You are on page 1of 10

High temperature electrical resistivity and Seebeck coefficient of Ge2Sb2Te5 thin films

L. Adnane, F. Dirisaglik, A. Cywar, K. Cil, Y. Zhu, C. Lam, A. F. M. Anwar, A. Gokirmak, and H. Silva

Citation: Journal of Applied Physics 122, 125104 (2017); doi: 10.1063/1.4996218


View online: http://dx.doi.org/10.1063/1.4996218
View Table of Contents: http://aip.scitation.org/toc/jap/122/12
Published by the American Institute of Physics
JOURNAL OF APPLIED PHYSICS 122, 125104 (2017)

High temperature electrical resistivity and Seebeck coefficient


of Ge2Sb2Te5 thin films
L. Adnane,1 F. Dirisaglik,2 A. Cywar,1 K. Cil,1 Y. Zhu,3 C. Lam,3 A. F. M. Anwar,1
A. Gokirmak,1 and H. Silva1
1
Department of Electrical and Computer Engineering, University of Connecticut, Storrs, Connecticut 06269, USA
2
Department of Electrical and Electronics Engineering, Eskisehir Osmangazi University, Eskisehir 26480, Turkey
3
IBM Watson Research Center, Yorktown Heights, New York 10598, USA
(Received 14 July 2017; accepted 10 September 2017; published online 25 September 2017)
High-temperature characterization of the thermoelectric properties of chalcogenide Ge2Sb2Te5 (GST)
is critical for phase change memory devices, which utilize self-heating to quickly switch between
amorphous and crystalline states and experience significant thermoelectric effects. In this work, the
electrical resistivity and Seebeck coefficient are measured simultaneously as a function of temperature,
from room temperature to 600  C, on 50 nm and 200 nm GST thin films deposited on silicon dioxide.
Multiple heating and cooling cycles with increasingly maximum temperature allow temperature-
dependent characterization of the material at each crystalline state; this is in contrast to continuous
measurements which return the combined effects of the temperature dependence and changes in the
material. The results show p-type conduction (S > 0), linear S(T), and a positive Thomson coefficient
(dS/dT) up to melting temperature. The results also reveal an interesting linearity between dS/dT and
the conduction activation energy for mixed amorphous-fcc GST, which can be used to estimate one
parameter from the other. A percolation model, together with effective medium theory, is adopted to
correlate the conductivity of the material with average grain sizes obtained from XRD measurements.
XRD diffraction measurements show plane-dependent thermal expansion for the cubic and hexagonal
phases. Published by AIP Publishing. [http://dx.doi.org/10.1063/1.4996218]

I. INTRODUCTION layer is grown on a low doped p-type Si substrate. 300 nm


deep trenches of the 2 mm  2 mm area are opened into SiO2
Ge2Sb2Te5 (GST) has been the most studied phase-
by optical lithography and reactive ion etching (RIE).
change material for phase change memory (PCM) devices.
Tungsten is then deposited by physical vapor deposition
The atomic structure of this chalcogenide material is very sta-
(PVD), and the contacts are formed using second optical
ble in crystalline phases, and it presents an electrical resistiv-
lithography and RIE steps. Bottom tungsten contacts were
ity that is 5–6 orders of magnitude lower compared to the
found to provide better adhesion and contact to the GST
resistivity of the material in the amorphous phase. The face
films, and the misalignment arising from the two lithography
centered cubic (fcc) phase of the material and the hexagonal
steps is not important for these large-scale structures. Thin
(hcp) phase are characterized by very high conductivities due
GST films are deposited by elemental co-sputtering of Ge,
to high carrier density.1,2 Thermoelectric transport—the cou-
Sb, and Te and capped by a 10 nm layer of SiO2 to prevent
pling of electronic and thermal transport—plays an important
evaporation and oxidation (Fig. 1). GST has been found to
role in the operation of PCM devices due to high current den-
form good Ohmic contacts with most metals.11
sities and temperature gradients3–8 and has been observed as
The measurement setup used for simultaneous S(T) and
asymmetric amorphization of GST line cells and polarity R(T) measurements up to high temperatures was described
dependent operation of PCM cells in general.3 However, data in detail in Ref. 12. The bottom metal contacts are probed
on thermal conductivity and Seebeck coefficient of phase- with tungsten probes through the soft GST film and are con-
change materials are still limited. In this work, we report the nected to an HP 4145B semiconductor parameter analyzer.
evolution of the Seebeck coefficient S(T) and the electrical A distance of 20 mm between the two contacts is sufficient
resistance R(T) characteristics of GST as the material is pro- to achieve a temperature difference of DT  10  C. The
gressively changed from amorphous to crystalline fcc and measurements were performed in an enclosed chamber under
then to crystalline hcp. The Seebeck coefficient of a material low vacuum (8  104 kPa) with nitrogen flow to minimize
represents the open circuit voltage generated across a temper- the oxidation of the chuck and probe tips. Two K-type ther-
ature gradient along a sample, which can be approximated as mocouples clamped very close to the contacts for each GST
S ¼ dðV0 Þ=dðDTÞ for small temperature gradients.9,10 sample were used to monitor the temperature on both sides
of the sample. The brass-alloy chuck is 10 mm thick, pro-
II. EXPERIMENTAL SETUP AND MEASUREMENT
viding a good thermal inertia and linear temperature profile
PROCEDURE
on the surface between the probe locations. Two resistive
The electrical resistivity and Seebeck coefficient of GST heaters inserted in the chuck are used to set the temperature
are measured simultaneously using GST thin films with tung- gradient (Fig. 2). In order to reach very high temperatures
sten bottom contacts. A 1 lm thick silicon dioxide (SiO2) (800  C), an inductive heater coil positioned below the

0021-8979/2017/122(12)/125104/9/$30.00 122, 125104-1 Published by AIP Publishing.


125104-2 Adnane et al. J. Appl. Phys. 122, 125104 (2017)

FIG. 1. Top-view optical image and


schematic of the 100 nm thin film GST
sample. The film is deposited on top of
2 mm  2 mm W bottom contacts.

setup is used to generate a high frequency alternating mag- GST behaves like a p-type semiconductor,13 and the
netic field that results in heating of the chuck through contact conductivity ðrÞ can be expressed as
with steel surrounding plates. The temperature gradient can
also be controlled with the inductive heater by moving the rðT Þ ¼ elp ðTÞpðTÞ; (1)
coil from one side to another. The surrounding steel plates
form an oven-like enclosure for heat confinement and heat- where e is the elementary charge, p is the carrier density, and
ing of the probes and probe arms to reduce the probe cooling lp is the low field hole mobility. Both the carrier density and
effect. the mobility are temperature dependent.
When a temperature gradient is set across the sample at a The measured resistance of the sample is scaled to resis-
given stabilized temperature, the I-V characteristics obtained tivity using the resistivity value of the sample measured at
with the semiconductor parameter analyzer simultaneously room temperature using the van der Pauw (vdP) method14
provide the Seebeck voltage, which is the x-axis intercept of and assuming that the geometry factor does not change with
the line, and the resistance between the two contacts, which temperature. The resistivity of a square shaped GST thin film
is the slope of the line (Fig. 3). sample with known film thickness, from the same wafer, is
In the experiments, the maximum temperature differ- measured at room temperature using four probes to contact
ence recorded between the probe arm and the surface of the the film surface [Fig. 5(a)]. According to the vdP technique,
chuck does not exceed 150  C at a chuck temperature of the resistivity of the sample is given by
600  C. The probe needles (Cascade Microtech PTT-24/ R14; 23 þ R12;43
4–25 tungsten needles) are 20 mm long, and its tip diame- q ¼ 4:532 tf ; (2)
2
ter is 2.5 lm. An approximated geometry of a GST sample
with electrical contact (2D cylindrically symmetric tungsten where t is the thickness of the film, R14,23 ¼ V14/I23, and
probe tip on sample tungsten bottom contact) is simulated in R12,43 ¼ V12/I43, and the values of the coefficient f are tabu-
COMSOL to estimate the effect of the cooling by the probe lated for arbitrary values of the resistance ratio R14,23/
arms via a 10 lm diameter tungsten contact on the surface of R12,43.14
the GST film (Fig. 1). A worst-case scenario is simulated by The measured resistivity of the 100 nm thick fcc-GST
setting the temperature of the top surface of a 1 lm thick film in a square shape [configuration in Fig. 5(a)], at room
tungsten contact on the sample surface at 200  C, with the temperature, previously heated up to 200  C, is 504 6 3
bottom surface temperature of the sample (surface of the mXcm (assuming a 5% error on the film thickness).
chuck) at 400  C. The simulations show that outside the A second sample from the same wafer, with 4 inline
small area where the probe tips come into contact (10 lm metal contacts, as shown in Fig. 5(b), is also heated up to
diameter), the temperature of the sample surface is the same
as the chuck temperature (Fig. 4).

FIG. 2. Measurement setup schematic. (1) K type thermocouple, (2) probe


arm with a tungsten probe, (3) GST sample, (4) brass chuck, (5) cartridge
heater, (6) inductive heater coil, (7) ceramic glass base, and (8) surrounding FIG. 3. Typical I-V characteristic obtained at T ¼ 400  C and DT ¼ 4  C.
steel plates. The steel plate on the right is drawn partially transparent to The x-axis intercept (V0) is the Seebeck voltage, and the inverse of the slope
show the components inside. is the resistance between the two contacts.
125104-3 Adnane et al. J. Appl. Phys. 122, 125104 (2017)

FIG. 4. Numerical modeling of the cooling of the sample by the probes. (a)
Temperature of the sample simulated after 60 s sample heating to 400  C.
FIG. 6. GST resistivity versus temperature measurements on the 200 nm
(b) Temperature of the simulated probe tip region while the tip surface is
GST film using the 4-point in-line measurement, scaled from R(T) using the
maintained at 200  C. (c) Temperature on the surface of the sample across
vdP measurement at room temperature after heating and cooling down the
the metal contact.
sample from 200  C and from 300  C at the same rates.

200  C, with the same heating rate (5  C/min), to measure configuration shown in Fig. 5(c) is used for simultaneous
the R(T) characteristic during heating and cooling back to R(T) and S(T) measurements of two different samples during
room temperature. The resistance between the two inner con- the same run. The temperature difference DT ¼ Thot  Tcold
tacts on the surface of the sample is given by is the temperature difference between the two probe loca-
tions, which is varied between 10  C and þ10  C at given
R ¼ V23 =I14 : (3)
stable average temperatures TAVG¼(Thot þ Tcold)/2.
The results obtained from the 2 previous measurements are
III. RESULTS AND DISCUSSION
used to determine the resistivity value of the film at 200  C
A. Continuous R(T) measurements
qroomT
q200 ¼ R200 : (4) Continuously increasing temperature measurements of
RroomT
R(T) were performed on 50 nm, 100 nm, and 200 nm (target
This procedure is repeated for a second anneal temperature thicknesses) GST thin films up to the melting temperature
(300  C) and the R(T) characteristic obtained while heating and scaled to resistivity q(T) using the room-temperature
matches the R(T) obtained during the previous cooling from value after anneal at 200  C as described above (Fig. 7).
200  C (Fig. 6). The obtained resistivities at 200  C and During these measurements, the temperature was increased at
300  C are 75 mXcm and 11 mXcm, respectively. a constant rate of 5  C/min. The two transitions, from amor-
The measured R(T) characteristics for different samples phous to fcc (155  C) and from fcc to hcp (365  C), corre-
are then scaled to resistivity (Fig. 7) using the resistivity spond to the turning points in the curves. The as-deposited
value at 200  C, which assumes the same geometry factor
measured at room temperature.
The sign of the Seebeck coefficient indicates the
majority carrier type (S > 0 for p-type conduction). The

FIG. 5. Schematics of measurement settings showing the position and the


numbering of electrical contacts for (a) the configuration for the room tem- FIG. 7. GST resistivity versus temperature. All the curves represent the 4-
perature measurement and (b) 4-point configuration for the R(T) measure- point resistance measurements obtained with the same heating rate of
ment. (c) 2-point configuration for R(T) and S(T) measurements on 2 5  C/min, scaled to resistivity using the resistivity value of the 200 nm film
different samples. at 200  C (dashed green line, data from Fig. 6).
125104-4 Adnane et al. J. Appl. Phys. 122, 125104 (2017)

amorphous films have a room-temperature resistivity of


8.7  102 Xcm, in agreement with reported values of
8  102 to 9  102 Xcm.4 The films typically start breaking
up and become discontinuous after melting (600  C). The
resistivity of the 100 nm GST film in the hcp phase is calcu-
lated to be slightly higher than that of the 50 nm and 200 nm
films after the second transition phase which can be attributed
to variations in film thicknesses. The drop in resistivity at
585  C as shown in Fig. 7 (inset) indicates the melting of
the film. The liquid resistivity of 1.14 mXcm observed after
this point is also in agreement with previously reported values
for GST.6,15 A film thickness reduction of 5% due to mate-
rial density changes after the first transition has been
reported16 but is not taken into account in our calculations of
resistivity.
The R(T) curves of the as-deposited amorphous GST
films show a drop in resistivity of more than five orders of
magnitude as the material transitions from amorphous to fcc
and hcp crystalline phases. In GST devices, this ratio is only
approximately four orders of magnitude.15 This difference is
attributed to the melt-quench amorphization process in
small-scale devices through electrical pulses, which leads to
lower atomic disorder and hence lower amorphous resistivity
to start with, compared to as-deposited amorphous films.

B. Simultaneous S(T) and R(T) measurements


In order to characterize the temperature-dependent resis-
tance R(T) and Seebeck coefficient S(T) of the material at
each crystalline state as the material progressively changes
from amorphous to crystalline, measurements during multi-
ple heating and cooling cycles with increasingly maximum
temperature were performed (Figs. 8 and 9). The R(T) and FIG. 8. Resistance measured in consecutive cycles to increasingly maximum
temperature for (a) 50 nm and (b) 200 nm GST films. The average tempera-
S(T) characteristics were obtained simultaneously from 2- ture and temperature gradients are regulated at each step for 20 min while
point I-V measurements performed on 2 samples [Fig. 5(c)] measuring the resistance and the Seebeck coefficient. The insets in the
using a parameter analyzer. Due to the required thermal gra- graphs show the resistance vs. 1/kT, from which the activation energies are
obtained. The dashed lines in the insets represent the resistance values of the
dient, and for practical experiment times, the measurements samples at the second transition temperature (Rc0). Each point is the average
are performed in 10  C increments from 40  C up to 300  C, of 15 measurement points, and for all tests but the first, the errors are
in 20  C increments above 300  C, and in 50  C increments smaller than the data markers.
for the last measurement (up to 540  C). The final tempera-
ture was set to 180  C for the first cycle and was increased
by 20  C for each subsequent cycle. gap measurements.21 The activation energy for conduction
The R(T) and S(T) results obtained simultaneously for obtained for the amorphous phase from the first cycle,
50 nm and 200 nm GST films are shown in Figs. 8 and 9. The E ¼ 0.417 eV, is in agreement with reported values,22 and it
R(T) characteristics show the expected exponential decrease decreases as the material crystallizes in the following cycles
in the resistance with increasing temperature, within each of increasingly maximum temperatures. Given the large car-
crystalline state, as the material progressively crystallizes rier concentration in GST,23,24 the overall decrease in the
from amorphous to fcc and from fcc to hcp.17,18 The conduc- resistance from one annealing cycle to another is expected to
tivity of the film below the glass transition (100  C)19 fol- be mostly due to the increase in mobility with crystallization.
lows an Arrhenius dependence (Fig. 8 insets) After the fcc to hcp transition (last four R(T) curves), the
material shows a metallic behavior with the resistance
r ¼ r0 eE=kT ; (5) increasing with temperature due to mobility degradation, in
agreement with previous reports.5,16,25
where E is the conduction activation energy, k is the Once the material starts crystallizing, and for each state
Boltzmann constant, T is the absolute temperature, and r0 is (corresponding to the material annealed at the previous cycle
defined as the minimum metallic conductivity.20 In the amor- maximum temperature), the Seebeck coefficient increases
phous phase, the activation energy E corresponds to the linearly with temperature until further crystallization occurs
energy for sub-band hopping mechanism which does not beyond the previous anneal temperature. Linear fits of these
occur in fcc and hcp-GST based on UV/visible/NIR band regions of S(T) curves, with fixed zero intercept (0 mV/K at
125104-5 Adnane et al. J. Appl. Phys. 122, 125104 (2017)

FIG. 10. Seebeck coefficient versus temperature measurement data on


50 nm and 200 nm thin films. The dashed lines represent the linear fit of the
data with the 0 lV/K intercept. The numbers in the legend represent the tem-
perature in Celsius to which the sample was previously annealed.

linear dependence of dS/dT with E, together with the linear


S(T) with S ¼ 0 V/K at 0 K, can still be used to empirically
estimate one transport parameter from the other. A few low-
temperature measurements of conductivity (to obtain the
activation energy) can be used to determine dS/dT, and
hence, S(T) or a single measurement of Seebeck coefficient
can be used to determine the conductivity activation energy
E. The different slopes of dS/dT vs. E for the 50 nm and
200 nm films may be due to a mismatch of the activation
energies for the two thicknesses since the anneal temperature
FIG. 9. Seebeck coefficient versus temperature measured simultaneously may not result in the same crystallinity state [Fig. 11(a)].
with the R(T)s in Fig. 8 for (a) 50 nm and (b) 200 nm GST films. The insets The positive Seebeck coefficient obtained for the GST
show the slopes dS/dT versus the conductivity of the sample at room tem- films confirms the p-type conduction of the material up to
perature. Standard deviations for each point are shown as error bars, but
these are not visible in this scale.
the maximum measured temperature of 800 K. The S(T)
decrease with temperature for the amorphous phase, as well
0 K, Fig. 10), result in very small relative standard deviation as the increase with temperature for the stable hcp phase, is
errors on the slopes (0.6–3.7%). in agreement with previous reports.5,6 For the fcc phase,
The observed linear dependence of the Seebeck coeffi- however, these multiple stepped measurements show that the
Seebeck coefficient also increases with temperature for each
cient with temperature is consistent with the Boltzmann
crystalline state, up to the maximum temperature reached in
approximation transport model for degenerate semiconduc-
the previous cycle. The decrease in S(T) with temperature
tors, which results in26–28
observed for amorphous GST and fcc-GST after certain tem-
 2=3 perature is therefore due to crystallization rather than bipolar
8p2 kB2  p
S¼ 2
m p T ; (6) conduction which explains the S(T) turn-around in stable
3eh 3p
semiconductors as they approach the intrinsic regime. A pos-
where kB is the Boltzmann constant, h is the Plank constant, itive or negative slope in S(T) determines the direction of
e is the elementary charge, p is the carrier density, mp* is the asymmetry of the thermal profile (Thomson effect) that has
carrier effective mass, and T is the absolute temperature. been observed in phase-change memory devices and has
Figure 11(a) shows the conductivity activation energy important implications related to power and reliability of the
and dS(T)/dT as a function of anneal temperature for the devices. The positive S(T) slope we observe for the crystal-
50 nm and 200 nm films. The derivative dS/dT, / mp*/p2/3, line material (both fcc and hcp) is consistent with observa-
shows a linear dependence on the conductivity activation tions of asymmetric amorphization of PCM line cells toward
energy for the amorphous-fcc mixed phase, where this the higher potential terminal.3,29,30
energy is positive [Fig. 11(b)]. dS/dT is closely related to the
C. Average grain size from XRD measurements
electrical conductivity, e2psp/mp*, where sp is the carrier
relaxation time. Decoupling the relative changes in the car- Room temperature XRD measurements on pre-annealed
rier concentration, effective mass, and relaxation time as the samples and in-situ XRD measurements (at stabilized tem-
material crystallizes would require Hall measurements. The peratures during anneal)31 were performed to correlate R(T)
125104-6 Adnane et al. J. Appl. Phys. 122, 125104 (2017)

FIG. 12. XRD patterns of the 100 nm GST film annealed at different temper-
atures and then cooled down from 400  C to obtain the pattern for hcp-GST
at 30  C. The peaks around 47.5 and 55 , present in all patterns, likely orig-
inate from the sample holder and are not related to the GST sample.

0:9k
gs ¼ ; (7)
b cos h

where k is the X-ray wavelength and h is the diffraction


angle. XRD measurements have been previously reported for
FIG. 11. (a) Activation energy and dS/dT as a function of anneal tempera- hcp-GST annealed up to 330  C, and the average grain size
ture. The activation energies for the 50 nm film are slightly higher than those was found to be 40.3 nm.41
for the 200 nm film, probably due to the difference in the crystallinity of the
The data for average grain size versus temperature were
two films annealed at the same temperature. (b) Derivative of Seebeck ver-
sus temperature dS/dT versus the conductivity activation energy E for 50 nm interpolated with a third order polynomial fit to correlate
and 200 nm GST thin films, showing a linear dependence for the amor- XRD grain size results with R(T) and S(T) measurements
phous-fcc mixed phase region. Standard deviations for dS/dT are shown as obtained after different anneal temperatures (Fig. 13). Grain
error bars. The dashed lines show the linear fits for the two samples, using
all points of positive conduction activation energy. sizes calculated from impedance spectroscopy (IS) measure-
ments by Huang et al., assuming a brick model,42 fall in the
same range but show a significantly different trend with the
and S(T) with grain sizes in mixed-phase GST. Figure 12 anneal temperature (downward curvature, square symbols in
shows the evolution of XRD patterns with temperature for Fig. 13).
an as-fabricated amorphous 100 nm GST film as the chuck
temperature was increased from room temperature to
D. Percolation model
585  C in 100  C steps and decreased back to room tempera-
ture. The peaks in the pattern of the amorphous sample are Since the increase in average grain size in fcc-GST is
repeated in all patterns, suggesting that these originate from only from 18 nm to 30 nm, the drastic change in conduc-
the silicon substrate and sample holder. The fcc phase of the tivity observed at the amorphous to fcc transition is likely
sample at 200  C and 300  C is identified by the peaks at due to percolation paths that form in the material as both the
25.5 , 29.4 , 42.4 , and at 52.6 , while the hcp phase exhib- number of grains and the grain size increase. The electrical
its different peaks at the angles 21.4 , 25.7 , 28.9 , 40.1 , conductivity of an inhomogeneous material formed by two
42.8 , 48.5 , and 52.9 corresponding to different planes, materials with different conductivities was described in the
similar to what has been previously observed.32–36 After late 1800s by Rayleigh43 and Wiener44 and then formulated
cooling down the sample to room temperature, a new pat- into the effective-medium theory by Bruggeman45 for vari-
tern is acquired, and it shows a small shift to the right in all ous shapes and by Landauer46 for spheres of conductivity rc
hcp peaks, which is due to the thermal expansion of the embedded in a material of conductivity ra. The model was
material at high temperatures.37 From these results, the adopted recently for mixed phase amorphous-crystalline
highest increase in the hcp planes spacing at 400  C is chalcogenides such as GST.47 In this work, we model the
1.45% for the (004) direction. mixed phase amorphous-fcc GST as spheres of fully crystal-
The average grain size at different temperatures was line fcc-GST embedded in amorphous GST and use the aver-
obtained from the XRD patterns using Scherrer’s equa- age grain sizes obtained from XRD measurements as the
tion38,39 which relates the full-width-half-max (FWHM) b to diameter of the spheres. The conductivity of the mixed mate-
the grain size (gs)40 rial in this case is given by45–48
125104-7 Adnane et al. J. Appl. Phys. 122, 125104 (2017)

FIG. 14. Calculated GST mixed-phase resistivity as a function of the crystal-


FIG. 13. Average grain size calculated from XRD measurements on the line fraction using Eq. (8) with amorphous and crystalline GST resistivities
200 nm thin GST film. In-situ measurements were done in 25  C steps from qa ¼ 494 Xcm and qc ¼ 0.01 Xcm, respectively.
150  C to 585  C (green circles). Room temperature measurements were
also done on different samples (from the same wafer) pre-annealed at differ- R ¼ R0 expðE=kB T Þ; (11)
ent temperatures for 10 minutes, with a heating rate of 2  C/min (red stars).
Interpolated data were calculated using the third order polynomial fit (black
dashed line) to obtain grain sizes for the anneal temperature values used in
where R0 and E are the pre-factor and conduction activation
the S(T) and R(T) measurements. Grain sizes calculated from room- energy of the mixed material. The parameters for amorphous
temperature impedance spectroscopy measurements by Huang et al. (blue GST Ra0 and Ea are obtained from the first R(T) measure-
squares) are in the same range but show a significantly different trend with ment on the amorphous film (Fig. 8)
anneal temperature.42
Ra ¼ Ra0 expðEa =kB T Þ: (12)
 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 The resistance of fully crystalline fcc-GST is assumed to be
r ¼ 2rp  rq þ ð2rp  rq Þ2 þ 8ra rc (8)
4 constant with temperature (E ¼ 0 eV) since the activation
energy is observed to change from positive values for the
with
mixed amorphous-fcc phase to negative values for the mixed
rp ¼ ð1  f Þra þ f rc ; fcc-hcp phase (see dashed lines in Fig. 8, insets). This con-
stant resistance value for fully crystalline fcc-GST is taken
rq ¼ ð1  f Þrc þ f ra ; as the resistance at the second transition temperature (the
zero of the second derivative of the R(T) ’envelope’ formed
where rc is the conductivity of the crystalline spheres, ra is
by the last data point from each cycle)
the conductivity of the amorphous matrix, and f is the crys-
tallinity fraction of the material (fraction of the crystalline Rc ¼ Rc0 : (13)
volume to the total volume of the sample). Using rc and ra
obtained from our R(T) measurements (r ¼ 1/q, qa ¼ 494 The different parameters, Ra0, Ea, and Rc0, are given in Table I
Xcm, and qc ¼ 0.01 Xcm, explained below), the calculated for the 50 nm and 200 nm thick films and are used to calculate
mixed phase resistivity of GST (Fig. 14) shows a sharp drop the crystallinity fraction f of the sample as a function of tem-
of 3 orders of magnitude at the percolation threshold of perature from Eq. (10).
33% crystalline fraction. Figure 15(a) shows the calculated crystallinity fraction f
Solving Eq. (8) for f yields along with the corresponding experimental S(T) slopes of
the mixed-phase material as a function of anneal temperature
ðr  ra Þð2r þ rc Þ for the 50 nm and 200 nm thick films. The crystallinity frac-
f ¼ ; (9)
3rðrc  ra Þ tion of the 200 nm thin film is slightly higher than that of the
50 nm film. Figure 15(b) also shows the experimental dS/dT
which can be written in terms of the resistances assuming a (/ m/p2/3) and room-temperature conductivity rRT (/ ms/p)
constant geometry factor C, r¼C/R as

ðRa  RÞðR þ 2Rc Þ TABLE I. Amorphous and crystalline resistance pre-factors and activation
f ¼ ; (10)
3RðRa  Rc Þ energies extracted from the R(T) measurements.

where R, Ra, and Rc are the resistances of the mixed phase, 50 nm thin film 200 nm thin film
the amorphous phase, and the fully crystalline fcc GST,
Ra0 (X) 77.05 56.99
respectively. Ea (meV) 419.8 376.7
The resistance of the mixed phase material at each crys- Rc0 (X) 3558 1104
tallinity fraction f follows an Arrhenius dependence:
125104-8 Adnane et al. J. Appl. Phys. 122, 125104 (2017)

FIG. 16. Crystallinity fraction f for the 200 nm GST thin film sample as a
function of the cubic grain size (gs)3 obtained from XRD patterns for the
mixed phase amorphous-fcc region, pre-annealed at increasing anneal tem-
peratures (180  C to 320  C). The intermediate linear region suggests growth
dominated crystallization for this anneal temperature range.

of the crystals is reached and the increase in crystallinity


is mainly due to crystallization of the remaining gaps
between the crystals.

IV. SUMMARY
Simultaneous measurements of the temperature depen-
FIG. 15. (a) Calculated crystalline fraction f and experimental dS/dT as a
dent electrical resistance and Seebeck coefficient of GST
function of anneal temperature for the 50 nm and 200 nm thin GST films. (b)
Experimental dS/dT (/ m/p2/3) and room-temperature conductivity rRT (/ (Ge2Sb2Te5) thin films were performed up to 540  C.
ms/p) versus calculated crystalline fraction f for 50 nm and 200 nm thin film Repeated measurements to increasingly maximum tempera-
samples. tures allow characterization of the properties of each state (in
contrast to continuous measurements which show the convo-
versus calculated crystalline fraction f for the 50 nm and
luted effects of the temperature dependence of the transport
200 nm thin film samples. These relations can now be used to
parameters and material crystallization during heating, as
estimate S(T) and r(T) for a given amorphous-fcc mixed phase
also shown here up to melting temperature). The measured
material from a single, room-temperature value of the conduc-
resistance was scaled to resistivity using the 4 point room-
tivity. rRT can be used to determine f and the corresponding
temperature resistivity measurement performed on a sample
dS/dT and hence S(T), which can then be used to determine the
of the same film. The R(T) characteristics measured at differ-
conduction activation energy E (Fig. 11) to obtain the full r(T).
ent crystalline states follow an Arrhenius dependence with a
Since the model for conductivity we have used assumes
decreasing activation energy as the material crystallizes. The
spherical crystals in an amorphous matrix and f is proportional
S(T) results show p-type conduction until melting tempera-
to (gs)3 times the number of the grains in the sample, it is
ture and linear S(T) characteristics for each state above glass
interesting to look at the relationship between f and (gs)3
transition temperature, in agreement with the degenerate
obtained from XRD (Fig. 16), which appears to show three
semiconductor transport model. The measurements also
distinct regions for crystallization:
show that both the mixed amorphous-fcc and the mixed
(1) 180  C < T < 220  C: the crystallinity of the sample fcc–hcp phases exhibit a constant positive Thomson coeffi-
increases rapidly with slow grain growth (Df/Dgs3 cient (dS/dT). A linear relationship between the slope of
¼ 1.1  104 nm3), suggesting that nucleation in the S(T) characteristics and the activation energy of the
material is dominant. Arrhenius conduction for the GST material is also observed
(2) 220  C < T < 300  C: f changes linearly with (gs)3 at a for the mixed amorphous-fcc phase. This observation can be
lower rate compared to the previous region (Df/ useful to estimate the temperature-dependent conductivity or
Dgs3¼5.1  105 nm3), suggesting that the number of Seebeck coefficient of GST from a single measurement of
crystals is approximately constant and the increase in the Seebeck coefficient or a few conductivity measurements
crystallinity is mainly related to the growth of the at low temperature points (to obtain the conduction activa-
crystals. tion energy). In-situ XRD measurements on a 100 nm
(3) 300  C < T < 320  C: the crystallinity again increases thin GST film sample showed the plane-dependent thermal
rapidly with a small increase in grain size (Df/ expansion of the hcp phase, with a maximum thermal expan-
Dgs3¼1.1  104 nm3), implying that the critical size sion observed for the (004) plane.
125104-9 Adnane et al. J. Appl. Phys. 122, 125104 (2017)

11
A percolation model of conducting spheres in a lower L. W.-W. Fang, Z. Zhang, R. Zhao, J. Pan, M. Li, L. Shi, T.-C. Chong,
and Y.-C. Yeo, J. Appl. Phys. 108, 53708 (2010).
conductivity matrix (which predicts the sharp drop observed 12
L. Adnane, N. Williams, H. Silva, and A. Gokirmak, Rev. Sci. Instrum.
in the resistance with a relatively small change in the grain 86, 105119 (2015).
13
size, at a critical crystalline density threshold) was applied to A. Pirovano, A. A. L. Lacaita, A. Benvenuti, F. Pellizzer, and R. Bez,
the measured data to relate the electrical transport parame- IEEE Trans. Electron Devices 51, 452 (2004).
14
L. J. Van der Pauw, Philips Tech. Rev. 20, 220–224 (1958).
ters and the crystalline grain sizes obtained from XRD meas- 15
K. Cil, F. Dirisaglik, L. Adnane, M. Wennberg, A. King, A. Faraclas, M.
urements. The R(T) results of mixed phase amorphous-fcc B. Akbulut, Y. Zhu, C. Lam, A. Gokirmak, and H. Silva, IEEE Trans.
are then expressed in terms of the amorphous and fully crys- Electron Devices 60, 433 (2013).
16
talline fcc temperature-dependent resistance values using the I. Friedrich, V. Weidenhof, W. Njoroge, P. Franz, and M. Wuttig, J. Appl.
Phys. 87, 4130 (2000).
effective-medium theory to obtain the temperature depen- 17
D. Fugazza, D. Ielmini, G. Montemurro, and A. L. Lacaita, in Proceedings
dent crystallinity fraction of the material. of the International Electron Devices Meeting (IEEE, 2010), pp.
This study focused on the properties of GST up to the 29.3.1–29.3.4.
18
C. David Wright, M. Armand, and M. M. Aziz, IEEE Trans. Nanotechnol.
second phase transition (fcc–hcp), which is the more techno-
5, 50 (2006).
logically relevant range as phase-change memory devices do 19
J. A. Kalb, M. Wuttig, and F. Spaepen, J. Mater. Res. 22, 748 (2007).
20
not experience the slow fcc–hcp transition during switching. N. F. Mott and E. A. Davis, Electronic Processes in Non-Crystalline
A similar percolation and effective medium theory model Materials (Clarednon Press, Oxford, 1971), p. 591.
21
F. Rao, Z. Song, Y. Cheng, M. Xia, K. Ren, L. Wu, B. Liu, and S. Feng,
can however be applied to the mixed fcc–hcp phase after the Acta Mater. 60, 323 (2012).
second transition for a fuller understanding of the crystalliza- 22
A. Redaelli, A. Pirovano, I. Tortorelli, D. Ielmini, and A. L. Lacaita, IEEE
tion dynamics and transport properties in chalcogenide Electron Device Lett. 29, 41 (2008).
23
glasses. H.-K. K. Lyeo, D. G. Cahill, B.-S. S. Lee, J. R. Abelson, M.-H. H. Kwon,
K.-B. B. Kim, S. G. Bishop, and B. Cheong, Appl. Phys. Lett. 89, 151904
(2006).
ACKNOWLEDGMENTS 24
A. Voraud, M. Horprathum, P. Eiamchai, P. Muthitamongkol, B.
Chayasombat, C. Thanachayanont, A. Pankiew, A. Klamchuen, D.
The GST films were deposited at IBM T.J. Watson Naenkieng, T. Plirdpring, A. Harnwunggmoung, A. Charoenphakdee, W.
Research Center and characterized at UConn. This work was Somkhunthot, and T. Seetawan, J. Alloys Compd. 649, 380 (2015).
25
partially supported by the U.S. National Science Foundation T. Siegrist, P. Jost, H. Volker, M. Woda, P. Merkelbach, C.
Schlockermann, and M. Wuttig, Nat. Mater. 10, 202 (2011).
through Award Nos. ECCS 0925973 and ECCS 1150960. 26
M. Cutler and J. F. Leavy, Phys. Rev. 133, A1153 (1964).
The characterization and analysis efforts of L.A. were 27
G. J. Snyder and E. S. Toberer, Nat. Mater. 7, 105 (2008).
28
supported by the U.S. Department of Energy, Office of J. P. Heremans, V. Jovovic, E. S. Toberer, A. Saramat, K. Kurosaki, A.
Science, Basic Energy Sciences under Award No. DE- Charoenphakdee, S. Yamanaka, and G. J. Snyder, Science (80-.) 321,
1457 (2008).
SC005038. L.A. also acknowledges a graduate fellowship 29
F. Dirisaglik, G. Bakan, A. Faraclas, A. Gokirmak, and H. Silva, Int. J.
from the Graduate Assistance in Areas of National Need High Speed Electron. Syst. 23, 1450004 (2014).
30
(GAANN). F.D. and K.C. acknowledge graduate fellowships A. Faraclas, G. Bakan, L. Adnane, F. Dirisaglik, N. E. Williams, A.
Gokirmak, and H. Silva, IEEE Trans. Electron Devices 61, 372 (2014).
from the Republic of Turkey Ministry of National Education. 31
K. Cil, Dr. Diss. (2015).
A.C. was supported through an NSF GRFP fellowship. The 32
D. Dimitrov and H.-P. Shieh, Mater. Sci. Eng. B 107, 107 (2004).
33
authors would also like to thank Simone Raoux, Norma Sosa, N. Kato, I. Konomi, Y. Seno, and T. Motohiro, Appl. Surf. Sci. 244, 281
and Matthew BrightSky for their contributions to sample (2005).
34
Y. Yin, H. Sone, and S. Hosaka, Jpn. J. Appl. Phys., Part 1 44, 6208
fabrication at the IBM T.J. Watson Research Center. (2005).
35
A. Voraud, M. Horprathum, P. Eiamchai, P. Muthitamongkol, C.
1
Z. Sun, S. Kyrsta, D. Music, R. Ahuja, and J. M. Schneider, Solid State Thanachayanont, W. Somkhunthot, and T. Seetawan, Surf. Coat. Technol.
Commun. 143, 240 (2007). 291, 15 (2016).
2 36
D. Ielmini, in Proceedings of the International Conference on K. S. Bang, Y. J. Oh, and S.-Y. Lee, J. Electron. Mater. 44, 2712 (2015).
Simulations of Semiconductor Processes and Devices (SISPAD) (IEEE, 37
C.-A. Jong, W. Fang, C.-M. Lee, and T.-S. Chin, Jpn. J. Appl. Phys., Part 1
2009), pp. 1–8. 40, 3320 (2001).
3 38
D. T. Castro, L. Goux, G. A. M. Hurkx, K. Attenborough, R. Delhougne, A. L. Patterson, Phys. Rev. 56, 978 (1939).
J. Lisoni, F. J. Jedema, M. A. A. In’t Zandt, R. A. M. Wolters, D. J. 39
J. I. Langford and A. J. C. Wilson, IUCr J. Appl. Crystallogr. 11, 102
Gravesteijn, M. A. Verheijen, M. Kaiser, R. G. R. Weemaes, and D. J. (1978).
40
Wouters, IEEE Int. Electron Devices Meet. 315 (2007). B. Cullity, Elements of X-ray Diffraction (Addison–Wesley, Reading, MA,
4
H.-S. P. Wong, S. Raoux, S. Kim, J. Liang, J. P. Reifenberg, B. Rajendran, 1967), p. 262.
41
M. Asheghi, and K. E. Goodson, Proc. IEEE 98, 2201 (2010). Y. Hu, H. Zou, J. Zhang, J. Xue, Y. Sui, W. Wu, L. Yuan, X. Zhu, S.
5
J. Lee, T. Kodama, Y. Won, M. Asheghi, and K. E. Goodson, J. Appl. Song, and Z. Song, Appl. Phys. Lett. 107, 263105 (2015).
42
Phys. 112, 14902 (2012). Y.-H. Huang, Y.-J. Huang, and T.-E. Hsieh, J. Appl. Phys. 111, 123706
6
T. Kato and K. Tanaka, Jpn. J. Appl. Phys., Part 1 44, 7340 (2005). (2012).
7 43
N. Ciocchini, M. Laudato, and A. Leone, IEEE Trans. Electron Devices L. Rayleigh, London Edinburgh Dublin Philos. Mag. J. Sci. 34, 481
62, 3264 (2015). (1892).
8 44
J. Lee, M. Asheghi, and K. E. Goodson, Nanotechnology 23, 205201 O. Wiener, Phys. Z. 5, 332 (1904).
45
(2012). D. A. G. Bruggeman, Ann. Phys. 416, 636 (1935).
9 46
R. R. Heikes and R. W. Ure, Thermoelectricity: Science and Engineering R. Landauer, J. Appl. Phys. 23, 779 (1952).
47
(Interscience Publishers, 1961). D.-H. Kim, F. Merget, M. Laurenzis, P. H. Bolivar, and H. Kurz, J. Appl.
10
G. S. Nolas, J. Sharp, J. Goldsmid, and H. J. Goldsmid, Thermoelectrics: Phys. 97, 83538 (2005).
48
Basic Principles and New Materials Developments (Springer Science and R. C. McPhedran and D. R. McKenzie, Proc. R. Soc. London, Ser. A 359,
Business Media, 2013). 45 (1978).

You might also like