Reconstructing phase diagrams from local measurements via gaussian processes: mapping the temperature-composition space to confidence

Nature

Reconstructing phase diagrams from local measurements via gaussian processes: mapping the temperature-composition space to confidence"


Play all audios:

Loading...

ABSTRACT We show the ability to map the phase diagram of a relaxor-ferroelectric system as a function of temperature and composition through local hysteresis curve acquisition, with the


voltage spectroscopy data being used as a proxy for the (unknown) microscopic state or thermodynamic parameters of materials. Given the discrete nature of the measurement points, we use


Gaussian processes to reconstruct hysteresis loops in temperature and voltage space, and compare the results with the raw data and bulk dielectric spectroscopy measurements. The results


indicate that the surface transition temperature is similar for all but one composition with respect to the bulk. Through clustering algorithms, we recreate the main features of the bulk


diagram, and provide statistical confidence estimates for the reconstructed phase transition temperatures. We validate the method by using Gaussian processes to predict hysteresis loops for


a given temperature for a composition unseen by the algorithm, and compare with measurements. These techniques can be used to map phase diagrams from functional materials in an automated


fashion, and provide a method for uncertainty quantification and model selection. SIMILAR CONTENT BEING VIEWED BY OTHERS REVEALING THE HIDDEN STRUCTURE OF DISORDERED MATERIALS BY


PARAMETERIZING THEIR LOCAL STRUCTURAL MANIFOLD Article Open access 24 May 2024 THE CLUSTER DECOMPOSITION OF THE CONFIGURATIONAL ENERGY OF MULTICOMPONENT ALLOYS Article Open access 19 July


2024 A FRAMEWORK FOR QUANTIFYING UNCERTAINTY IN DFT ENERGY CORRECTIONS Article Open access 29 July 2021 INTRODUCTION Phase diagrams serve as descriptors of material behavior and material


properties, and their forms are intrinsically linked to the underpinning physics driving the system. Traditionally, phase diagrams are determined from property measurements across material


compositions, when the anomalies in specific heat, lattice parameter, polarization, or other macroscopic parameters are interpreted as signatures of phase transitions, allowing the phase


boundaries to be drawn.1 Recently, progressively large attention has been focused on extracting this information from theoretical or combinatorial data.2,3,4,5,6,7 The groups involved in


synthesis of materials via combinatorial methods have recently turned to using machine-learning-based approaches to characterize the structural phases present, applying clustering methods to


large volumes of X-ray diffraction data. These techniques have been explored by several groups worldwide in an effort to accelerate experimental materials design and


discovery.8,9,10,11,12,13 They rely on either unsupervised or semi-supervised learning, to group (cluster) diffraction patterns based on some given similarity metric, enabling the


distinguishing of phase boundaries across the composition spread.14 However, the same techniques are seldom applied to local measurements:15 for instance, while studies of piezoresponse


force microscopy (PFM) of piezo/ferroelectric samples as a function of temperature are routine, studies of functional response (such as hysteresis loops) as a function of both temperature


and composition are rare. Partly, the difficulty is related to the appropriate methods of analysis of large multidimensional data sets across temperature and voltage space in a consistent


fashion. Fundamentally, there is appreciable difficulty in accurately determining the phase transition temperature from local measurements, given that they (usually) do not provide access to


the order parameter directly and often are not quantitative. Among the scanning probe microcopy methods, piezoresponse force microscopy offers an example of technique which intrinsically


quantitative due to specific voltage-dependent contact mechanics.16,17,18,19 The PFM response measures the remnant effective piezocoefficient, which is proportional to polarization order


parameter in proper ferroelectrics. However, it can also arise from a number of other mechanisms that can greatly complicate the interpretation of the measured hysteresis loops.20


Furthermore, in relaxor ferroelectrics,21,22,23,24 the nature of order parameter is significantly more complex,25,26,27 as therefore is the nature of PFM signal formation. However, the


ability to reconstruct the phase diagram from local spectroscopic PFM measurements could be useful, because such experiments are often simpler to perform (machines and software are


commercially available), and can be carried out on extremely small sample sizes (<~20 nm). More importantly, these measurements provide access to the phase transition at the surface,


which can differ significantly from the bulk due to symmetry breaking, chemical segregation, adsorbates, etc.28 and thus enables the study of how the phase transition is affected by these


factors. As a secondary difficulty in constructing phase diagrams from experimental data, the uncertainty in these measurements is often hard to compute. Since measurement points in the


composition-temperature space are discrete, demarcating the phase boundary requires interpolation, typically using functional forms such as high-order polynomials, or piecewise fitting such


as cubic splines. However, these approaches have the disadvantage in that they impose a form of the function on the data, and cannot provide robust uncertainty quantification (the


uncertainty depends on the choice of fitting model). Gaussian process (GP) methods are a class of machine learning, and have been utilized to tackle this problem, and provide an elegant (if


computationally expensive) method that allows for regressions of data without the need to postulate an underlying functional form. The GP regression is non-parametric, naturally extends to


higher-dimensional spaces, and provides robust uncertainty quantification.29 In this letter, we show the use of these methods in exploring the phase diagram of a relaxor-ferroelectric


system, (1−_x_) Pb(Fe0.5Nb0.5)O3−_x_Ni0.65Zn0.35Fe2O4 (_x_ = 0, 0.10, 0.20, and 0.30). The four distinct compositions are investigated using band-excitation piezoresponse spectroscopy


(BEPS),30,31 yielding hysteresis loops as a function of temperature for all four samples below and above the Curie temperature. Basic clustering methods allow coarse determination of the


phase diagram, which is compared to the bulk phase transition temperatures as deduced from dielectric spectroscopy measurements. Gaussian process regression is then used to model the


response as a function of temperature (_T_) and voltage (_V_), providing uncertainty bounds in the temperature space, and allow for the prediction of the response at arbitrary points in the


2D _T_−_V_ space, resulting in a full reconstruction of the phase diagram. Clustering methods are carried out on the GP-predicted loops and compared to the results from the raw hysteresis


loops, indicating similar trends. The results point to a reduction in the surface phase transition temperature with respect to the bulk for the _x_ = 0.2 composition, with little difference


seen for all other compositions investigated. Further, we use the method to predict hysteresis loops for a composition unseen by the algorithm, and compare it with experimental results,


confirming the validity of the approach. The method shown here is general for determining phase diagrams from local response measurements, allows for uncertainty in the data to be accurately


determined, and represents an important step toward automated high-throughput characterization via local spectroscopic measurements in scanning probe microscopy. RESULTS The (1−_x_)


Pb(Fe0.5Nb0.5)O3−_x_Ni0.65Zn0.35Fe2O4 (_x_ = 0, 0.10, 0.20, and 0.30) films investigated in this study will be abbreviated as PFN, PN1, PN2, and PN3, respectively (see “Methods” for sample


preparation details). We abbreviate Pb(Fe0.5Nb0.5)O3 as PFN and Ni0.65Zn0.35Fe2O4 as NZFO. This system is a perovskite/spinel composite that shows promise in magnetoelectric applications.


The Pb(Fe0.5Nb0.5)O3−Ni0.65Zn0.35Fe2O4 composite structures have been chosen for the present study to observe enhanced ferroelectric and magnetic properties with strong magnetoelectric


coupling above room temperature. PFN is a well-known multiferroic material with a ferroelectric Curie temperature between 379 and 385 K having high dielectric constant, low loss tangent, and


high piezoelectric constant; it also shows magnetic ordering below RT (TN ~150–200 K).32,33,34,35 Recently, the existence of ferroelectric, ferromagnetic, and ferroelastic properties in PFN


with interesting ferroelectric and magnetic properties have been reported.36 On the other hand, Nickel zinc ferrites are soft magnetic materials with high saturation magnetization, low


coercivity, high resistivity, reasonable magnetostriction, low dielectric losses, high dielectric constant, and high magnetic Curie temperatures. Ni0.65Zn0.35Fe2O4 has been chosen for the


present work as this composition exhibits the combination of highest saturation magnetization and high magnetostriction in the entire Ni–Zn series with a magnetic Curie temperature of ~710 


K.37 As such, exploration of the phase diagram of this system in terms of ferroelectric properties is important in material optimization, and provides the basis for our work. To examine the


phase purity and crystalline quality of PFN, PN1, PN2, and PN3 thin films of 70 nm thickness grown on LaNiO3 (LNO)-buffered (LaAlO3)0.3 (Sr2AlTaO6)0.7 (LSAT) substrate, X-ray diffraction


(XRD) measurements were carried out at room temperature and the XRD patterns for PFN sample are shown in Fig. 1a. The _θ_−2_θ_ large-angle XRD patterns (20–80°) showed only the diffraction


peaks from the substrate and (00 _l_) pseudocubic reflections from PFN and NZFO confirms that these films are highly _c_-axis oriented in nature, and that the secondary phases are absent.


The XRD peaks with symbols *, #, and & correspond to the peaks of LSAT/LNO (overlapped), PFN, and NZFO, respectively. The superposition of both the crystallographic phases of PFN and


NZFO are observed in the XRD patterns, confirming the formation of the composites. The intensity of the peaks corresponding to NZFO phase continues increasing with the increase in ferrite


concentration in the composites. No intermediate phases, i.e., extra reflection peaks that would be indicative of secondary phases and peaks from lead-deficient pyrochlore phases were


detected, indicating the high purity and high quality of the thin films (see supplementary for XRD scans of all samples). The atomic force microscopy topography image for the PFN sample is


shown in Fig. 1b, along with the vertical PFM amplitude and phase image in Fig. 1c, d. The results indicate that the film is extremely flat, and monodomain in the as-grown state, but can be


switched. All composite thin films also show similar behaviors. To probe the ferroelectric phase transitions and to explore the effect of concentration of the magnetic material (NZFO) on the


nature of phase transition in composite thin films, the dielectric permittivity of PFN, PN1, PN2, and PN3 as a function of temperature have been measured in a wide range of frequencies and


is shown in Fig. 1e–h, respectively. It is observed that for all compositions the dielectric constant increases with increasing temperature, attains its maximum value at a certain


temperature and then deceases. The temperature at which maximum value of dielectric constant is observed corresponds to the ferroelectric to paraelectric phase transition temperature (_T_C).


Single crystal and ceramic samples of PFN exhibit diffuse type of ferroelectric phase transition having _T_c ranging from 379 to 385 K.38 From the dielectric spectra, the values of _T_c


were determined (see “Methods”) for PFN, PN1, PN2, and PN3, which were found to be 400, 430, 440, and 460 K, respectively. The significantly larger values of the transition temperature as


compared to the bulk may be attributed to the compressive epitaxial strain,39 which is expected to be on the order of ~4.6%. Additionally, the transition temperature increases with increase


of the magnetic material (NZFO) concentration. This behavior, i.e., the increase of ferroelectric _T_c with increase of magnetic content has been theoretically predicted for 1–3 type


multiferroic composite thin films.40 All the measured thin films exhibit a diffuse (broadening) type of phase transition behavior. The spectra can be fit to a modified Curie Weiss law to


describe the diffuseness of the phase transition41 and is given by the relation $$\frac{1}{\varepsilon } - \frac{1}{{\varepsilon _{\rm{m}}}} = \frac{{(T - T_{\rm{m}})^\gamma }}{{C_1}}$$ (1)


where _γ_ corresponds to the degree of relaxation and its value varies from 1 < _γ_ < 2. For the limiting value of _γ_ = 1, the above equation reduced to the Cure Weiss law which is


applicable to normal ferroelectrics. But in the case of _γ_ = 2, the quadratic dependence is applicable to relaxor-ferroelectric materials.41,42 It should be noted that when _γ_ varies from


1 to 2, this parameter is a measure of the degree of relaxation. Simple models that assume microscopic compositional fluctuations, where the micro-regions have similar dielectric properties


but different Curie temperatures, can reproduce the _γ_ = 2 case, but in real situations this is actually more complex. From fitting of the dielectric spectra, the value of _γ_ for PFN, PN1,


PN2, and PN3 are found to be 1.98 ± 0.05, 1.95 ± 0.01, 1.80 ± 0.01, and 1.97 ± 0.02, respectively (reported uncertainties are 1 s.d.). Thus, the PN2 sample appears to stand out from the


other three samples in that the diffuseness of the phase transition is actually lowered. Elaboration of the reasons behind this difference is provided in the “Discussion” section. Having


established the main features of the bulk phase diagram of the system, we next turned to local measurements of the hysteresis loops using BEPS. Briefly, this involves applying a DC train of


pulses comprising a triangular amplitude envelope, with BE packets applied after each DC pulse to determine the strain state (piezoresponse) of the system. Here, we applied BE packets both


during (on-field) and after (off-field) the DC pulse application, but from hereon we will focus on the off-field piezoresponse only. Measurements were performed on a 5 × 5 spectroscopic grid


across a [1 μm × 1 μm] region of each sample, at different temperatures. At each point in the spectroscopic grid, two hysteresis loops were acquired, resulting in 50 hysteresis loops at a


given temperature for each sample investigated. The average of the 50 hysteresis loops, plotted as a function of temperature for each of the four compositions studied, are shown in Fig. 2.


The data indicate that all four samples show a phase transition, although the collapse of the hysteresis loops (and onset of the paraelectric phase) is more rapid with temperature for the


PFN than for the other three compositions. CLUSTERING To map the phase diagram based on the shape of the hysteresis loops, we turned to the _k_-means clustering algorithm. This is a common


method used to cluster data in high dimensional spaces, and aims to segment data into _k_ sets (clusters), such that the within-cluster sum of squares is minimized.43 We performed the


_k_-means clustering method (as implemented in the open source scikit-learn python package44) on the BEPS loop data, with _k_ _=_ 3 (clusters). It should be noted that each composition was


investigated at similar (but slightly different) temperatures, but the data were grouped together before clustering (this is explored later). The results are plotted in Fig. 3a, b, with the


labels in Fig. 3a and the cluster centers plotted in Fig. 3b. Given that there are 50 hysteresis loops at each temperature and composition, the _k_-means clustering algorithm assigns each


hysteresis loop a label as belonging to one of the three clusters. In the RGB labels diagram, the color of each pixel in the phase diagram is indicative of the weights, i.e., entirely red


pixels signify that all 50 loops are classified as belonging to cluster 0, entirely green are classified as cluster 1, and entirely blue are cluster 2. The average response within each


cluster (“Cluster Centers”, Fig. 3b) indicates that cluster 1 is effectively the paraelectric phase, while the red cluster signifies the ferroelectric phase. Meanwhile, a cluster with much


larger response is present for the PFN, due to the superior piezoelectric characteristics of this composition and is colored in blue. Remarkably, the simple clustering algorithm is capable


of automatically demarcating the phase diagram from the given data set. GAUSSIAN PROCESSES MODELING As noted above, each sample was investigated at slightly different temperatures, which is


visualized in Fig. 4a. To adequately compare each measurement across samples, one needs to resample the measured hysteresis loops in the temperature–voltage space. This resampling can be


performed using a number of methods, ranging from functional fits to multidimensional interpolation techniques. To surmount difficulties associated with choosing the appropriate form of a


functional fit or the unknown statistical confidence of interpolated data, we utilized Gaussian processes modeling (GP). The latter is a multidimensional and non-parametric Bayesian modeling


technique used in various statistical learning tasks such as regression and classification. In regression, one aims to find an optimized function _f_ that maps the inputs _X_ (i.e., vectors


of temperature and DC bias) to output _Y_ (hysteresis loops), given some observations _Y_ of the latter at some temperature and bias data points. Briefly, in Gaussian processes modeling,


the input vectors are projected into a higher-dimensional space via a mapping \(\phi _1 = \exp \left( { - \frac{1}{{2l}}^2\left| {x - x\prime } \right|^2} \right)\). This allows to model


both linear as well as nonlinear functions, _f_ and is now given by $${\it{f}}\left({\boldsymbol{x}} \right){\mathrm{ = }}\phi


{\mathrm{(}}{\boldsymbol{x}}^{\it{T}}{\mathrm{)}}{\boldsymbol{w}}{\mathrm{,}}\,{\mathrm{and}}\,{\boldsymbol{w}}{\mathrm{\sim


}}{\boldsymbol{N}}{\mathrm{(}}{\boldsymbol{m}}{\mathrm{,}}\,{\it{\Sigma }}{\mathrm{)}},$$ (2) where _W_ are (unknown) parameters that define _f_. The parameters of the model are assumed to


follow a (prior) multivariate normal distribution _N_ with mean values _m_, and covariance matrix \(\Sigma\). The covariance matrix defined by \(\Sigma = \sigma


^2K({\boldsymbol{x}},{\boldsymbol{x}}\prime )\) is fully determined once the kernel _K_ is chosen, while the variance _σ_ is estimated by maximum likelihood using the observations _Y_


(incidentally, the choice of _K_ also fully specifies the mapping \(\phi\)). The most popular kernel is the Gaussian correlation function or radial basis function (RBF), given by $$K\left(


{{\boldsymbol{x}},{\boldsymbol{x}}\prime } \right) = {\mathrm{exp}}\left( { - \frac{1}{{2\ell ^2}}\left| {{\boldsymbol{x}} - {\boldsymbol{x}}\prime } \right|^2} \right),$$ (3) where \(\ell\)


is a hyper-parameter of the GP model. To take into account the noise inherent in each measurement, we used a RBF kernel with a white kernel, i.e., $$K = \alpha \phi _1 + \beta \phi _2$$


where \(\phi _1 = \exp \left( { - \frac{1}{{2l}}^2\left| {x - x\prime } \right|^2} \right)\) is the RBF Kernel in (3), and _ϕ_2 is a noise kernel given by $$\phi _2 = \left\{


{\begin{array}{*{20}{l}} {{\it{\epsilon }},x = x\prime } \hfill \\ {0,else} \hfill \end{array}} \right.$$ (4) where \({\it{\epsilon }}\) is a noise level. We choose \(\alpha\) = 0.1 and _β_ 


= 0.5, with bounds on the noise level \({\it{\epsilon }}\) ([1 × 10−10, 1 × 10−2]). Once an optimal value of the hyper-parameter is found (by maximizing the log-likelihood of the model), one


arrives at an optimal function _f_, which can be used to predict hysteresis loops at new temperature and bias points. The predictions of hysteresis loops using GP regression (and the RBF


kernel) along the temperature axis for the PN2 case is seen in Fig. 4b, for different individual voltages, along with a confidence interval (2 s.d.) filled in blue. This uncertainty


quantification is an advantageous feature of GP regression, and can be useful for applying or developing models to describe the nature of the phase transition. The GP regression is also


fully generalizable across higher dimensions, although the computational cost is higher given that it is not a sparse method. It would appear that the uncertainties represented in Fig. 4b


are rather large, and are a result of the sparsity of data points. In a sense, this is the price to be paid for unknown functional forms; while other methods (such as polynomial or spline


interpolation) could work, the uncertainty estimates would not be accurate (given that this uncertainty depends on the model choice, which is largely arbitrary, especially for ferroelectric


hysteresis loops45). However, the ability to generalize and produce predictions with quantified uncertainty is of critical importance, and not available through the other standard means.


Shown in Fig. 5 are the GP regressions on the 2D surfaces of the average (i.e., spatial average) piezoresponse as a function of voltage and temperature for all four compositions. In Fig.


5a–d, the raw data and the GP-predicted loops as a function of temperature are plotted for each composition; in general, the GP-predicted loops closely align with the raw data for each


composition. The full GP-predicted 2D surface plots are shown in Fig. 6a–d for the four compositions, and can be regenerated for arbitrary values using the model. The distinct behaviors and


closure of the hysteresis loops with increasing temperature are readily observed in this representation. Given the wrapped nature of the 3D plots, we also provide the surface as a function


of voltage step, effectively unfolding the hysteresis loops along the voltage axis (see supplementary material). With the GP-predicted data sets for each composition, we were then able to


perform _K_-means on the full predicted data set containing the loops at different compositions and temperatures. In other words, we took the data in Fig. 6 from the Gaussian process


prediction and concatenated them into a single 2D matrix, and subsequently performed _k_-means clustering. The cluster centers and the label map is shown in Fig. 7a, b, and again clearly


demarcates the line between the paraelectric and ferroelectric phases. Note that here we performed this for _k_ = 4 clusters (as opposed to 3 earlier), because we found it more accurately


depicted the changed loop behavior of the PN2 sample (nonetheless, trends are similar for _k_ = 3 clusters, shown in supplementary, with the main difference being slightly lowered _T_c


values, though with a more drastic reduction for the PN2 case). Based on the label matrix, we can determine the phase transition temperatures for the surface, and compare it with the phase


transition temperature of the bulk, as shown in Fig. 7c. The surface phase transition temperatures are marked as crosses, while the bulk phase diagram is plotted as colored circles (the


latter are taken from the dielectric spectroscopy data in Fig. 1). As with Figs. 3a and 7a, the most interesting feature of this phase diagram is the behavior of PN2, which appears markedly


different from the other compositions. USING GP FOR PREDICTION IN COMPOSITION SPACE Perhaps of most importance is the ability to predict functionality as a function of


composition,9,11,46,47,48 given that this can potentially reduce the time spent on exploring the composition space (which is generally more time consuming than performing measurements at


different conditions, for instance). We attempt to do this on our existing system, for the PN2 composition, given hysteresis loops at a single temperature for the PFN, PN1, and PN3


compositions. The first step is to perform the GP in the composition space given the input data (three hysteresis loops, one for each composition). The results of the GP are shown in Fig.


8a, with the mean plotted as a black line and the variance estimates as filled blue regions. We can then use the trained model to predict the hysteresis loop for the missing PN2 composition,


which is shown in Fig. 8b. Note that the response is plotted against the voltage step (i.e., time), not voltage, to more easily visualize the upper and lower confidence bounds (see


supplementary for plot as a function of voltage). The actual measurements of the PN2 composition are shown as filled circles with associated standard deviation (arising from spatial


variation in the measured hysteresis loops). In all cases, the actual measured values are within 1 s.d. of the prediction bound, suggesting good agreement and validation of the GP model.


DISCUSSION The phase transition temperatures as measured by PFM are slightly different from the bulk, especially for the PN2 sample. This is likely due to a combination of the fact that PFM


is not measuring the polarization directly (piezoresponse is strain), whereas macroscopic dielectric spectra measure the film polarization as well as interfacial polarization; furthermore,


PFM is inherently surface sensitive, with associated issues relating to change of the surface crystal structure, chemical segregation, or symmetry breaking, which can all affect the nature


and temperature of the transition in relaxor-ferroelectric systems.28 Indeed, we found that the PN2 transition appears more diffuse in the local measurements, as evidenced by the more


gradual transition from green to red in Fig. 3a, whereas it shows a less diffuse transition than the other compositions in the dielectric spectroscopy. There can be several reasons for the


discrepancy in the phase transition temperature of the PN2 sample with respect to the others; in addition to the basic substrate strain-related effects, given this is a composite, one can


expect effects from the strain impact of one phase on another, as well as possible interdiffusion of cations across the interfaces in this system leading to more spatial variability. It is


also possible that there is an interplay between the ferroelectric and magnetic order parameters of the system that can affect the _T_c in a nonlinear fashion in the composition space,


especially as the parent compound PFN is known to exhibit weak ferromagnetism up to 400 K.49 Additional surface-sensitive local PFM investigations can target the influence of surface-based


modifications of the sample, e.g., local tip pressure, to uncover the main driving mechanisms behind the locally diffuse phase transition in PN2. The results confirm the utility of the


clustering approach in demarcating the important features of the phase diagram, as well as illustrate the differences between the surface and bulk measurements in this system. An interesting


future direction in the application of Gaussian processes in the reconstruction of phase diagrams from local measurements is the prediction of the onset of diffuse transitions in


relaxor-ferroelectric systems as a function of composition. Traditionally, an accurate determination of the dependence of piezoelectric and dielectric properties in relaxors on the


dependence of ionic substitution requires the synthesis of a large number of materials. As we demonstrated, however, Gaussian processes machine-learning techniques can accurately reconstruct


phase diagrams in the presence of a finite and small subset of temperature and voltage measurements. Therefore, one may explore the feasibility of accurately predicting the ionic


substitution ratio at which relaxor-ferroelectric behavior takes place from a small and finite number of material compositions via an application of GP machine learning. The local and global


ferroelectric phase transitions of a relaxor-ferroelectric system were probed using dielectric and band-excitation PFM measurements. Through acquisition of hysteresis loops as a function of


temperature for four distinct compositions, we compiled a large data set of the functional response of the system. Basic unsupervised clustering algorithms applied to the data set


facilitated the reconstruction of the temperature-composition phase diagram. Given the discrete and non-uniform nature of the measured parameter space, as well as inherent nonlinear


dependence of the hysteresis loops on thermodynamic potentials, we employed Gaussian process regression to reconstruct the data in both the 1D case (temperature) and the 2D case (temperature


and voltage), accompanied with confidence quantification without imposing functional forms. Clustering was performed on the GP-reconstructed data set to reproduce the phase diagram, with


results compared to bulk dielectric spectroscopy and allowing comparison between surface and bulk transition temperatures. These techniques can be used for automated determination of phase


diagrams with uncertainty quantification, across a wide range of imaging modalities. METHODS PULSED LASER DEPOSITION Phase pure ceramic targets of Pb(Fe0.5Nb0.5)O3 and


(1−_x_)Pb(Fe0.5Nb0.5)O3−_x_Ni0.65Zn0.35Fe2O4 (_x_ = 0.10, 0.20, and 0.30) composites were synthesized by high-temperature solid-state reaction technique. The detailed synthesis conditions


and structural properties of the above-mentioned ceramic targets are reported elsewhere.21 PFN, PN1, PN2, and PN3 thin films were grown on LaNiO3 (LNO) bottom electrode deposited on


commercially available (LaAlO3)0.3 (Sr2AlTaO6)0.7 (LSAT) (100) substrates using optimized pulsed laser deposition (PLD) with an excimer laser (KrF, 248 nm). Initially, LNO bottom electrode


of 50 nm was deposited on LSAT substrate at 700 °C in an oxygen pressure of 200 mTorr, followed by annealing in same oxygen ambient of 300 Torr for 30 min at same temperature and then slowly


cooled down to room temperature. PFN, PN1, PN2, and PN3 thin films were deposited at a fixed temperature of 600 °C under oxygen ambient of 20 mTorr. The laser pulse repetition rate was 2 Hz


with a laser energy density ~1.5 J/cm2. Later, the PFN and other composite thin films were annealed in an oxygen atmosphere of 300 Torr for 30 min at 700 °C and later cooled down to room


temperature slowly. The thickness of all above-mentioned thin films were kept constant ~70 nm for the comparison of physical functionalities. X-RAY DIFFRACTION The phase purity and


crystallinity of PFN, PN1, PN2, and PN3 thin films were examined using high-resolution X-ray diffraction (HRXRD) using CuKα radiation with wavelength of _λ_ = 1.5405 Å operated at a scan


rate of 1°/min over the angular range (2_θ_) of 20–80 at room temperature. The thicknesses of all the thin films were measured utilizing XP-200 profilometer along with Filmetrics F-20.


DIELECTRIC CHARACTERIZATION Pt top electrodes of area ~10−4 cm2 and thickness ~40 nm were grown by dc sputtering utilizing a metal shadow square mask for electrical characterization.


Temperature-dependent dielectric parameters, i.e., capacitance, loss tangent, phase angles, and impedance were measured from 100 to 550 K in a wide frequency range using an impedance


analyzer HP4294A with MMR Technologies K-20 programmable temperature controller with fixed ac voltage amplitude of 0.1 V. To determine the ferroelectric-paraelectric phase transition


temperature, we plotted the the derivative of inverse of dielectric permittivity (1/_ε_) against temperature for the frequency 10 kHz. When the derivative tends to zero, it is known as the


phase transition temperature, and the phase transition temperature of PFN, PN1, PN2, and PN3 were found to be 400, 430, 440, and 460 K, respectively. BAND-EXCITATION PIEZORESPONSE


SPECTROSCOPY The PFM spectroscopy measurements were carried out by moving the tip across a pre-defined coordinate (_x_,_y_) grid on the surface of the samples and perturbing it with a DC


bias waveform, and then measuring the strain (piezoresponse) as a function of frequency via band-excitation (BE) technique. BE technique includes the generation of a band of AC frequencies


near the contact resonance of the cantilever and measure the electromechanical response over time with subsequent Fourier transformation back to the frequency domain to generate the


frequency-dependent response. This measurement is repeated for many time steps after every DC bias pulse; thus, at each (_x_,_y_) position, local spectroscopic information is obtained as a


function of perturbation voltage (_V_) and excitation frequency (_f_), characterizing the system’s mechanical response (vertical deflection) _R_ = _R_(_x_, _y_, _V_). The response _R_ at


each (_x_, _y_, _V_) step is fitted to a simple harmonic oscillator (SHO) model yielding the amplitude (_A_), phase (_φ_), quality factor (_Q_), and resonant frequency (_ω_) corresponding


with the response. The thin film was glued onto a sample plate with silver epoxy. The PFM experiments were carried out using moderately stiff Budget Sensors ElectriMulti75-G cantilevers (_k_


 ∼ 1 N/m) and a free resonance frequency (in air) of ∼75 kHz, on a multimode (Veeco) AFM equipped with a Nanonis controller. National Instruments DAQ cards were utilized for signal


generation and acquisition for the band-excitation measurements, which were performed using scripts written in Labview v11 and python 3. All analyses were also performed utilizing python 3,


including scikit-learn44 for Gaussian Process implementation and Pycroscopy (https://pycroscopy.github.io/pycroscopy/) for BEPS data analysis. Temperature-dependent PFM spectroscopic


measurements were carried out in a wide temperature range of 300–600 K with a high precision temperature controller. In order to avoid the moisture effect, all the thin films were heated up


to 50 °C for 30 min prior to the measurements. DATA AVAILABILITY The data and Jupyter notebook that details the analysis are included in the supplementary files accompanying this manuscript.


DISCLAIMER This manuscript has been authored by UT-Battelle, LLC, under contract no. DE-AC0500OR22725 with the U.S. Department of Energy. The United States Government retains and the


publisher, by accepting the article for publication, acknowledges that the United States Government retains a non-exclusive, paid-up, irrevocable, worldwide license to publish or reproduce


the published form of this manuscript, or allow others to do so, for the United States Government purposes. The Department of Energy will provide public access to these results of federally


sponsored research in accordance with the DOE Public Access Plan (http://energy.gov/downloads/doe-public-access-plan). REFERENCES * Hillert, M. _Phase Equilibria, Phase Diagrams and Phase


Transformations: Their Thermodynamic Basis_ (Cambridge University Press, Cambridge, 2007). * Diéguez, O. et al. Ab initio study of the phase diagram of epitaxial BaTiO3. _Phys. Rev. B_ 69,


212101 (2004). Article  Google Scholar  * Tavares, J., da Gama, M. T., Teixeira, P., Weis, J. & Nijmeijer, M. Phase diagram and critical behavior of the ferromagnetic Heisenberg fluid


from density-functional theory. _Phys. Rev. E_ 52, 1915 (1995). Article  Google Scholar  * Li, L., Muckerman, J. T., Hybertsen, M. S. & Allen, P. B. Phase diagram, structure, and


electronic properties of (Ga 1− x Zn x)(N 1− x O x) solid solutions from DFT-based simulations. _Phys. Rev. B_ 83, 134202 (2011). Article  Google Scholar  * Zhao, J.-C. Combinatorial


approaches as effective tools in the study of phase diagrams and composition–structure–property relationships. _Prog. Mater. Sci._ 51, 557–631 (2006). Article  Google Scholar  * Daniels, J.


E., Jo, W., Rödel, J., Honkimäki, V. & Jones, J. L. Electric-field-induced phase-change behavior in (Bi 0.5 Na 0.5) TiO 3–BaTiO 3–(K 0.5 Na 0.5) NbO 3: a combinatorial investigation.


_Acta Mater._ 58, 2103–2111 (2010). Article  Google Scholar  * Potyrailo, R. et al. Combinatorial and high-throughput screening of materials libraries: review of state of the art. _ACS Comb.


Sci._ 13, 579–633 (2011). Article  Google Scholar  * Bunn, J. K. et al. Generalized machine learning technique for automatic phase attribution in time variant high-throughput experimental


studies. _J. Mater. Res._ 30, 879–889 (2015). Article  Google Scholar  * Hattrick-Simpers, J. R., Gregoire, J. M. & Kusne, A. G. Perspective: composition–structure–property mapping in


high-throughput experiments: turning data into knowledge. _APL Mater._ 4, 053211 (2016). Article  Google Scholar  * Stein, H. S., Jiao, S. & Ludwig, A. Expediting combinatorial data set


analysis by combining human and algorithmic analysis. _ACS Comb. Sci._ 19, 1–8 (2017). Article  Google Scholar  * Kusne, A. G. et al. On-the-fly machine-learning for high-throughput


experiments: search for rare-earth-free permanent magnets. _Sci. Rep._ 4, 6367 (2014). Article  Google Scholar  * LeBras, R. et al. in _The 2nd International Workshop on Constraint Reasoning


and Optimization for Computational Sustainability_ (CROCS at CPAIOR-10, Bologna, Italy, 2010). * Green, M. et al. Fulfilling the promise of the materials genome initiative with


high-throughput experimental methodologies. _Appl. Phys. Rev._ 4, 011105 (2017). Article  Google Scholar  * Iwasaki, Y., Kusne, A. G. & Takeuchi, I. Comparison of dissimilarity measures


for cluster analysis of X-ray diffraction data from combinatorial libraries. _NPJ Comput. Mater._ 3, 4 (2017). Article  Google Scholar  * Laanait, N., Zhang, Z. & Schlepütz, C. M.


Imaging nanoscale lattice variations by machine learning of x-ray diffraction microscopy data. _Nanotechnology_ 27, 374002 (2016). Article  Google Scholar  * Karapetian, E., Kachanov, M.


& Kalinin, S. V. Nanoelectromechanics of piezoelectric indentation and applications to scanning probe microscopies of ferroelectric materials. _Philos. Mag._ 85, 1017–1051 (2005).


Article  Google Scholar  * Shin, J. et al. Simultaneous elastic and electromechanical imaging by scanning probe microscopy: Theory and applications to ferroelectric and biological materials.


_J. Vac. Sci. Technol. B_ 23, 2102–2108 (2005). Article  Google Scholar  * Kalinin, S. V., Karapetian, E. & Kachanov, M. Nanoelectromechanics of piezoresponse force microscopy. _Phys.


Rev. B_ 70, 184101 (2004). Article  Google Scholar  * Kalinin, S. V. & Gruverman, A. _Scanning Probe Microscopy: Electrical and Electromechanical Phenomena at the Nanoscale_ (Springer,


New York, 2006). * Vasudevan, R. K., Balke, N., Maksymovych, P., Jesse, S. & Kalinin, S. V. Ferroelectric or non-ferroelectric: Why so many materials exhibit “ferroelectricity” on the


nanoscale. _Appl. Phys. Rev._ 4, 021302 (2017). Article  Google Scholar  * Shvartsman, V. et al. Polar nanodomains and local ferroelectric phenomena in relaxor lead lanthanum zirconate


titanate ceramics. _Appl. Phys. Lett._ 86, 202907 (2005). Article  Google Scholar  * Rodriguez, B. J., Jesse, S., Bokov, A. A., Ye, Z.-G. & Kalinin, S. Mapping bias-induced phase


stability and random fields in relaxor ferroelectrics. _Appl. Phys. Lett._ 95, 092904 (2009). Article  Google Scholar  * Shvartsman, V. V., Kleemann, W., Łukasiewicz, T. & Dec, J.


Nanopolar structure in Sr x Ba 1− x Nb 2 O 6 single crystals tuned by Sr∕ Ba ratio and investigated by piezoelectric force microscopy. _Phys. Rev. B_ 77, 054105 (2008). Article  Google


Scholar  * Shvartsman, V. & Kholkin, A. Domain structure of 0.8 Pb (Mg 1/3 Nb 2/3) O 3− 0.2 PbTiO 3 studied by piezoresponse force microscopy. _Phys. Rev. B_ 69, 014102 (2004). Article 


Google Scholar  * Glinchuk, M. D. & Stephanovich, V. A. Dynamic properties of relaxor ferroelectrics. _J. Appl. Phys._ 85, 1722–1726 (1999). Article  Google Scholar  * Westphal, V.,


Kleemann, W. & Glinchuk, M. D. Diffuse phase-transitions and random-field-induced domain states of the relaxor ferroelectric PBMG1/3NB2/3O3. _Phys. Rev. Lett._ 68, 847–850 (1992).


Article  Google Scholar  * Vugmeister, B. E. & Rabitz, H. Coexistence of the critical slowing down and glassy freezing in relaxor ferroelectrics. _Phys. Rev. B_ 61, 14448–14453 (2000).


Article  Google Scholar  * Kholkin, A. et al. Surface domain structures and mesoscopic phase transition in relaxor ferroelectrics. _Adv. Funct. Mater._ 21, 1977–1987 (2011). Article  Google


Scholar  * Rasmussen, C. E. _Evaluation of Gaussian Processes and Other Methods for Non-linear Regression_ (University of Toronto, Toronto, 1999). * Vasudevan, R. K., Jesse, S., Kim, Y.,


Kumar, A. & Kalinin, S. V. Spectroscopic imaging in piezoresponse force microscopy: New opportunities for studying polarization dynamics in ferroelectrics and multiferroics. _MRS


Commun._ 2, 61–73 (2012). Article  Google Scholar  * Jesse, S., Maksymovych, P. & Kalinin, S. V. Rapid multidimensional data acquisition in scanning probe microscopy applied to local


polarization dynamics and voltage dependent contact mechanics. _Appl. Phys. Lett._ 93, 112903 (2008). Article  Google Scholar  * Correa, M., Kumar, A., Priya, S., Katiyar, R. & Scott, J.


Phonon anomalies and phonon-spin coupling in oriented PbFe 0.5 Nb 0.5 O 3 thin films. _Phys. Rev. B_ 83, 014302 (2011). Article  Google Scholar  * Kumar, A., Katiyar, R., Rinaldi, C.,


Lushnikov, S. G. & Shaplygina, T. A. Glasslike state in PbFe 1/2 Nb 1/2 O 3 single crystal. _Appl. Phys. Lett._ 93, 232902 (2008). Article  Google Scholar  * Correa, M., Kumar, A.,


Katiyar, R. & Rinaldi, C. Observation of magnetoelectric coupling in glassy epitaxial PbFe 0.5 Nb 0.5 O 3 thin films. _Appl. Phys. Lett._ 93, 192907 (2008). Article  Google Scholar  *


Peng, W. et al. Spin-lattice coupling in multiferroic Pb (Fe 1/2 Nb 1/2) O 3 thin films. _Appl. Phys. Lett._ 94, 012509 (2009). Article  Google Scholar  * Carpenter, M. A. et al. Elastic and


magnetoelastic relaxation behaviour of multiferroic (ferromagnetic + ferroelectric + ferroelastic) Pb (Fe0. 5Nb0. 5) O3 perovskite. _J. Phys. Cond. Mat._ 27, 285901 (2015). Article  Google


Scholar  * Pradhan, D. K. et al. Studies on dielectric, optical, magnetic, magnetic domain structure, and resistance switching characteristics of highly c-axis oriented NZFO thin films. _J.


Appl. Phys._ 122, 033902 (2017). Article  Google Scholar  * Pradhan, D. K., Barik, S. K., Sahoo, S., Puli, V. S. & Katiyar, R. Investigations on electrical and magnetic properties of


multiferroic [(1− x) Pb (Fe0. 5Nb0. 5) O3− x Ni0. 65Zn0. 35Fe2O4] composites. _J. Appl. Phys._ 113, 144104 (2013). Article  Google Scholar  * Schlom, D. G. et al. Elastic strain engineering


of ferroic oxides. _Mrs Bull._ 39, 118–130 (2014). Article  Google Scholar  * Lu, X., Wang, B., Zheng, Y. & Ryba, E. Critical phase transition temperatures of 1–3 type multiferroic


composite thin films. _J. Phys. D Appl. Phys._ 40, 1614 (2007). Article  Google Scholar  * Smolenskii, G. & Agranovskaya, A. Dielectric polarization of a number of complex compounds.


_Sov. Phys. Solid State_ 1, 1429–1437 (1960). Google Scholar  * Uchino, K. & Nomura, S. Critical exponents of the dielectric constants in diffused-phase-transition crystals.


_Ferroelectrics_ 44, 55–61 (1982). Article  Google Scholar  * Hartigan, J. A. & Wong, M. A. Algorithm AS 136: a k-means clustering algorithm. _Journal of the Royal Statistical Society_.


_Ser. C_ 28, 100–108 (1979). Google Scholar  * Pedregosa, F. et al. Scikit-learn: machine learning in Python. _J. Mach. Learn. Res._ 12, 2825–2830 (2011). Google Scholar  * Jesse, S.,


Baddorf, A. P. & Kalinin, S. V. Switching spectroscopy piezoresponse force microscopy of ferroelectric materials. _Appl. Phys. Lett._ 88, 062908 (2006). Article  Google Scholar  * Suram,


S. K., Haber, J. A., Jin, J. & Gregoire, J. M. Generating information-rich high-throughput experimental materials genomes using functional clustering via multitree genetic programming


and information theory. _ACS Comb. Sci._ 17, 224–233 (2015). Article  Google Scholar  * Ding, S. et al. Combinatorial development of bulk metallic glasses. _Nat. Mater._ 13, 494–500 (2014).


Article  Google Scholar  * Thienhaus, S., Naujoks, D., Pfetzing-Micklich, J., Konig, D. & Ludwig, A. Rapid identification of areas of interest in thin film materials libraries by


combining electrical, optical, X-ray diffraction, and mechanical high-throughput measurements: a case study for the system Ni–Al. _ACS Comb. Sci._ 16, 686–694 (2014). Article  Google Scholar


  * Majumder, S. et al. Dielectric and magnetic properties of sol-gel-derived lead iron niobate ceramics. _J. Appl. Phys._ 99, 024108 (2006). Article  Google Scholar  Download references


ACKNOWLEDGEMENTS The PFM and Gaussian process portion of this research was supported by the U.S. Department of Energy, Office of Science, Materials Sciences and Engineering Division (R.K.V.


and S.V.K.). The synthesis and characterization of samples work was supported by DoD-AFOSR (Grant #FA9550-16-1-0295). D.K.P. and S.K. acknowledge IFN (NSF Grant No. EPS-01002410) for


fellowship. The scanning probe microscopy studies were conducted at the Center for Nanophase Materials Sciences, which is a US DOE Office of Science User Facility. N.L. acknowledges support


from the Eugene P. Wigner Fellowship program at Oak Ridge National Lab. D.K.P. and R.S.K. acknowledge CNMS facilities through CNMS user Proposal ID: CNMS2014-095. E.S. acknowledges support


under the Cooperative Research Agreement between the University of Maryland and the National Institute of Standards and Technology Center for Nanoscale Science and Technology, Award


70NANB10H193, through the University of Maryland. AUTHOR INFORMATION Author notes * Dhiren K. Pradhan Present address: Geophysical Laboratory, Carnegie Institution for Science, Washington,


DC, 20015, USA * Shalini Kumari Present address: Department of Physics and Astronomy, West Virginia University, Morgantown, WV, 26506, USA AUTHORS AND AFFILIATIONS * Department of Physics


and Institute of Functional Nanomaterials, University of Puerto Rico, San Juan, PR, 00936, USA Dhiren K. Pradhan, Shalini Kumari & Ram S. Katiyar * Institute for Research in Electronics


and Applied Physics, University of Maryland, College Park, MD, 207425, USA Evgheni Strelcov * Department of Physics & Astronomy, National Institute of Technology, Rourkela, 769008, India


Dillip K. Pradhan * Center for Nanophase Materials Sciences, Oak Ridge National Laboratory, Oak Ridge, TN, 37831, USA Sergei V. Kalinin, Nouamane Laanait & Rama K. Vasudevan * Institute


for Functional Imaging of Materials, Oak Ridge National Laboratory, Oak Ridge, TN, 37831, USA Sergei V. Kalinin, Nouamane Laanait & Rama K. Vasudevan Authors * Dhiren K. Pradhan View


author publications You can also search for this author inPubMed Google Scholar * Shalini Kumari View author publications You can also search for this author inPubMed Google Scholar *


Evgheni Strelcov View author publications You can also search for this author inPubMed Google Scholar * Dillip K. Pradhan View author publications You can also search for this author


inPubMed Google Scholar * Ram S. Katiyar View author publications You can also search for this author inPubMed Google Scholar * Sergei V. Kalinin View author publications You can also search


for this author inPubMed Google Scholar * Nouamane Laanait View author publications You can also search for this author inPubMed Google Scholar * Rama K. Vasudevan View author publications


You can also search for this author inPubMed Google Scholar CONTRIBUTIONS D.K.P. and S.K. synthesized and characterized the samples. Piezoresponse force microscopy measurements were


performed at ORNL by the advice of R.K.V, E.S., N.L., and S.V.K. The data were analyzed by D.K.P., E.S., D.K.P., N.L., and R.K.V. R.S.K. and S.V.K. supervised this research work. N.L.


developed the Gaussian process modeling, analyzed data, and co-wrote the paper. R.K.V. conceived of the idea, analyzed data, and co-wrote the paper. All authors assisted in writing and


commented on the manuscript. CORRESPONDING AUTHOR Correspondence to Rama K. Vasudevan. ETHICS DECLARATIONS COMPETING INTERESTS The authors declare no competing financial interests.


ADDITIONAL INFORMATION PUBLISHER'S NOTE: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. ELECTRONIC SUPPLEMENTARY


MATERIAL SUPPLMENTARY INFORMATION JUPYTER NOTEBOOK FOR ANALYSIS PN1 - DATASET PN2 - DATASET PN3 - DATASET PFN - DATASET RIGHTS AND PERMISSIONS OPEN ACCESS This article is licensed under a


Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit


to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are


included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and


your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this


license, visit http://creativecommons.org/licenses/by/4.0/. Reprints and permissions ABOUT THIS ARTICLE CITE THIS ARTICLE Pradhan, D.K., Kumari, S., Strelcov, E. _et al._ Reconstructing


phase diagrams from local measurements via Gaussian processes: mapping the temperature-composition space to confidence. _npj Comput Mater_ 4, 23 (2018).


https://doi.org/10.1038/s41524-018-0078-7 Download citation * Received: 14 August 2017 * Revised: 14 March 2018 * Accepted: 29 March 2018 * Published: 25 April 2018 * DOI:


https://doi.org/10.1038/s41524-018-0078-7 SHARE THIS ARTICLE Anyone you share the following link with will be able to read this content: Get shareable link Sorry, a shareable link is not


currently available for this article. Copy to clipboard Provided by the Springer Nature SharedIt content-sharing initiative


Trending News

Drought dries up copper canyon waterfall although some blame mining

A waterfall in the Copper Canyon in Ocampo, Chihuahua, has dried up due to the severe drought affecting the area. The 24...

The bcl-2 pro-survival protein a1 is dispensable for t cell homeostasis on viral infection

ABSTRACT The physiological role of the pro-survival BCL-2 family member A1 has been debated for a long time. Strong mRNA...

5 signs a house is a money pit

The spring home-buying season is in full swing, and if forecasts are correct it’s going to be a busy one. Despite mortga...

Waste-to-biofuels technology ready for global rollout

Breakthrough technology developed by Enerkem converts a multitude of waste material into renewable fuel and chemicals, m...

Small doses of alcohol affect driving skills | science news

Science News was founded in 1921 as an independent, nonprofit source of accurate information on the latest news of scien...

Latests News

Reconstructing phase diagrams from local measurements via gaussian processes: mapping the temperature-composition space to confidence

ABSTRACT We show the ability to map the phase diagram of a relaxor-ferroelectric system as a function of temperature and...

Happy chocolate day 2025: wishes, messages, images, whatsapp and facebook status to share with your loved ones

MAKE CHOCOLATE DAY 2025 EXTRA SPECIAL! SHARE SWEET WISHES, MESSAGES, AND IMAGES WITH YOUR LOVED ONES. GET THE BEST WHATS...

Listen and subscribe to take on today podcast - aarp

TAKE ON TODAY PODCAST EPISODE 1 Bob Edwards and Sarah Lock, AARP Senior Vice President for Policy, discuss ways to impro...

Rebel base taken, mozambique says

MAPUTO, Mozambique — Security forces have captured a major rebel base near the border with South Africa, the official AI...

Erratum: measurements of elastic anisotropy due to solidification texturing and the implications for the earth's inner core

_Nature_ 389, 60–63 (1997) Owing to a typographical error, this Letter appeared under the title “Measurements of electri...

Top