.

Time-domain calibrations are possible in principle, butrequire specialized instrumentation and would add a greatdeal of complexity to a measurement system [36]. We areunaware of any attempts to use time-domain calibrationsin this context. Instead, Refs. [34,35] relied on a circuitmodel that described the coupling between pulse width, thelinear capacitor charging time, and the switching dynam-ics of the material they measured. As a consequence, thephysical interpretation of their measurements depended ontheir model.

In principle, frequency-domain measurements [Fig. 1(b)]can circumvent these obstacles because one can correct forsignal distortion (Sec. III A). Frequency-domain measure-ments quantify dynamics in terms of harmonics and fre-quency mixing products generated by nonlinear processeswithin a material. Most frequency-mixing measurementsuse either low-frequency (<1 GHz) [25,26,37], or opti-cal [1,21,38] techniques. These frequency ranges allow foruseful approximations that simplify nonlinear-permittivitymeasurements. At low frequencies, one can use lumped-element circuit analysis; while at optical frequencies onecan use the slowly varying envelope approximation. In themicrowave regime, lumped-element models are increas-ingly difficult to apply at high frequencies. Larger testdevices can show distributed effects. These effects can bemitigated by using smaller devices, but then the calibrationis more sensitive to the difficult-to-model electromagneticfields near the probes. At the same time, the slowly varying

amplitude approximation is not viable due to sample sizeand dielectric loss.

Another challenge of microwave-frequency measure-ments is that the phase of the mixing products relative tothe incident signals, which carries vital information aboutthe dynamics, is difficult to determine. While the relativephase of two microwave signals at the same frequency canbe measured by mixing the two signals, cross-frequencyphase relationships are much harder to establish. Thesephase relationships are important because they are nec-essary to describe the shape of a time-domain waveforminfluenced by nonlinear processes. For example, the phaseof the mixing products distinguishes between nonlineargain and loss, and between increases and decreases inphase velocity.

Here, we describe a frequency-domain technique thatresolves the frequency-dependent complex nonlinear per-mittivity. Specifically, we use a large-signal networkanalyzer (LSNA) to demonstrate nonlinear-permittivitymeasurements from 0 to 40 GHz with cross-frequencyphase resolution and absolute power calibration. In ourexperiments, we fabricate waveguides on a nonlinearBa0.5Sr0.5TiO3 (BST) thin film, and measure the wavesgenerated by nonlinear mixing within the waveguide. Thekey contributions are cross-phase calibration, power cal-ibration, and our ability to interpret our results as funda-mental materials properties rather than as ad hoc circuitparameterizations [39–41].

This paper is organized as follows. Section II describeshow to relate nonlinear permittivity to experimentallyaccessible currents and voltages. Section III describes howwe perform calibrated nonlinear microwave measurementsof lithographically fabricated waveguide structures, andhow these measurements can determine linear and nonlin-ear permittivity. Section IV then describes how we expectthe measurements to look on the basis of phenomenolog-ical theory, and finally in Sec. V, we compare the resultsof a preliminary set of measurements to these theoreticalexpectations.

II. NONLINEAR CIRCUIT THEORY

Figure 2 illustrates how the nonlinear behavior of adielectric leads to small-signal frequency mixing and har-monic generation. The model we use here is given explic-itly in Sec. IV [42], but in this section we focus on itsqualitative behavior. To reflect the basic features of ourexperimental data (Sec. V), we chose a model that doesnot show any hysteresis but is nonlinear in the sense thatthe permittivity changes in response to a large appliedelectric field. Equivalently, this changing permittivity canalso be expressed as a nonlinear relationship between Dand E [Fig. 2(a)]. When a field is applied, the materialtakes some time to relax to a new configuration. The relax-ation dynamics lead to a difference between the response

044026-2

Page 3: Measurements of Nonlinear Polarization Dynamics in the

MEASUREMENTS OF NONLINEAR POLARIZATION... PHYS. REV. APPLIED 13, 044026 (2020)

E (arb. units)

E (

arb.

uni

ts)

D (

arb.

uni

ts)

d (a

rb. u

nits

)

Time (arb. units)

Time (arb. units) Time (arb. units)

D (

arb.

uni

ts)

0-8

00

-0.8

–2 –2

(a) (b)

(d)(c)

D( )

-4 0 4 8

–1

0

1

2

4 8 12

0

2

4

6

-0.4

0

0.4

0.8

4 8 12 0 4 8 12

0.5

1

1.5

2

D(2)D(1)

d(2)d(1)

FIG. 2. Illustration of the Volterra representation of dynamics.(a) Time-varying nonlinear relationship between the displace-ment field D(t) and the electric field E(t) superimposed with thedc D–E curve. (b) Two-tone input signal E(t). (c) First-order[d(1)(t), teal line] and second-order [d(2)(t), magenta line] con-tributions to the output signal. (d) Comparison of the first-orderapproximation D(1)(t) (teal line), second-order approximationD(2)(t) (purple line), and a solution obtained by direct numericalsolution of the equation of motion D(∞)(t) (cyan line).

to a time-varying field [the cyan line in Fig. 2(a)] and theresponse to a dc field (the black line). While the model doesnot include hysteresis, the time lag between the response ofthe material and the applied field leads to loops in the D–Eplot.

In our experiment, we characterize the dynamic non-linear behavior using small sinusoidal electric fields. Forreference, the amplitude of the sinusoidal signals is below0.03 V μm−1 root-mean-square, while the field necessaryto change the permittivity by about 33% is 8 V μm−1. Atlower frequencies it is possible to reconstruct a full D(E)

curve by applying a large, time-varying electric field andmeasuring harmonics of the input signal [43]. However, atmicrowave frequencies, small signals are convenient forseveral reasons. First, dielectric loss increases with fre-quency, and a microwave-frequency signal large enoughto significantly change the dielectric properties of a sam-ple will also change the sample through heating. Anotherdifficulty is that the nonlinear behavior in the measurementequipment can obscure the nonlinear response of the sam-ple. While the nonlinearity of the measurement equipmentcan be mitigated, this is typically only possible in a narrowfrequency range [44,45]. At the power levels we employ,

we only detect the leading-order nonlinear response, thesecond-order mixing products. We measure these mixingproducts in the presence of a large applied dc field, becausewithout this field, the second-order mixing products weseek to measure are suppressed by symmetry.

In Fig. 2, we show a simulation analogous to our exper-iment: a superposition of a dc field and two sinusoidalsignals at different frequencies ω1 and ω2 [Fig. 2(b)] isapplied to the material. Conceptually, one sinusoidal sig-nal alters the permittivity of the material, and the secondsignal probes this change in permittivity. The responseof the material includes a linear response, d(1), whichconsists of the Fourier components with frequencies ω1and ω2, as well as a second-order nonlinear response d(2)

with Fourier components ω1 + ω2, |ω1 − ω2|, 2ω1, 2ω2,and dc. In Fig. 2(c), we plot d(1) and d(2). Figure 2(d)compares three models of the response of the material: alinear approximation D(1) = Ddc + d(1), the second-ordernonlinear approximation D(2) = Ddc + d(1) + d(2), and asolution obtained by Euler integration D(∞). We see thatthe amplitude of the higher-frequency oscillations in D(∞)

is larger when E(t) is smaller. This asymmetry arisesbecause the permittivity of the material is higher with asmaller applied field. The second-order correction, d(2),restores this asymmetry (tuning) that the linear approxi-mation fails to capture. Note that the relative phase of thelinear response d(1) and the nonlinear response d(2) deter-mines whether the nonlinearity enhances the amplitude ofD(t) or diminishes it.

Formally, we model the behavior of a material by aVolterra-series expansion, which expresses a nonlinearprocess by a power-series expansion of the response ofa system in powers of the stimulus [46]. The nth termin the expansion, d(n)(t), is proportional to the nth powerof e(t). The relationship between these variables can beexpressed in either the frequency domain or the timedomain. The terms in the frequency-domain Volterra seriescan be written

d(1)

k (ω) = ε0ε(1)

jk (ω)ej (ω), (1)

d(2)

l (ω) = ε0

∫dω1

∫dω2 δ(ω − ω1 − ω2)

× ε(2)

jkl (ω1, ω2)ej (ω1)ek(ω2), (2)

d(n)

kn+1(ω) = ε0

∫dω δ(ω − �ω)ε

(n)

k1,...,kn+1(ω)

n∏j =1

ekj (ωj ),

(3)

where δ is the Dirac delta function, and ε(n)(ω) is thenth-order permittivity. In this formulation, the nonlinearpermittivity tensors depend on dc bias field, and in thecase of hysteretic materials, also depend on the past his-tory of the applied bias field. Equation (3) states that when

044026-3

Page 4: Measurements of Nonlinear Polarization Dynamics in the

AARON M. HAGERSTROM et al. PHYS. REV. APPLIED 13, 044026 (2020)

the material is stimulated by n sinusoidal signals whosefrequencies sum to ω, the amplitude of the mixing prod-uct d(n)(ω) is proportional to the nth-order permittivityε(n)(ω). To express the model more compactly, we defineω = (ω1, ω2, . . . , ωn) and �ω = ω1 + ω2 + · · · + ωn.

Using a Volterra-series approach, we develop an equiv-alent circuit theory to describe how a nonlinear dielec-tric perturbs a waveguide mode. We follow the logicof Ref. [47], which derives the well-known telegraphersequations from Maxwell’s equations in a single-modewaveguide. Reference [47] defines a voltage V and a cur-rent I that are proportional to the transverse electric fieldeT and the transverse magnetic field hT, respectively:

hT = I(z)I0

H(x, y), (4)

eT = V(z)V0

E(x, y). (5)

Here, H and E are normalized transverse fields that donot depend on z. According to Ref. [47], there is not auniversally accepted way to determine the normalizationconstants I0 and V0. However, they are partially determinedby a power-based normalization convention: in the situa-tion where there is a wave propagating in one direction,the power traveling down the waveguide is given by VI∗.

We make an additional simplifying assumption, which isthat our waveguide supports a transverse electric-magnetic(TEM) mode whose electric and magnetic fields are inthe plane perpendicular to the propagation direction. Thisapproximation is good so long as the gap width W ismuch less than the wavelength of the guided mode [48].To model the nonlinear material properties, we includethe first two terms in the Volterra-series expansion of d,and assume the materials involved have a linear magneticresponse. From energy-based arguments, in direct anal-ogy with Ref. [47], we obtain a generalized form of thetelegrapher’s equations:

ddz

V(ω, z) = −Z(ω)I(ω, z), (6)

ddz

I(ω, z) = −Y(1)(ω)V(ω, z) − N (ω, z), (7)

N (ω, z) =∫

dω δ(ω − �ω)Y(2)(ω)

2∏j =1

V(ωj , z). (8)

These equations are parameterized by a distributedadmittance Y(1)(ω), nonlinear admittance Y(2)(ω), andimpedance Z(ω). These coefficients are given byexpressions involving surface integrals over the x-y plane:

Z(ω) = iω|I0|2

∫dS

[μjk(ω)H∗

j (ω)Hk(ω)]

, (9)

Y(1)(ω) = iω|V0|2

∫dS

(1)

jk (ω)E∗j (ω)Ek(ω)

], (10)

Y(2) (ω1, ω2) = i(ω1 + ω2)

|V0|2V0

∫dS

×[ε

(2)

jkl (ω1, ω2)E∗j (ω1+ω2)Ek(ω1)El(ω2)

].

(11)

Equations (9) and (10) agree with the expressions givenin Ref. [47] in the special case that there are no electricor magnetic fields in the propagation direction. Refer-ence [47] did not consider the case of nonlinear mate-rials and did not offer any expression for Y(2). Similarexpressions to Eq. (11) can be derived for nonlinear mag-netic materials, or materials with a higher-order dielectricnonlinearity.

A. Solution of the nonlinear telegrapher’s equations

To a good approximation, we can solve Eqs. (6)–(8)perturbatively, treating Y(2) as a small parameter [49]. Inour experiment, the microwave-frequency electric fieldsin the waveguide are much smaller than the electric fieldrequired to change the permittivity of the nonlinear dielec-tric film by a substantial amount. We assume a solutionof the form V(ω, z) = V(1)(ω, z) + V(2)(ω, z), I(ω, z) =I (1)(ω, z) + I (2)(ω, z), where I (1) � I (2) and V(1) � V(2).The first-order voltage and current, V(1) and I (1), repre-sent the externally-applied signals, while the second-ordervoltage and current V(2) and I (2) represent the waves gen-erated by nonlinear mixing in the dielectric material. Tolowest order, we set Y(2) = 0, and Eqs. (6)–(8) reduce tothe well-known telegrapher’s equations. These equationsadmit traveling wave solutions:

V(1)(z) = V+e−γ z + V−eγ z, (12)

I (1)(z) = V+

Z0e−γ z − V−

Z0eγ z. (13)

Here, γ =√

ZY(1) is the propagation constant and Z0 =√Z/Y(1) is the characteristic impedance. The unknown

coefficients V+ and V− are determined by boundary condi-tions. Once these coefficients are determined, we approxi-mate the nonlinear term in Eq. (8) by

N (ω) ≈∫

dω δ(ω − �ω)Y(2)(ω)

2∏j =1

V(1)(ωj ). (14)

With this approximation, Eqs. (6)–(8) can be solved bya Green’s function approach. To construct a solution,we seek functions GV and GI that satisfy the following

044026-4

Page 5: Measurements of Nonlinear Polarization Dynamics in the

MEASUREMENTS OF NONLINEAR POLARIZATION... PHYS. REV. APPLIED 13, 044026 (2020)

differential equations:

dGV(z, z′)dz

= −Z(ω)GI (z, z′), (15)

dGI (z, z′)dz

= −Y(1)(ω)GV(z, z′) + δ(z − z′). (16)

Here, δ(z − z′) is the Dirac δ function and models apointlike current source at position z′. Once GI (z, z′) andGV(z, z′) have been computed, the voltage and current maybe determined by integrating the contributions of wavesgenerated by the nonlinear current source N (ω, z′) at eachposition along the transmission line (with length l).

V(2)(ω, z) = −∫ l

0dz′ GV(z, z′)N (ω, z′), (17)

I (2)(ω, z) = −∫ l

0dz′ GI (z, z′)N (ω, z′). (18)

B. Boundary conditions

We use boundary conditions based on the experimen-tally accessible quantities a and b to solve for both thefirst-order solution, V(1) and I (1), and for the form of theGreen’s function. The wave parameters a and b are definedwith respect to an arbitrary, real, reference impedance Zrefin terms of the current I and voltage V [47].

an ≡ V + ZrefI2

, (19)

bn ≡ V − ZrefI2

. (20)

The subscript n refer to ports 1 and 2, which we defineto be the left and right edges of the waveguide. Note thatthe current is defined as flowing away from the port, sofor a2 and b2, I has the opposite sign as in Eqs. (13). TheLSNA calibration that we describe in Sec. III allows us tomeasure the a and b waves at the edge of our waveguides.In our analysis, we solve for the coefficients V+ and V−assuming that the a wave at each end of the waveguide isgiven by its measured value.

We employ slightly different boundary conditions forthe Green’s function than the first-order solution. Green’sfunctions have homogeneous boundary conditions [50],and so we set the a wave to 0, rather than its measuredvalue. In principle, we should also account for an inho-mogeneous (proportional to a) contribution to the b wavesat the frequencies of the mixing products. In our case, wefind that the a waves at these frequencies are negligible.We expect the a waves to be small because the LSNA isnot sourcing any power at the frequencies of the mixingproducts, and we did not employ any highly reflective com-ponents in our setup. We found that including measured a

waves in our analysis seemed to add noise to our results,because in our experiment these waves are near the limitof detection.

If the left side of a waveguide is connected to a probe,we can assume as a boundary condition that a wave at z =0 has the value a1. We arrive at the following relationship:

V(0) + ZrefI(0) = 2a1. (21)

Meanwhile, if the right side of a waveguide is connectedto a probe, we can assume as a boundary condition that awave at z = l has the value a2.

V(l) − ZrefI(l) = 2a2. (22)

In our measurements, we employ both two-port transmis-sion lines, and one-port terminated transmission lines. Oneend of the terminated waveguide is connected to the probe,and the other is terminated by an open cirucit with a smallcapacitance compared to the rest of the transmission line.In terms of the impedance associated with this termina-tion, Zl, the boundary condition on the right end of theterminated waveguide is

V(l) = ZlI(l). (23)

The equations given in Sec. II A, with the boundary con-ditions in this section, allowed us to extract the nonlin-ear admittance from the measured wave parameters. InSec. III, we discuss how to calibrate our measurements sothat these boundary conditions can be applied.

III. EXPERIMENTAL APPROACH

A. Nonlinear calibration

To interpret LSNA measurements of our on-waferwaveguides, we need to characterize the relationshipbetween ar

n and brn, the raw signals measured by the

receivers of an LSNA, and an and bn, the forward andbackward waves at the edges of an on-wafer device. Thisrelationship (calibration) accounts for signal distortion byall of the circuitry between the receivers of the LSNA andthe sample, including the internal circuitry of the LSNA,the coaxial cables connecting the test ports of the LSNAto the probes, the bias tees and the wafer probes. We use alinear model to account for this distortion.

Figure 3 schematically shows the error model that weemploy. We define three reference planes: the raw mea-surements, a coaxial reference plane at the connector to thewafer probes, and an on-wafer reference plane. As we dis-cuss, this two-tiered structure is necessary for phase andamplitude calibrations. The 2 × 2 complex-valued matri-ces X1, Y1, X2, and Y2 are in T-parameter form, anddescribe how the a and b waves change between referenceplanes.

044026-5

Page 6: Measurements of Nonlinear Polarization Dynamics in the

AARON M. HAGERSTROM et al. PHYS. REV. APPLIED 13, 044026 (2020)

(a)

(b)

FIG. 3. Error model. (a) Physical location of the referenceplanes. The raw wave parameters, ar

n and brn, are measured by

the detectors of the LSNA. The wave parameters at the coaxialplane, ac

n and bcn are measured at the ends of the coaxial connec-

tors that attach to the wafer probes. The wave parameters of thedevice under test (DUT), an and bn, are measured at the waferprobe tips. (b) Schematic representation of the error model relat-ing the wave parameters at the different reference planes. Theerror model has 16 parameters defined independently at each fre-quency point: X1, Y1, X2 and Y2, are all 2 × 2 complex-valuedmatrices.

The LSNA calibration is a generalization of a tradi-tional linear network analyzer calibration that also allowsfor phase and power calibrations. We model the rela-tionship between the measured T parameters of a lineardevice (M ) and their true values (T) by M = X T Y. In ourtwo-tiered calibration model, X = X1 X2 and Y = Y2 Y1.By measuring a set of linear passive devices with known

(b)

(a)

FIG. 4. Measurement configurations. (a) For broadband mea-surements, we use coplanar waveguides, and excite the waveg-uide from both edges. (b) For narrow-band measurements, weuse a one-port configuration, with waveguides terminated by agap to ground. This allows us to use filters to minimize spurioussignals generated within the LSNA.

T parameters, we determine seven of the eight complexnumbers in X and Y [51]. The remaining ambiguity is afeature of any linear error model determined solely frommeasurements of passive devices [52]. One way to resolvethis ambiguity is to assume that either X or Y representsa reciprocal network [51]. However, the first-tier errorparameters X1 and Y1 include the LSNA, and the signalsmeasured by its receivers are subject to an unknown, andpotentially nonreciprocal, scaling in magnitude and fre-quency. To determine the phase and amplitude of the aand b waves, one must measure these quantities directlyat some reference plane [52].

We determine the phase and amplitude by measuringthese quantities at the coaxial reference plane immediatelybefore the wafer probes. This reference plane is neces-sary because we lack on-wafer phase and power standards.Apart from an overall phase and amplitude, we use ashort-open-load-thru (SOLT) calibration to determine theparameters in X1 and Y1 up to 40 GHz [53]. To establishthe cross-frequency phase, we use a comb generator thatproduces periodic pulses in the time domain. The combgenerator has a repetition rate of 10 MHz, and its Fourierspectrum consists of evenly spaced harmonics from 10MHz up to 40 GHz. We attach a comb generator to port 3 ofthe LSNA for the duration of the experiment to serve as aphase reference. Thus, the phase of any signal at ports 1 or2 could be determined by comparing to the phase of the ref-erence at port 3, as long as all of the frequencies involvedare multiples of 10 MHz. We also need to account for thefixed phase relationship between the phase reference andthe signals measured at ports 1 and 2, and this relation-ship is determined by measuring a second comb generatorattached to coaxial port 1. To calibrate the power of the aand b waves, we measure the power sourced by the LSNAat port 1 at the coaxial reference plane.

B. On-wafer calibration

In order to interpret our mixing products in physicalterms, we need to understand the relationship between themeasured a and b waves, and the guided modes of on-wafer waveguides. This involves measuring waveguideswith well-understood properties, and constructing an errormodel from these measurements. Our on-wafer calibrationstandards are shorts, loads, and transmission lines, allow-ing for a multiline thru-reflect-line (MTRL) calibration[51]. The on-wafer calibration standards had an identi-cal cross-sectional geometry to the test devices on theBST chip, but are patterned directly on a substrate withnegligible dielectric dispersion at microwave frequencies[tan(δ) < 0.005]. The calibration substrate also has a simi-lar relative permittivity (ε(1)′ ≈ 24) to the substrate that theBST film was deposited on (ε(1)′ ≈ 22). We do not needany phase or power calibration to construct our second-tier error parameters, X2 and Y2. These matrices represent

044026-6

Page 7: Measurements of Nonlinear Polarization Dynamics in the

MEASUREMENTS OF NONLINEAR POLARIZATION... PHYS. REV. APPLIED 13, 044026 (2020)

wave propagation through the wafer probes, which aremade of reciprocal materials. This reciprocity removes anyambiguity in X2 and Y2.

Transmission lines make good calibration standardsbecause their scattering parameters can be derived fromfundamental electromagnetic theory. If one chooses Zref =Z0, then the T-parameter model of a transmission line isdescribed by one unknown parameter, γ . Let M1 and M2be T-parameter measurements of two transmission lineswith lengths l1 and l2 = l1 + �l. Reference [51] showedthat the eigenvalues of matrix M1(M−1

2 ) are exp(±γ�l).This relationship means that, without any calibration, wecan determine the propagation constant γ from measure-ments of two transmission lines with different lengths. Bymeasuring a reflect standard and set of transmission lines,we can determine the error matrices X and Y [51,54].

In practice, it is more convenient to set the referenceimpedance Zref to 50 � than Z0. The coaxial cables andprobes leading up to the sample are designed to have acharacteristic impedance near 50 �, and hence it is a nat-ural choice for many measurements. To set the referenceimpedance, we make use of the fact that Z0 = γ /Y. Ourcalibration transmission lines are fabricated on a substratewith little to no dielectric dispersion. In this case, we makethe approximation Y1 = iωC0, where C0 is a frequencyindependent capacitance per unit length. This model ofthe admittance is justified by Eq. (10) under the assump-tion that the mode is TEM, and therefore the electric fieldis frequency-independent. To estimate the capacitance perunit length C0, we use the method described in Ref. [55].We then modify the error matrices X and Y to set thereference impedance to 50 � [56]. By insisting that thecapacitance is frequency independent, we implicitly deter-mine the constant V0 in Eq. (10). The power normalizationcondition determines the product I∗

0 V0, so the waveguidecurrent and voltage are unambiguously defined in terms ofthe guided mode electric and magnetic fields by the MTRLcalibration.

In principle, there is some error in the calibrationbecause the transmission lines on the BST test chip hasa higher capacitance per unit length than the calibrationstandards by a factor of about 3/2 [56]. The systematicerror introduced by this discrepancy should not influenceour linear-permittivity estimates because our estimator ofγ is based on the eigenvalues of measurements of pairs oftransmission lines, and is therefore independent of the cal-ibration [51]. To assess the magnitude of this uncertaintyin our nonlinear measurements, we modeled the effect ofa small parasitic capacitance to our analysis. As expected,this added error did not change our linear-permittivity esti-mates. Our nonlinear-permittivity estimates changed byless than the error bars. So, we conclude that the differencein distributed capacitance between our test structures andour calibration standards does not significantly degrade thequality of our data. We caution that for thicker films, or

films with higher permittivity, some modifications to thiscalibration approach may be necessary to account for thiseffect.

C. Measurement configurations

Figure 4 shows two experiment configurations we useto measure 2nd order nonlinear mixing products. We usethese two configurations because in our nonlinear mea-surements we encounter a trade-off between dynamic rangeand bandwidth. The dominant source of error in our non-linear measurements is the nonlinearity of the sources andreceivers in the LSNA, which generated their own mixingproducts at exactly the frequencies of interest. Conse-quently, in our frequency-swept nonlinear measurements[Fig. (4a)], we have about 10 dB of dynamic range eventhough the nonlinear signals from our sample are about 40dB above the noise floor. In our voltage-swept nonlinearmeasurements [Fig. (4b)], we improve the dynamic rangeby fixing the source frequencies, using one-port devices,and employing filters to suppress spurious signals from theLSNA.

In the one-port measurements of the terminated trans-mission lines, we use a low-pass filter to remove signalcontent at 2ω1, 2ω2, and ω1 + ω2 generated by the sources.Likewise, high-pass filters placed in front of the receiversattenuated the signals at ω1 and ω2 and prevented mix-ing in the receivers. With these filters, the dynamic rangeincreased to about 30 dB for the frequencies under test ina narrow band around 7.8 GHz.

D. Sample preparation

To facilitate our measurements, we fabricate coplanar-waveguide transmission-line structures and coplanarwaveguide offset open structures on the film (Fig. 5). Weprepare our sample through standard deposition and lithog-raphy techniques. The BST film is grown to a thickness of200 nm by pulsed laser deposition on the (001) surfaceof a (LaAlO3)0.3(SrTaAlO6)0.7 (LSAT) substrate heatedto 760 ◦C. The KrF ablation laser produced an incidentlaser fluence of 0.4 J/cm2 at a repetition rate of 5 Hz.The film growth occurred under a background oxygenpressure (PO2) of 40 Pa (30 mTorr). We pattern the teststructures on the BST film by photolithography, followedby electron-beam deposition of a 500-nm-thick gold layerwith a 10-nm-thick titanium adhesion layer. We fabricatedthe calibration standards on the (001) surface of a LaAlO3(LAO) substrate by the same process.

Photographs of the BST chip are shown in Fig. 5. Thetest devices on the BST thin film consisted of a collectionof CPW structures and offset open structures of varyinglengths: l = {0.420, 0.660, 0.880, 1.340, 2.240, 4.020,and 7.500} mm for the CPWs, and l = {0.720, 1.100, 1.620,2.260, 4.700, and 6.240} mm for the offset opens. TheCPWs had 20 μm wide center conductors, 200 μm wide

044026-7

Page 8: Measurements of Nonlinear Polarization Dynamics in the

AARON M. HAGERSTROM et al. PHYS. REV. APPLIED 13, 044026 (2020)

(b)

(c)

(a)

0.5 mm 0.5 mm

FIG. 5. Our BST sample. (a) Photograph of the BST chip withthe waveguide structures fabricated on it. Because the chip islarger than the microscope’s field of view, this photograph isstitched together from many smaller images. (b) Example of atwo-port device, a coplanar-waveguide (CPW) transmission linewith a length of 0.880 mm. (c) Example of a one-port device, acoplanar waveguide offset open with a length of 1.100 mm.

ground planes, and a 5 μm gaps between the center con-ductor and the ground planes. The offset open structureshave the same cross-sectional geometry as the CPWs, butare terminated at one end with a 40 μm gap to the groundplane.

E. Analysis overview

After performing the two-tiered calibration described inSec. III, we perform both a linear and nonlinear charac-terization of our BST film to determine the permittivitiesε(1) and ε(2) from the measured wave parameters. Wetreat these permittivities as complex frequency-dependentscalars because we do not have enough informationto independently determine all of the tensor elements.Because of the linear relationships between the admit-tances Y(1) and Y(2) and the permittivities ε(1) and ε(2), ourstrategy is to determine Y(1) and Y(2) from measurements,and then use simulations to relate these quantities to ε(1)

and ε(2).The first step is a linear characterization to determine

the circuit parameters Y(1) and Z as a function of dc biasvoltage and frequency. To do this, we measure a collectionof transmission lines fabricated on the materials we wishto characterize, and make use of the fact that, if M1 andM2 are T-matrix measurements of two transmission lineswhose lengths differ by �l, the eigenvalues of M1(M−1

2 )

are exp(±γ�l) (Sec. III B). On our BST thin-film sample,we fabricate seven CPWs, leading to 21 different mea-surements (six independent measurements) of the quantityγ�l, and use linear regression against �l to estimate γ .The linear admittance Y(1), is related to the propagationconstant by γ =

√ZY(1). Under the assumption of single-

mode TEM propagation, Z depends only on the CPWgeometry and metal conductivity [47,57]. We estimate thevalue of Z with the model described in Ref. [57]. From the

(a)

(b)

(c)

FIG. 6. Example of the nonlinear admittance estimation frommeasurements. (a) Measured mixing product, b(ω1 + ω2) of aCPW of length 1.340 mm. ω2 = 2π × 5 GHz. (b) bmodel(ω1 +ω2), the hypothetical value of the mixing product underthe condition Y(2)(ω1, ω2) = 1. The estimated uncertainty in(b) from uncertainty in Y(1) is smaller than the width ofthe line. (c) Y(2)(ω1, ω2) estimated by Y(2)(ω1, ω2) = b(ω1 +ω2)/bmodel(ω1 + ω2). Error bars in (a) and (c) represent a 95%confidence interval estimated from the distribution of Y(2) esti-mates from CPWs of different lengths. We omit error bars onindividual data points for visual clarity.

theoretical Z and measured γ , we can compute Y1. We usethese linear circuit parameters in the nonlinear analysis.

Figure 6 illustrates our process for estimating Y(2) frommixing product measurements. To measure the mixingproducts, we apply microwave power from both ends of theCPW as shown in Fig. 1(b). The incident power from port2 is about −4 dBm, and the incident power from port 1 var-ied with frequency between −10 and −23 dBm. Accordingto Eqs. (6)–(8), a waveguide excited by signals at ω1 andω2 generates voltage and current waves at ω1 + ω2. Thekey insight from Sec. II that enables our nonlinear analysisis that the amplitudes of these mixing products are directlyproportional to Y(2)(ω1, ω2). In other words, the complexamplitude of the measured mixing products, b(ω1 + ω2),

044026-8

Page 9: Measurements of Nonlinear Polarization Dynamics in the

MEASUREMENTS OF NONLINEAR POLARIZATION... PHYS. REV. APPLIED 13, 044026 (2020)

is given by

b(ω1 + ω2) = bmodel(ω1 + ω2)Y(2)(ω1, ω2), (24)

where bmodel(ω) is the hypothetical strength of the mix-ing products under the condition that Y(2) = 1. To estimateY2(ω), we measure the mixing product b2(ω) at ω1 + ω2for our seven transmission lines, and use linear regres-sion against bmodel(ω1 + ω2) to determine Y2(ω). We alsomeasure Y2 at the other second-order mixing frequencies:|ω1 − ω2|, 2ω1, and 2ω2, but we focus on ω1 + ω2 forbrevity.

The quantity bmodel is fully determined by the linearcharacterization. The first step to calculating bmodel is toreconstruct the spatial variation of the incident voltagewaves from the measured values a1(ω1), a2(ω1), a1(ω2),and a2(ω2). Using the boundary conditions from Sec. II B,we can determine the traveling wave amplitudes V+ andV−. From those amplitudes, and Eq. (12), we can estimatethe spatial shape of the nonlinear current source densityN (z) through Eq. (8). At this point, we solve for bmodelusing a Green’s function approach. The Green’s functionis given in Sec. II A, and the unknown coefficients in theGreen’s function are determined from Eqs. (15) to (16)with the boundary conditions from Sec. II B. With theGreen’s function known, and the nonlinear current densityN (z) known, the currents and voltages at the frequenciesof the mixing products are determined through Eqs. (17)and (18). These currents and voltages in turn can be usedto determine bmodel.

The last step in our analysis is to relate the admittancesY(1) and Y(2) to the permittivities ε(1) and ε(2). To sim-plify this relationship, we note that the electric field of aTEM mode is frequency independent. Therefore, the rela-tionship between the admittances Y(1) and Y(2) and thepermittivities ε(1) and ε(2) is also frequency independent.In our geometry, the linear admittance and the permittivityare approximately related by Y(1)(ω)/iω = c0 + c1ε

(1)(ω)

[47,58,59]. The constants c0 and c1 can be computedthrough quasistatic finite-element methods. The nonlinearadmittance Y(2)(ω) is linearly related to the strength of thedielectric nonlinearity, ε(2)(ω). To relate the nonlinear per-mittivity ε(2) to the admittance, Y(2), we assume that theelectric field in our waveguide (both microwave and dc)is given by the voltage between the center conductor andground plane, divided by the gap width, W. This approx-imation leads to the relation Y(2) = iω(c1/W)ε(2). Fromthe agreement between our data and model (Fig. 8), weestimate that the error introduced by this approximation isabout 20% in the worst-case scenario (strong dc bias).

IV. THERMODYNAMIC MODEL

We compare our results to a physical model that pre-dicts the nonlinear response of our BST thin film. To

be comparable with our experiment, the model needs todescribe the time dependence of the polarization, describethe change of the polarization in response to a large appliedfield, and be amenable to a small-signal Volterra-seriesexpansion. Mesoscale models meet these requirements,but require detailed knowledge of the nanostructure ofthe material. The small-signal requirement rules out mod-els like the Ishibashi-Orihara model, which assume theamplitude of the time-varying field is large enough to com-pletely pole the material [60,61]. Many models, like thespherical-random-bond-random-field model, are not suit-able because they do not describe the time dependence ofthe material [62].

In BST thin films, the relaxation dynamics of PNRslead to permittivity that varies over a broad range of fre-quencies, often modeled by a distribution of relaxationtimes [17–20]. To our knowledge, there is no universallyaccepted first-principles method to predict the distributionof relaxation times for a given material. Even a concep-tual picture of PNR dynamics remains a subject of currentresearch [13,14].

Given the complexity of a first-principles model, wetake a phenomenological approach to interpreting ourmeasurements. We chose a model that uses a heuristicinterpretation of PNR dynamics and the Landau-Ginzburg-Devonshire (LGD) free energy. This model, introduced inRef. [37], incorporates the salient features of our data: tun-ing of the permittivity with an applied electric field and adistribution of relaxation times. It was used to describe thetemperature dependence and frequency dependence of thethird harmonic up to 100 kHz in Pb(Mg1/3Nb2/3)O3. Weexpect this model to also apply to BST in the tens of GHzbecause BST has qualitatively similar tuning behavior anddispersion.

The model describes the tuning processes in terms ofthe LGD free energy. In its simplest form, the LGD freeenergy U can be expressed in terms of the polarizationP and the phenomenological parameters α and β as U =(1/2)αP2 + (1/4)βP4. The one-dimensional model is jus-tified because in our CPW structures, the electric fieldsare mostly in plane. The distribution of relaxation timesis modeled by a function f (t). The time evolution of thepolarization is described by

P(t) = ε0�ε

∫dt′ f (t − t′)

[E(t′) − βP3(t′)

], (25)

where �ε = (ε0α)−1. Essentially, Eq. (25) models thepolarization of a macroscopic region as a sum manycoupled relaxation processes within identical anharmonicfree-energy wells.

To model the nonlinear permittivity of our sample, wedevelop a perturbative expansion of Eq. (25) in pow-ers of the electric field. We begin by writing the electricfield as E(t) = Edc + e(t) and the polarization as P(t) =

044026-9

Page 10: Measurements of Nonlinear Polarization Dynamics in the

AARON M. HAGERSTROM et al. PHYS. REV. APPLIED 13, 044026 (2020)

Pdc + p (1)(t) + p (2)(t) + p (3)(t) + · · · . We then substitutethese series expansions into Eq. (25), and iterativelysolve for p (n) in terms of p (1), . . . , p (n−1). Setting e(t) = 0and p(t) = 0, Eq. (25) admits a closed-form solution forPdc if β > 0: Pdc = ε0�εENLT (Edc/ENL), with T (x) =sinh

[(1/3) sinh−1 (3 x)

][63,64]. We introduce the param-

eter ENL = (4α3/3β

)1/2 with units of electric field todescribe the strength of the nonlinearity. The function T(x)has a sigmoidal shape, with a slope of unity for small argu-ments. With the dc polarization determined, calculating theFourier transform of Eq. (25) yields

ε(1)(ω) = ε∞ + �εfE(ω), (26)

ε(2)(ω) = −4Pdc

ε0(ENL)2 fE(�ω)fE(ω1)fE(ω2). (27)

For convenience, we define the function fE(ω) =f (ω)/

[1 + 4T2(Edc/ENL)f (ω)

], where f (ω) is the Fourier

transform of f (t) in Eq. (25). We also add an additionalparameter ε∞, which describes the high-frequency limitof the permittivity [17,18]. Qualitatively, both the realand imaginary parts of ε1(ω) peak at Edc = 0 V m−1. Theparameter ENL can be interpreted as a characteristic dcfield strength necessary to cause appreciable tuning (about64% at dc). On the other hand, ε2(ω) vanishes at Edc =0 V m−1, and is antisymmetric about 0 V m−1. Note thatthe second-order permittivity is inversely related to ENL,so this parameter quantifies both the width of the tuningcurve and the strength of the second-order nonlinearity.

V. RESULTS AND DISCUSSION

As we expect from the model, the real and imaginaryparts of the frequency-dependent permittivity tuned withthe application of an electric field (Fig. 7). As we showin Fig. 1, we apply the bias voltage between the cen-ter conductor and the ground planes. This resulted in amostly in-plane dc and microwave electric field, which aresuperimposed onto one another. To check if the BST sam-ple had hysteresis, we sweep the electric field in a cycle0 V μm−1 → 8 V μm−1 → −8 V μm−1 → 8 Vμm−1 →0 Vμm−1. No hysteresis larger than the measurementuncertainties is present. Both the real and imaginary partsof the permittivity have a peak near 0 V μm−1, anddecrease as the magnitude of the bias field increases.

Equation (26) captures the basic features of the electricfield and frequency dependence of the data. We assumethat f (ω) is given by the Cole-Cole function f (ω) =[1 + (iωτ0)

a]−1 [19]. The parameter τ0 is a phenomeno-logical characteristic timescale. The parameter a describesthe shape of the distribution of relaxation times, with thevalue a = 1 corresponding to the Debye model (one relax-ation time), and smaller values describing wider distribu-tions. This parameter also describes the limiting behavior

(a)

(b)

NL

FIG. 7. Linear permittivity, ε(1) = ε(1)′ − iε(1)′′ as a function offrequency, and as a function of bias electric field at 20 GHz. Aninterval of width ENL centered around Edc = 0 V m−1 is indi-cated. The error bars (gray) are the 95% confidence interval fromthe standard error of the mean. (a) Real part, ε(1)′. (b) Imaginarypart, ε(1)′′.

of the dielectric loss at low frequencies ε(1)′′ ∝ ωa. Themodel illustrates the close link between the dielectric lossand the tuning. The parameters τ0 = 0.49 ps, a = 0.50,ENL = 10.7 V μm−1, ε∞ = 106, and �ε = 354 are esti-mated by a least-squares fit. Equation (25) states thatthe polarization relaxes to the lowest free energy by asuperposition of coupled gradient descent processes. Thefree-energy gradient is steeper at large dc fields, so therelaxation is faster. This field-dependent relaxation speedmanifests as a field-dependent loss tangent, and the imag-inary part of the permittivity tunes more strongly (66% at8 V μm−1) compared with the tuning of the real part (33%at 8 V μm−1).

As in the linear case, the model predicts the salient fea-tures of the dependence of ε(2)(ω1, ω2) on bias field andfrequency (Fig. 8). Note that the black line in Fig. 8 isa prediction, rather than a fit. The parameters in Eq. (27)are determined from the linear-permittivity data in Fig. 7.The model predicts the relaxationlike trend of ε(2)(ω1, ω2)

with frequency, but somewhat overestimates the value ofε(2)(ω1, ω2). More notably, even though we see no hys-teresis in the linear data, we see some hysteresis in thefield dependence of ε(2). Figure 8(c) shows that the abso-lute value of ε(2) takes differing values on increasing anddecreasing voltage sweeps, and that this difference is largerthan the measurement uncertainties [65].

044026-10

Page 11: Measurements of Nonlinear Polarization Dynamics in the

MEASUREMENTS OF NONLINEAR POLARIZATION... PHYS. REV. APPLIED 13, 044026 (2020)

(a)

(b)

(c)

FIG. 8. Second-order nonlinear permittivity, ε(2) = ε(2)′ −iε(2)′′, inferred from the ω1 + ω2 mixing product. In thefrequency-dependent measurements, we vary the angular fre-quency of one of the stimulus signals, ω1, while the otherstimulus frequency, ω2, is held constant at 2π × 5 GHz. In ourbias-swept measurements we set ω1 = 2π × 7.84 GHz and ω2 =2π × 7.76 GHz. To facilitate comparison to Fig. 7, we indicatean interval of width ENL centered around the Edc = 0 V μm−1.The error bars (gray) are the 95% confidence interval from thestandard error of the mean. (a) Real part, ε(2)′. (b) Imaginary part,ε(2)′′. (c) Absolute value of ε(2) near 0 V μm−1.

Extending this model to include spatial variation in thepolarization could improve the accuracy of the model andprovide physical insight. The electric field varies overthe scale of the CPW gaps, causing different regions ofthe film to experience different dc bias. This large-scaleinhomogeneity may impact the voltage dependence ofthe linear and nonlinear admittance. The polarization alsovaries on the scale of nanometers to hundreds of nanome-ters due to grain boundaries and PNRs. We speculate thatinteractions among PNRs might lead to different nonlinearcoupling [other forms of the P3 term in Eq. (25)]. Whilewe conclude that Eq. (25) predicts most of the featuresof our data, we cannot exclude other models of nonlinear

coupling. The samples that we characterize have rela-tively low dielectric loss, leading to a fairly flat frequencydependence for ε(2)(ω1, ω2). In the future, we hope to eval-uate more lossy samples and extend our measurementsto higher frequencies. In this report, our main findingis that the measurement technique allows for broadband,calibrated, and model-independent measurements of mate-rial nonlinearity.

VI. CONCLUSIONS

In conclusion, we demonstrate the ability to measurenonlinear permittivity up to 40 GHz and develop a ther-modynamic model informed by the data. We expect thistechnique to apply to a broad range of materials becausethe measurements and analysis used to produce Figs. 7 and8 made no physical assumptions about the sample. Thesenonlinear-permittivity measurements provide a measur-able quantity, independent of linear permittivity, that canbe used to probe the complex dynamics of ferroelectricand related materials. The dynamics of these materialsarise due to an interaction of many mechanisms includingelectrostatic energy, gradient energy, strain, and flexoelec-tricity. Nonlinear-permittivity measurements are a way totest models that include these different physical mecha-nisms. There are endless possibilities for manipulating thecoupling between these effects to control domain structuresand dielectric properties. For these materials to be usedin high-speed electronics applications, it is necessary tounderstand their linear and nonlinear dynamics under anapplied field.

The experimental data presented here is availableonline [66].

ACKNOWLEDGMENTS

We thank Dylan F. Williams, Alirio S. Boaventura,and Richard Chamberlin for advice and training onLSNA measurements. We also thank Angela Stelson, AricSanders, Thomas Mitchell Wallis, and Samuel Berwegerfor careful and constructive feedback on the manuscriptpreparation. Financial assistance from Award Numbers70NANB16H003, 70NANB17H301, 70NANB18H165,and 70NANB18H006 from U.S. Department of Commerceat National Institute of Standards and Technology (NIST)supported this work. Some of this research was performedwhile Aaron M. Hagerstrom held an NRC Research Asso-ciateship award at NIST. Contributions to this article byworkers at NIST, an agency of the U.S. Government, arenot subject to U.S. copyright.

[1] M. E. Lines and A. M. Glass, Principles and Applicationsof Ferroelectrics and Related Materials (Oxford UniversityPress, Oxford, UK, 2001).

044026-11

Page 12: Measurements of Nonlinear Polarization Dynamics in the

AARON M. HAGERSTROM et al. PHYS. REV. APPLIED 13, 044026 (2020)

[2] A. Gruverman and A. Kholkin, Nanoscale ferroelectrics:Processing, characterization and future trends, Rep. Prog.Phys. 69, 2443 (2005).

[3] S.-C. Chang, A. Naeemi, D. E. Nikonov, and A. Gru-verman, Theoretical Approach to Electroresistance in Fer-roelectric Tunnel Junctions, Phys. Rev. Appl. 7, 024005(2017).

[4] S.-C. Chang, U. E. Avci, D. E. Nikonov, S. Manipatruni,and I. A. Young, Physical Origin of Transient NegativeCapacitance in a Ferroelectric Capacitor, Phys. Rev. Appl.9, 014010 (2018).

[5] A. I. Khan, K. Chatterjee, B. Wang, S. Drapcho, L. You,C. Serrao, S. R. Bakaul, R. Ramesh, and S. Salahuddin,Negative capacitance in a ferroelectric capacitor, Nat.Mater. 14, 182 (2015).

[6] G. Catalan, D. Jiménez, and A. Gruverman, Ferroelectrics:Negative capacitance detected, Nat. Mater. 14, 137 (2015).

[7] V. V. Zhirnov and R. K. Cavin, Nanoelectronics: Negativecapacitance to the rescue? Nat. Nanotechnol. 3, 77 (2008).

[8] S. Salahuddin and S. Datta, Use of negative capacitanceto provide voltage amplification for low power nanoscaledevices, Nano Lett. 8, 405 (2008).

[9] R. A. York, Multifunctional Adaptive Microwave Circuitsand Systems (Scitech Publishing, Raleigh, NC, 2009).

[10] A. K. Tagantsev, V. O. Sherman, K. F. Astafiev, J.Venkatesh, and N. Setter, Ferroelectric materials formicrowave tunable applications, J. Electroceram. 11, 5(2003).

[11] M. Sazegar, Y. Zheng, C. Kohler, H. Maune, M. Nikfalazar,J. R. Binder, and R. Jakoby, Beam steering transmitarrayusing tunable frequency selective surface with integratedferroelectric varactors, IEEE Trans. Antennas Propag. 60,5690 (2012).

[12] E. G. Erker, A. S. Nagra, Y. Liu, P. Periaswamy, T. R.Taylor, J. Speck, and R. A. York, Monolithic ka-bandphase shifter using voltage tunable BaSrTiO3 parallel platecapacitors, IEEE Microw. Guided Wave Lett. 10, 10 (2000).

[13] H. Takenaka, I. Grinberg, S. Liu, and A. M. Rappe, Slush-like polar structures in single-crystal relaxors, Nature 546,391 (2017).

[14] A. A. Bokov and Z.-G. Ye, Recent progress in relaxor fer-roelectrics with perovskite structure, J. Mater. Sci. 41, 31(2006).

[15] Z. Gu et al., Resonant domain-wall-enhanced tunablemicrowave ferroelectrics, Nature 560, 622 (2018).

[16] H. Akamatsu, Y. Yuan, V. A. Stoica, G. Stone, T. Yang,Z. Hong, S. Lei, Y. Zhu, R. C. Haislmaier, J. W. Free-land, L.-Q. Chen, H. Wen, and V. Gopalan, Light-ActivatedGigahertz Ferroelectric Domain Dynamics, Phys. Rev. Lett.120, 096101 (2018).

[17] T. Tsurumi, J. Li, T. Hoshina, H. Kakemoto, M. Nakada,and J. Akedo, Ultrawide range dielectric spectroscopy ofBaTiO3-based perovskite dielectrics, Appl. Phys. Lett. 91,182905 (2007).

[18] T. Teranishi, T. Hoshina, H. Takeda, and T. Tsurumi, Polar-ization behavior in diffuse phase transition of BaxSr1−xTiO3ceramics, J. Appl. Phys. 105, 054111 (2009).

[19] J. C. Booth, I. Takeuchi, and K.-S. Chang, Microwave-frequency loss and dispersion in ferroelectric Ba0.3Sr0.7TiO3thin films, Appl. Phys. Lett. 87, 082908 (2005).

[20] V. Goian, S. Kamba, N. Orloff, T. Birol, C. H. Lee, D.Nuzhnyy, J. C. Booth, M. Bernhagen, R. Uecker, and D.G. Schlom, Influence of the central mode and soft phononon the microwave dielectric loss near the strain-induced fer-roelectric phase transitions in Srn+1TinO3n+1, Phys. Rev. B90, 174105 (2014).

[21] L. M. Garten, M. Burch, A. S. Gupta, R. Haislmaier, V.Gopalan, E. C. Dickey, and S. Trolier-McKinstry, Relaxorferroelectric behavior in barium strontium titanate, J. Am.Ceram. Soc. 99, 1645 (2016).

[22] M. Tyunina and J. Levoska, Coexistence of ferroelectricand relaxor properties in epitaxial films of Ba1−xSrxTiO3,Phys. Rev. B 70, 132105 (2004).

[23] C. Hubert, J. Levy, E. J. Cukauskas, and S. W. Kirchoefer,Mesoscopic Microwave Dispersion in Ferroelectric ThinFilms, Phys. Rev. Lett. 85, 1998 (2000).

[24] Z. G. Lu and G. Calvarin, Frequency dependence of thecomplex dielectric permittivity of ferroelectric relaxors,Phys. Rev. B 51, 2694 (1995).

[25] M. Tyunina and M. Savinov, Dynamic nonlinearity inepitaxial BaTiO3 films, Phys. Rev. B 94, 054109 (2016).

[26] S. Hashemizadeh and D. Damjanovic, Nonlinear dynam-ics of polar regions in paraelectric phase of Ba1−xSrxTiO3 ceramics, Appl. Phys. Lett. 110, 192905 (2017).

[27] C.-H. Lee et al., Exploiting dimensionality and defect miti-gation to create tunable microwave dielectrics, Nature 502,532 (2013).

[28] P. Shafer, P. García-Fernández, P. Aguado-Puente, A. R.Damodaran, A. K. Yadav, C. T. Nelson, S.-L. Hsu, J. C.Wojdeł, J. Íñiguez, L. W. Martin, E. Arenholz, J. Junquera,and R. Ramesh, Emergent chirality in the electric polariza-tion texture of titanate superlattices, Proc. Natl. Acad. Sci.115, 915 (2018).

[29] Q. Li, C. T. Nelson, S.-L. Hsu, A. R. Damodaran, L.-L.Li, A. K. Yadav, M. McCarter, L. W. Martin, R. Ramesh,and S. V. Kalinin, Quantification of flexoelectricity inPbTiO3/SrTiO3 superlattice polar vortices using machinelearning and phase-field modeling, Nat. Commun. 8, 1468(2017).

[30] J. Mangeri, Y. Espinal, A. Jokisaari, S. P. Alpay,S. Nakhmanson, and O. Heinonen, Topological phasetransformations and intrinsic size effects in ferroelectricnanoparticles, Nanoscale 9, 1616 (2017).

[31] D. Karpov, Z. Liu, T. D. S. Rolo, R. Harder, P. V. Bal-achandran, D. Xue, T. Lookman, and E. Fohtung, Three-dimensional imaging of vortex structure in a ferroelectricnanoparticle driven by an electric field, Nat. Commun. 8,280 (2017).

[32] A. K. Tagantsev, I. Stolichnov, N. Setter, J. S. Cross, andM. Tsukada, Non-Kolmogorov-Avrami switching kinet-ics in ferroelectric thin films, Phys. Rev. B 66, 214109(2002).

[33] Y. So, D. Kim, T. Noh, J.-G. Yoon, and T. Song, Polar-ization switching kinetics of epitaxial Pb(Zr0.4Ti0.6)O3 thinfilms, Appl. Phys. Lett. 86, 092905 (2005).

[34] J. Li, B. Nagaraj, H. Liang, W. Cao, C. H. Lee, andR. Ramesh, Ultrafast polarization switching in thin-filmferroelectrics, Appl. Phys. Lett. 84, 1174 (2004).

[35] J. Li, H. Liang, B. Nagaraj, W. Cao, C. H. Lee, and R.Ramesh, Application of an ultrafast photonic technique

044026-12

Page 13: Measurements of Nonlinear Polarization Dynamics in the

MEASUREMENTS OF NONLINEAR POLARIZATION... PHYS. REV. APPLIED 13, 044026 (2020)

to study polarization switching dynamics of thin-filmferroelectric capacitors, J. Lightwave Technol. 21, 3282(2003).

[36] D. F. Williams, T. S. Clement, P. D. Hale, and A. Dienstfrey,in 2006 68th ARFTG Conference: Microwave Measure-ment (IEEE, Broomfield, CO, USA, 2006), p. 1.

[37] A. E. Glazounov and A. K. Tagantsev, Phenomenolog-ical Model of Dynamic Nonlinear Response of RelaxorFerroelectrics, Phys. Rev. Lett. 85, 2192 (2000).

[38] R. W. Boyd, Nonlinear Optics (Academic press, San Diego,CA, USA, 2003).

[39] J. Mateu, C. Collado, N. Orloff, J. C. Booth, E.Rocas, A. Padilla, and J. M. O’Callaghan, Third-orderintermodulation distortion and harmonic generation in mis-matched weakly nonlinear transmission lines, IEEE Trans.Microw. Theory Tech. 57, 10 (2009).

[40] J. Mateu, J. C. Booth, and S. A. Schima, Frequency tuningand spurious signal generation at microwave frequenciesin ferroelectric SrTiO3 thin-film transmission lines, IEEETrans. Microw. Theory Tech. 55, 391 (2007).

[41] J. C. Booth, R. H. Ono, I. Takeuchi, and K.-S. Chang,Microwave frequency tuning and harmonic generation inferroelectric thin film transmission lines, Appl. Phys. Lett.81, 718 (2002).

[42] The model is given by Eq. (25). We express the electricfield and the displacement field in a normalized, dimension-less form with Enorm = E/ENL and Dnorm = αD/ENL. Tomodel the dynamics, we set f = [1 + iωτ0]−1 with τ0 = 1.The first- and second-order responses, d(1) and d(2), arecalculated from Eqs. (26) and (27).

[43] D. Niermann, C. P. Grams, M. Schalenbach, P. Becker, L.Bohaty, J. Stein, M. Braden, and J. Hemberger, Domaindynamics in the multiferroic phase of MnWO4, Phys. Rev.B 89, 134412 (2014).

[44] F. Bergmann, M. Letz, H. Maune, and G. Jakob, Highsensitivity characterization of the nonlinear electric suscep-tibility of a glass ceramic in the microwave range, Appl.Phys. Lett. 114, 212903 (2019).

[45] H. Tamura, J. Hattori, T. Nishikawa, and K. Wakino, Thirdharmonic distortion of dielectric resonator materials, Jpn. J.Appl. Phys. 28, 2528 (1989).

[46] J. J. Bussgang, L. Ehrman, and J. W. Graham, Analysisof nonlinear systems with multiple inputs, Proc. IEEE 62,1088 (1974).

[47] R. B. Marks and D. F. Williams, A general waveguidecircuit theory, J. Res. Natl. Inst. Stand. Technol. 97, 533(1992).

[48] V. Radisic, D. R. Hjelme, A. Horrigan, Z. B. Popovic, andA. R. Mickelson, Experimentally verifiable modeling ofcoplanar waveguide discontinuities, IEEE Trans. Microw.Theory Tech. 41, 1524 (1993).

[49] A. M. Hagerstrom, E. Marksz, C. J. Long, J. C. Booth,and N. D. Orloff, in 90th ARFTG Microwave MeasurementSymposium (ARFTG) (IEEE, Boulder, CO, USA, 2017),p. 1.

[50] I. Stakgold and M. J. Holst, Green’s Functions and Bound-ary Value Problems (John Wiley & Sons, Hoboken, NJ,USA, 2011).

[51] R. B. Marks, A multiline method of network analyzercalibration, IEEE Trans. Microw. Theory Tech. 39, 1205(1991).

[52] D. E. Root, J. Verspecht, J. Horn, and M. Marcu,X-Parameters: Characterization, Modeling, and Design ofNonlinear RF and Microwave Components (CambridgeUniversity Press, Cambridge, UK, 2013).

[53] J. A. Jargon, C. Cho, D. F. Williams, and P. D. Hale, in2015 85th Microwave Measurement Conference (ARFTG)(IEEE, Phoenix, AZ, 2015), p. 1.

[54] D. F. Williams, J. C. Wang, and U. Arz, An opti-mal vector-network-analyzer calibration algorithm, IEEETrans. Microw. Theory Tech. 51, 2391 (2003).

[55] N. D. Orloff, J. Mateu, A. Lewandowski, E. Rocas, J. King,D. Gu, X. Lu, C. Collado, I. Takeuchi, and J. C. Booth,A compact variable-temperature broadband series-resistorcalibration, IEEE Trans. Microw. Theory Tech. 59, 188(2011).

[56] D. F. Williams, U. Arz, and H. Grabinski, in 1998 IEEEMTT-S International Microwave Symposium Digest (Cat.No. 98CH36192) (IEEE, Baltimore, MD, USA, 1998), Vol.3, p. 1917.

[57] W. T. Weeks, L. L. H. Wu, M. F. McAllister, and A.Singh, Resistive and inductive skin effect in rectangularconductors, IBM J. Res. Dev. 23, 652 (1979).

[58] C. A. E. Little, N. D. Orloff, I. E. Hanemann, C. J. Long,V. M. Bright, and J. C. Booth, Modeling electrical double-layer effects for microfluidic impedance spectroscopy from100 kHz to 110 GHz, Lab Chip 17, 2674 (2017).

[59] E. Carlsson and S. Gevorgian, Conformal mapping ofthe field and charge distributions in multilayered sub-strate CPWs, IEEE Trans. Microw. Theory Tech. 47, 1544(1999).

[60] H. Orihara, S. Hashimoto, and Y. Ishibashi, A theory of dehysteresis loop based on the avrami model, J. Phys. Soc.Jpn. 63, 1031 (1994).

[61] Y. Ishibashi and H. Orihara, A theory of de hysteresis loop,Integr. Ferroelectr. 9, 57 (1995).

[62] R. Pirc and R. Blinc, Spherical random-bond–random-fieldmodel of relaxor ferroelectrics, Phys. Rev. B 60, 13470(1999).

[63] D. R. Chase, L.-Y. Chen, and R. A. York, Modeling thecapacitive nonlinearity in thin-film BST varactors, IEEETrans. Microw. Theory Tech. 53, 3215 (2005).

[64] A. Hagerstrom, E. Marksz, C. Long, J. Booth, I.Takeuchi, and N. Orloff, in 2018 IEEE MTT-S InternationalMicrowave Workshop Series on Advanced Materials andProcesses for RF and THz Applications (IMWS-AMP 2018)(IEEE, Ann Arbor, MI, USA, 2018).

[65] Note that the uncertainties in Fig. 8(c) are smaller than theuncertainties in the right panels of Figs. 8(a) and 8(b). Inthe narrow-band, voltage-swept measurement, we have arelatively large uncertainty in the phase of our incident sig-nals because the filters we are using attenuated the signalsfrom our phase standard. The data in Figs. 8(a) and 8(b) arecalculated with a regression that considered amplitude andphase, while the data in Fig. 8(c) are calculated with a dif-ferent regression that only considered the amplitude of themixing product.

[66] A. M. Hagerstrom, E. J. Marksz, X. Zhang, X. Lu, C. J.Long, J. C. Booth, I. Takeuchi, and N. D. Orloff, Datafor: “Measurements of nonlinear polarization dynamicsin the tens of gigahertz,” https://doi.org/10.18434/M31963(2020).

044026-13