10.10.2014 Views

Local polarization dynamics in ferroelectric materials

Local polarization dynamics in ferroelectric materials

Local polarization dynamics in ferroelectric materials

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Rep. Prog. Phys. 73 (2010) 056502<br />

Figure 30. (a) Antipol<strong>in</strong>g effect observed on epitaxial PZT. (b)<br />

Schematic representation of the effect: (left) pol<strong>in</strong>g at high voltage<br />

and (right) collection of surface charges and formation of opposite<br />

field, lead<strong>in</strong>g to <strong>polarization</strong> switch<strong>in</strong>g. The doma<strong>in</strong> wall between<br />

the upper and lower regions reduces the electrostatic energy of the<br />

head-to-head configuration. In the center schematic the contribut<strong>in</strong>g<br />

potentials from the spherical model are presented. (c) Illustrates the<br />

surface charge distribution at the beg<strong>in</strong>n<strong>in</strong>g and at the end of a<br />

bias<strong>in</strong>g experiment, and after the bias is turned off. (a), (b)<br />

Reproduced with permission from [304], copyright 2005, American<br />

Physical Society.<br />

known by now [305–307]. While the electrochemistry<br />

of classical <strong>ferroelectric</strong>s such as PZT and BTO <strong>in</strong><br />

ultrath<strong>in</strong> films and <strong>in</strong> the presence of water vapor is not<br />

studied, experimental observations suggest that tip-<strong>in</strong>duced<br />

electrochemical processes are extremely common. Typically,<br />

their onset is associated with the formation of ‘switched’<br />

patterns with low contrast and highly disordered and broad<br />

doma<strong>in</strong>s walls. Often the process is associated with the<br />

formation of topographic features follow<strong>in</strong>g the bias pattern.<br />

For high biases and humidities, the complete breakdown of the<br />

film and the formation of particulates are observed. Recent<br />

measurements suggest that the onset of the electrochemical<br />

processes is related to the leakage current flow. Note that<br />

these processes are closely related to the transport properties<br />

of <strong>ferroelectric</strong>s, actively studied <strong>in</strong> the context of resistive<br />

switch<strong>in</strong>g [42, 267, 268, 308], and hence must be considered<br />

<strong>in</strong> this context.<br />

3.3. Theory of tip-<strong>in</strong>duced <strong>polarization</strong> switch<strong>in</strong>g<br />

The analysis of the tip-<strong>in</strong>duced switch<strong>in</strong>g behavior by<br />

PFM necessitates the comparison between macroscopic and<br />

microscopic switch<strong>in</strong>g as a first step of theoretical analysis.<br />

The sem<strong>in</strong>al work by Landauer has demonstrated that<br />

the experimentally observed switch<strong>in</strong>g fields correspond to<br />

S V Kal<strong>in</strong><strong>in</strong> et al<br />

impossibly large (∼10 3 –10 5 kT) values for the nucleation<br />

activation energy [79]. This discrepancy is resolved by<br />

postulat<strong>in</strong>g the presence of discrete switch<strong>in</strong>g centers <strong>in</strong>itiat<strong>in</strong>g<br />

low-field nucleation. The spatial and energy distribution of<br />

the nucleation center has become the central component of all<br />

statistical theories for <strong>polarization</strong> switch<strong>in</strong>g <strong>in</strong> macroscopic<br />

<strong>ferroelectric</strong>s [66, 95]. Notably, this behavior is universal<br />

to all first order phase transitions <strong>in</strong>clud<strong>in</strong>g solidification,<br />

martensitic transformations and phenomena such as plastic<br />

deformation and electrochemical reactions, <strong>in</strong> which the<br />

formation of a new phase is always <strong>in</strong>itiated at a specific<br />

defect site or is controlled by defect <strong>dynamics</strong>. Given that<br />

typical spac<strong>in</strong>gs between extended defects are on the order<br />

of 10–100 nm, even high-quality epitaxial films will have<br />

multiple defects with<strong>in</strong> a typical device structure such as a<br />

micrometer-sized capacitor.<br />

In comparison, <strong>in</strong> an SPM experiment, the field is<br />

concentrated to a small (10–30 nm) volume with<strong>in</strong> a material<br />

that may possess small number of defects or be (extended)<br />

defect free. This opens the pathway for prob<strong>in</strong>g defectmediated<br />

switch<strong>in</strong>g <strong>in</strong> <strong>ferroelectric</strong>s on a s<strong>in</strong>gle defect<br />

level. The orig<strong>in</strong>al work by Landauer [79] utilized the<br />

rigid <strong>ferroelectric</strong> approximation (<strong>polarization</strong> is P or −P ,<br />

doma<strong>in</strong> walls are <strong>in</strong>f<strong>in</strong>itely th<strong>in</strong>) to describe thermo<strong>dynamics</strong><br />

of doma<strong>in</strong> formation <strong>in</strong> a uniform field. This approach was<br />

extended <strong>in</strong> a series of groundbreak<strong>in</strong>g works by Molotskii<br />

et al [309–313] to the analysis of the doma<strong>in</strong> formation<br />

<strong>in</strong>duced by the <strong>in</strong>homogeneous electric field of a biased SPM<br />

probe. This thermodynamic analysis was further developed<br />

by Morozovska et al to account for the f<strong>in</strong>ite electric field<br />

below the probe, surface and bulk screen<strong>in</strong>g and presence of<br />

defects [211, 314]. Below, we develop a general description<br />

of switch<strong>in</strong>g thermo<strong>dynamics</strong> <strong>in</strong> the rigid approximation.<br />

3.3.1. General formalism of the problem. The free energy of<br />

a nucleat<strong>in</strong>g <strong>ferroelectric</strong> doma<strong>in</strong> is [314]<br />

(r,U) = d (r) + Dom (r,U) ,<br />

(3.1a)<br />

where the doma<strong>in</strong> geometry is described by the N-dimensional<br />

parameter vector, r (which can be, e.g., coord<strong>in</strong>ates of all<br />

po<strong>in</strong>ts on a doma<strong>in</strong> wall, or the radius and length <strong>in</strong> the semiellipsoidal<br />

approximation) and U is the electric bias applied to<br />

the local probe <strong>in</strong> the proximity of the surface. The first term<br />

<strong>in</strong> equation (3.1a) is the contribution from the frozen disorder<br />

(i.e. random electric fields <strong>in</strong> the material), d (r), with<strong>in</strong> the<br />

volume of the doma<strong>in</strong>. The second term <strong>in</strong> equation (3.1a) is<br />

the thermodynamic work of switch<strong>in</strong>g<br />

Dom (r,U) = S (r) + p (r,U) + D (r) .<br />

(3.1b)<br />

It comprises the contributions of positive doma<strong>in</strong> wall energy,<br />

S (r) = ψ S S, where S is the doma<strong>in</strong> surface and ψ S is the<br />

doma<strong>in</strong> wall energy density, p (r,U) is the energy density<br />

due to the probe field and D (r) is the de<strong>polarization</strong> energy.<br />

The analysis of the switch<strong>in</strong>g process can be simplified<br />

for a rigid piezoelectric, for which effective material constants<br />

are <strong>in</strong>dependent of the electric field. In this case, the<br />

<strong>in</strong>teraction energy with the probe tip electric field, E p 3 (x),<br />

30

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!