Computational Materials Science: F. Salazar, Luis A. Pérez

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Theoretical study of electronic and mechanical properties of GeC nanowires

F. Salazar
1
, Luis A. Prez

Instituto de Fsica, Universidad Nacional Autnoma de Mxico, Apartado Postal 20-364, 01000 Mxico, DF, Mexico
a r t i c l e i n f o
Article history:
Received 1 February 2012
Received in revised form 30 May 2012
Accepted 30 May 2012
Available online 28 June 2012
Keywords:
GeC nanowires
Electronic properties of nanowires
Mechanical properties of nanowires
a b s t r a c t
In this work, we present a density functional study of the electronic band structure, the Youngs modulus
and Poisson ratio of hydrogen-passivated germanium carbide (GeC) nanowires, with diamond structure
and grown along the [001] and [111] crystallographic directions, as function of their diameter. The
results obtained are compared with those calculated for the corresponding silicon carbide (SiC), germa-
nium (Ge) and silicon (Si) passivated nanowires. The band gaps of GeC and SiC passivated nanowires are
quite similar between them and larger than those of the Ge and Si nanowires. For all studied nanowires,
the Youngs modulus have a lower value than the bulk one and it increases as function of the diameter
converging to the bulk value. Both carbide nanowires (SiC and GeC) show higher Youngs modulus values
than the nanowires without carbon (Si and Ge). Moreover, the nanowires grown along [001] direction
show a lower Youngs modulus in comparison to those grown along [111] direction, since the covalent
bonds are oriented along this direction. Finally, the results show that for both crystallographic directions
the carbide nanowires have a more compact structure in comparison to the corresponding non-carbide
ones, and then the former nanowires have a higher resistance to be deformed.
2012 Elsevier B.V. All rights reserved.
1. Introduction
Semiconductor nanowires (NWs) have attracted the attention of
scientists in the last two decades due to their potential technolog-
ical applications in electronic and optoelectronic devices [13], as
well as their promising thermoelectrical properties [4,5] which
could be used to generate clean energy. Experimental procedures
have been developed to control the axial and radial (homogeneous
and heterogeneous) growth of NWs along different crystallo-
graphic directions with diameters ranging from hundred of nano-
meters to 10 nm [6]. Even more, it is possible to synthesize NWs
whose surface is chemically steady (passivated surface) [6]. There
are excellent reviews where the different experimental techniques
for their synthesis are described [7,8]. A passivated surface is usu-
ally a fundamental step to obtain a semiconductor behavior in
NWs [9], for example, bare silicon NWs can have either metallic
or semimetallic behavior since the atoms in the surface of the
NW could form dimmers and then their coordination number
could change [10]. The energy band gap is highly sensitive to the
growth orientation [11], surface atoms and NW diameter [12,11].
Cruz-Irisson et al. showed how new bands appear in the energy
bandgap when one or more hydrogen atoms are removed from
the surface of passivated silicon carbide NWs [13]. On the other
hand, in passivated silicon NWs it seems that the bandgap does
not to depend on the cross section of the NW when the surface-
to-volume ratio is conserved [14]. The mechanical behavior of
NWs is another interesting and intriguing property. The Youngs
modulus of thin metallic NWs is larger than the bulk value and
diminishes to this value when the diameter is increased [15,16],
the same behavior was found in CuO [17] and WO
3
NWs [18]. In
contrast, the Youngs modulus of semimetallic Si, SiC, and GaN
NWs is lower than the bulk value and converges to that when
the diameter is increased [15,1921]. Since there is no consensus
to explain this behavior, theoretical studies are important to
understand the physics involved.
On the other hand, density functional calculations suggest that
the GeC alloy, in zincblende phase, is stable at high pressures and it
has an indirect band gap of 2.46 eV [22]. Moreover, GeC might
present peculiar properties for optoelectronic applications due to
its wide band gap [23]. From the experimental side, Ge
x
C
1x
has
been prepared, by deposition, as thin lms of amorphous material,
which have coating properties with good environmental durability
and it can be deposited with a refractive index between 2 and 4
[24]. In this paper we present a theoretical study of electronic
and mechanical properties of hydrogen-passivated GeC NWs using
density functional theory. The results obtained are compared with
those calculated for Si, Ge, and SiC NWs. In Section 2 we describe
the morphologies of the studied NWs, the methodology used and
the lattice constants obtained for each NW. In Section 3, we ana-
lyze the band structures of GeC NWs and the band gaps in compar-
ison with those of Si, Ge, and SiC hydrogen-passivated NWs. In
0927-0256/$ - see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.commatsci.2012.05.066

Corresponding author. Tel.: +52 55 56225183; fax: +52 55 56225008.


E-mail address: lperez@sica.unam.mx (L.A. Prez).
1
Present address: ESIME-Culhuacan, Instituto Politcnico Nacional, Av. Santa Ana
1000, 04430 Mxico, DF, Mexico.
Computational Materials Science 63 (2012) 4751
Contents lists available at SciVerse ScienceDirect
Computational Materials Science
j our nal homepage: www. el sevi er . com/ l ocat e/ commat sci
Section 4 we present the Youngs modulus and the Poissons ratio
for the NWs studied in the previous sections. Finally, the results are
summarized in Section 5.
2. Systems and methodology
In this work we studied Si, Ge, SiC and GeC NWs with zinc-
blende structure and with their surface passivated with hydrogen
atoms. The crystallographic directions of growth of the studied
NWs are [001] and [111]. The NWs are modeled by means of peri-
odic supercells along the growth direction and with a vacuum
space of ve lattice constants in the plane perpendicular to that
direction. In this work, four diameters were considered for the
[001] direction and three diameters for the [111] direction. The
electronic and mechanical properties of these NWs were obtained
by using density functional theory (DFT) within the local density
approximation (LDA), as implemented in the SIESTA code [25],
which employs standard norm-conserving pseudopotentials [26]
in their fully nonlocal form [27]. The pseudopotentials for Si, Ge
and C were generated with the atomic valence-electron congura-
tions s
2
p
2
. The core radii (in atomic units) are as follows: s (1.89), p
(1.89), d (1.89) for Si; s (2.06), p (2.85), d (2.58) for Ge; and s (1.25),
p (1.25), d (1.25) for C. In the present calculations we used double-f
s,p-basis and single d orbital and an energy cutoff of 150 Ry to de-
ne the real space grid for numerical integrations. All the NWs
were let free to relax until the HellmannFeynman forces were less
than 1 meV/.
In order to check the suitability of these pseudopotentials, we
performed calculations for bulk Si, Ge and C with diamond struc-
ture. We have found that the predicted lattice constants are
5.41 , 5.66 , and 3.57 , respectively. The corresponding experi-
mental values are 5.43 , 5.658 , and 3.567 [28]. Moreover, the
calculated SiSi, GeGe and CC bulk bond lengths are 2.34 ,
2.45 , and 1.54 , respectively, while the corresponding experi-
mental values are 2.35 , 2.45 , and 1.54 [28]. Both physical
quantities agree with the corresponding experimental values with-
in an error less than 0.5% for Si, Ge and C.
Fig. 1ad shows the side and cross-section views of the four ini-
tial morphologies of GeC NWs along the [001] direction, where the
blue
2
, gray, and white spheres represent Ge, C and H atoms, respec-
tively. The same geometries were used for Si, Ge, and SiC NWs.
After the relaxation process, the diameter is obtained by projecting
the spatial coordinates in the xy plane and the distances between
the hydrogen atoms in opposite sides are averaged (a and b) as
shown the Fig. 1a. The diagonal of the square formed by a and b
is taken as the diameter of the NW. In all cases the cross section
is taken as a circumference whose diameter is the result of the pre-
vious process.
Likewise, the initial morphologies for GeC NWs grown along
[111] direction are depicted in Fig. 2 for the three diameters con-
sidered. The same structures were used for [111] Si, Ge, and SiC
NWs. The diameter of these NWs was taken as the average length
of the hydrogen atoms in opposite sides.
3. Electronic properties
The band structure of the NWs was obtained for the closed tra-
jectory LCZ and LCZL for [001] and [111] growth directions
respectively, where L and C are the usual crystallographic points
and Z indicates a point along the growth direction of the NW.
Fig. 3 shows the band structures of hydrogen-passivated [001]
GeC NWs. The diameters (d) considered and their corresponding
energy gaps (D) are (a) d = 7.32 , D = 4.86 eV, (b) d = 11.91 ,
D = 3.32 eV, (c) d = 16.40 , D = 2.62 eV, and (d) d = 20.88 ,
D = 2.24 eV. As expected from quantum connement, the energy
gap augments when the wire diameter diminishes. Moreover, for
all the [001] GeC NWs considered, the energy gap is direct. It is
worth mentioning that for the Si, Ge, and SiC NWs grown along
the [001] direction, the behavior is the same that for GeC NWs.
A similar behavior is found for hydrogen-passivated GeC NWs
grown along [111] direction, which band structures are presented
in Fig. 4 and their diameters and corresponding energy gaps being
Fig. 1. (Color online) Side and cross-section views of the initial structures of the
hydrogen-passivated [001] GeC NWs studied in this work. The blue, gray, and
white spheres represent Ge, C and H atoms, respectively. The average length
between opposite hydrogens are labeled as a and b, which form a square whose
diagonal length is taken as the diameter of the NW.
2
For interpretation of color in Figs. 1, 2, 6, and 7, the reader is referred to the web
version of this article.
Fig. 2. Side and cross-section views of the initial structures of the hydrogen-
passivated [111] GeC NWs studied in this work. The blue, gray, and white spheres
represent Ge, C and H atoms, respectively.
48 F. Salazar, L.A. Prez / Computational Materials Science 63 (2012) 4751
(a) d = 6.18 , D = 5.20 eV, (b) d = 9.90 , D = 3.46 eV and (c)
d = 13.63 , D = 2.88 eV.
In Fig. 5a the band gap as a function of the diameter for Si, Ge,
SiC and GeC [0 0 1] NWs is shown. For the smallest diameter
considered, the SiC NW has the larger band gap which is very close
to that of the corresponding GeC NW, whereas the smallest band
gap correspond to the Ge NW. For larger diameters, the band gap
of the SiC NWs is slightly smaller than that of the GeC NWs, while
for Si and Ge NWs, the band gaps are very similar. Fig. 5b shows the
band gap as a function of the diameter for hydrogen-passivated
[1 1 1] Si, Ge, SiC and GeC NWs. In this case, the SiC and GeC band
gaps are always larger than those of Si and Ge NWs.
(b) (c) (d)
(a)
Fig. 3. Band structures of hydrogen-passivated [001] GeC NWs. The Fermi energy is
taken as zero (horizontal line). The diameters of the NWs are (a) d = 7.32 , (b)
d = 11.91 , (c) d = 16.40 , and (d) d = 20.88 .
(a) (b) (c)
Fig. 4. Band structures of hydrogen-passivated [111] GeC NWs with diameters (a)
d = 6.18 , (b) d = 9.90 , and (c) d = 13.63 . The Fermi energy is taken as zero
(horizontal line).
(a)
(b)
Fig. 5. (Color online) Band gap as a function of the diameter for (a) [001] and (b)
[111] nanowires. The band gap for bulk SiC, Si, GeC and Ge are indicated. Notice
that the packing of SiC and GeC NWs is higher than that of Si and Ge NWs, as shown
by the difference in diameters between wires with the same initial morphology.
Table 1
Growth direction, diameter (d), band gap (D) and type of band gap ([D] direct or [I] indirect) of the hydrogen-passivated NWs studied in this work.
Nanowire Growth direction d () D (eV) Type of band gap Growth direction d () D (eV) Type of band gap
Si [001] 8.79 4.429 D 7.19 3.774 I
14.08 2.571 D [111] 11.62 2.434 I
19.38 1.772 D 16.03 1.751 I
24.67 1.318 D
Ge [001] 9.09 4.110 I 7.43 3.598 I
14.63 2.565 I [111] 12.07 2.279 D
20.18 1.799 D 16.69 1.629 I
25.28 1.348 D
SiC [001] 7.14 4.914 D 5.98 5.259 D
11.44 3.312 D [111] 9.52 3.557 D
15.51 2.573 D 13.06 2.884 D
19.94 2.162 D
GeC [001] 7.32 4.860 D 6.18 5.201 D
11.91 3.312 D [111] 9.90 3.463 D
16.40 2.622 D 13.63 2.880 D
20.88 2.238 D
F. Salazar, L.A. Prez / Computational Materials Science 63 (2012) 4751 49
We also analyzed the type of band gap for the NWs studied and
the results are presented in Table 1, where the type of band gap,
direct (D) or indirect (I), for each NW is indicated together with
its diameter and band gap. Notice that, in contrast to the corre-
sponding bulk materials, the hydrogen-passivated SiC and GeC
NWs have a direct band gap for both growth directions and all
diameters considered. The same behavior has been obtained before
for SiC NWs by using accurate semi-empirical calculations [13]. On
the other hand, for passivated Si NWs, the nature of their band gap
depends on the growth direction, being direct for [001] NWs and
indirect for [111] NWs. Moreover, for large-diameter [001] passiv-
ated Ge NWs, the band gap is direct while for small-diameter ones,
the band gap is indirect. Likewise, for [111] passivated Ge NWs,
the nature of the band gap also depends on the diameter, being
indirect for both the smallest and largest diameter studied, and di-
rect for the intermediate one.
Finally, observe in Fig. 5 that, although the initial morphologies
were the same for GeC, SiC, Si, and Ge NWs in both [001] and
[111] directions, the diameters of the relaxed structures are quite
different between them. In particular, the SiC NWs in both direc-
tions have the most compact structure, followed by the GeC NWs
with slightly larger diameters. In contrast, the Ge NWs have the
largest diameters in both growth directions followed by the Si
NWs whose diameters are slightly smaller than those of Ge NWs.
This fact becomes important regarding their mechanical properties
which are studied in the next section.
4. Mechanical properties
Two representative elastic constants of nanowires are the
Youngs modulus, which is a measure of longitudinal elasticity,
and the Poissons ratio, which relates transverse to longitudinal
strain in response to a given stress. The Youngs modulus (Y) is
the ratio between the stress and strain and it can be calculated as:
Y
1
V
@
2
E
@e
2

e0
1
where E is the total energy, V is the equilibrium volume and e = DL/L
is the axial strain, being DL the change in the nanowires equilib-
rium length L. In general, the calculations consist of taking a section
of a nanowire and using periodic boundary conditions on a super-
cell geometry with sufcient lateral separation among neighboring
wires to simulate an isolated, innitely long wire. Then the nano-
wire is subjected to strains along the axial direction. At each strain
the positions of all atoms in the supercell are fully relaxed without
constraints, using a conjugate gradients minimization technique.
From these calculations the total energy and atomic positions, as
a function of the axial strain imposed on the nanowire, are obtained
from which one can calculate the Youngs modulus by tting a
second degree polynomial to E(e). The cross section of all studied
NWs is considered as a circumference whose diameter is obtained
following the procedure in Section 2.
Fig. 6a and b shows Y as a function of the radius for [001] and
[111] Si, Ge, SiC and GeC hydrogen-passivated NWs. Observe that
SiC (green squares) and GeC (open blue squares) NWs have a much
higher Y than Si and Ge NWs. This result can be explained if we
keep in mind that the carbide structures are more compact than
those corresponding to Si (black circles) and Ge (open red circles)
NWs as we saw in the previous Section, and then the unit cell vol-
ume of the carbide structures is smaller than that of Si and Ge
NWs. On the other hand, notice that the values of Y for passivated
Si and Ge NWs are below the bulk value for Si (doted black line)
and Ge (doted red line) respectively, and the tendency indicates
that in both cases they will converge to this limit value when the
radius is increased. In Fig. 6b, a similar behavior is observed for
[111] passivated nanowires. It is worth mentioning that the
results obtained in this work for Si NWs are in good agreement
with the DFT results obtained by B. Lee et al. [29], as shown in
Fig. 6a.
By comparing the results obtained for GeC NWs in both growth
directions, it can be understood the important role of the orienta-
tion of the bonds in the wire structure. Since the radius of the wires
grown along [111] direction is shorter than that of those grown
along [001] direction, the Youngs modulus along the former direc-
tion is higher than along the latter direction. Moreover, since the
covalent bonds are just along the [111] direction, the effect of
the strain along this direction is lesser than along [001] direction
where the strain is not parallel to the covalent bonds, and then
these NWs are more susceptible to be deformed. Finally, in all
cases, the results show that for both growth directions the SiC
NWs have a higher Youngs modulus than GeC NWs.
On the other hand, the Poissons ratio (m), which is the ratio of
transverse contraction strain (e
trans
) to longitudinal extension
strain (e) in the direction of stretching force, can be expressed as:
m
e
trans
e

1
e
R R
0
R
0
; 2
where R
0
is the equilibrium or zero-strain radius of the NW and R is
the radius at strain e.
The Poissons ratio was calculated with the same relaxed cong-
urations used to obtain the Youngs modulus. As we mentioned, all
NWs were relaxed for n different longitudinal strain forces. For
each strain (e
j
), a Poissons ratio m
0
j

was obtained and then we
averaged these n quantities, i.e.,
m
1
n
X
n
j1
m
0
j
3
5 8 11 14
0
100
200
300
3 5 7 9
0
120
240
360
Y
o
u
n
g
'
s

M
o
d
u
l
u
s

(
G
P
a
)
Radius
Si Bulk
Ge Bulk
Si
Ge
SiC
GeC
Y
o
u
n
g
'
s

M
o
d
u
l
u
s

(
G
P
a
)
Si Bulk
(a)
(b)
Radius
Fig. 6. (Color online) Youngs modulus as function of the radius for Si (black circles),
Ge (open red circles), SiC (green squares), and GeC (open blue squares) hydrogen-
passivated nanowires grown along (a) [001] and (b) [111] directions. The open
black squares in (a) correspond to the results of another DFT study [29] for Si NWs.
50 F. Salazar, L.A. Prez / Computational Materials Science 63 (2012) 4751
Fig. 7 shows m as a function of the radius for NWs grown along the
(a) [001] and (b) [111] directions. For all systems m diminishes
when the radius augments. Also, notice that the values of m for
GeC (open blue squares) are very close to those of SiC (solid green
squares). Moreover, Si (black circles) NWs have the higher m fol-
lowed by Ge (open red circles) NWs. Notice that the non-carbide
NWs have a higher value of m than the carbide NWs, in contrast with
what occurs for the Youngs modulus where it is clear that the car-
bide NWs have higher values.
5. Summary
In this paper we presented a theoretical study of the electronic
and mechanical properties of hydrogen-passivated GeC nanowires
grown along [001] and [111] crystallographic directions. As ex-
pected, the band gaps of the GeC NWs considered in this work
are larger than the bulk value as a consequence of the quantum
connement, and they converge to this value when the diameter
is increased. Moreover, the band gaps for both types of passivated
GeC NWs are direct and have almost the same value for similar
diameters, as occurs in the case of passivated SiC NWs. On the
other hand, for a given cross section geometry, the SiC and GeC
NWs have a smaller radius than the corresponding Si and Ge
NWs since the C atom has a smaller size. Furthermore, the Youngs
modulus of SiC NWs is higher than that of the GeC NWs, for both
studied crystallographic directions. In contrast, for the case of Si
and Ge NWs, their Youngs moduli are almost the same for all stud-
ied diameters. One important point is that the Youngs modulus
values show a clear difference between carbide and non-carbide
passivated NWs. Finally, the Poissons ratios of Si and Ge NWs
are higher than those of the GeC and SiC NWs, which are very close
between them.
Acknowledgements
This work was supported by DGAPA-UNAM under Grant
IN102511. Computations have been performed at KanBalam super-
computer of DGTIC-UNAM. F.S. acknowledges the UNAM postdoc-
toral fellowship.
References
[1] D.K. Ferry, Science 319 (5863) (2008) 579580.
[2] X. Duan, Y. Huang, J. Wang, C.M. Lieber, Nature 409 (66) (2001) 6669.
[3] C. Yang, C.J. Barrelet, F. Capasso, C.M. Lieber, Nano Letters 6 (12) (2006) 2929
2934.
[4] A.I. Boukai, Y. Bunimovich, J. Tahir-Kheli, J.-K. Yu, W.a. Goddard, J.R. Heath,
Nature 451 (7175) (2008) 168171.
[5] A.I. Hochbaum, R. Chen, R.D. Delgado, W. Liang, E.C. Garnett, M. Najarian, A.
Majumdar, P. Yang, Nature 451 (7175) (2008) 163167.
[6] D.D.D. Ma, C.S. Lee, F.C.K. Au, S.Y. Tong, S.T. Lee, Science (New York N.Y.) 299
(5614) (2003) 18741877.
[7] Y. Xia, P. Yang, Y. Sun, Y. Wu, B. Mayers, B. Gates, Y. Yin, F. Kim, H. Yan,
Advanced Materials 15 (5) (2003) 353389.
[8] L. Cademartiri, G.a. Ozin, Advanced Materials 21 (9) (2009) 10131020.
[9] R. Rurali, Reviews of Modern Physics 82 (1) (2010) 427449.
[10] R. Rurali, N. Lorente, Physical Review Letters 94 (2) (2005) 026805.
[11] J.-A. Yan, L. Yang, M.Y. Chou, Physical Review B 76 (11) (2007) 115319.
[12] P.W. Leu, B. Shan, K. Cho, Physical Review B 73 (19) (2006) 195320.
[13] A. Miranda, J. Cuevas, A. Ramos, M. Cruz-Irisson, Journal of Nano Research 5
(2009) 161167.
[14] D. Yao, G. Zhang, B. Li, Nano Letters 8 (12) (2008) 45574561.
[15] C.-C. Rhlig, M. Niebelschtz, K. Brueckner, K. Tonisch, O. Ambacher, V.
Cimalla, Physica Status Solidi (B) 247 (10) (2010) 25572570.
[16] G. Jing, H. Duan, X. Sun, Z. Zhang, J. Xu, Y. Li, J. Wang, D. Yu, Physical Review B
73 (23) (2006) 235409.
[17] E.P.S. Tan, Y. Zhu, T. Yu, L. Dai, C.H. Sow, V.B.C. Tan, C.T. Lim, Applied Physics
Letters 90 (16) (2007) 163112.
[18] K.H. Liu, W.L. Wang, Z. Xu, L. Liao, X.D. Bai, E.G. Wang, Applied Physics Letters
89 (22) (2006) 221908.
[19] Y. Zhu, F. Xu, Q. Qin, W.Y. Fung, W. Lu, Nano Letters 9 (11) (2009) 39343939.
[20] T.-Y. Zhang, M. Luo, W.K. Chan, Journal of Applied Physics 103 (10) (2008)
104308.
[21] Y. Chen, I. Stevenson, R. Pouy, L. Wang, D.N. McIlroy, T. Pounds, M.G. Norton,
D.E. Aston, Nanotechnology 18 (13) (2007) 135708.
[22] R. Pandey, M. Rerat, C. Darrigan, M. Causa, Journal of Applied Physics 88 (11)
(2000) 6462.
[23] A. Mahmood, L.E. Sansores, Journal of Materials Research 20 (05) (2005) 1101
1106.
[24] C.J. Kelly, J.S. Orr, H. G, Leonard T. T, A.H. Lettington, in: F. Yonezawa, T.
Ninomiya (Eds.), Application of Germanium Carbide in Durable Multilayer IR
Coatings, Proceedings of SPIE, vol. 122, 1990, p. 1275.
[25] J. Soler, E. Artacho, J. Gale, A. Garca, J. Junquera, P. Ordejn, D. Snchez-Portal,
Journal of Physics: Condensed Matter 14 (11) (2002) 2745.
[26] N. Troullier, J.L. Martins, Physical Review B 43 (3) (1991) 19932006.
[27] L. Kleinman, D. Bylander, Physical Review Letters 48 (20) (1982) 14251428.
[28] C. Kittel, Solid State Physics, 7th ed., 1996.
[29] B. Lee, R.E. Rudd, Physical Review B 75 (4) (2007) 041305(R).
(a)
(b)
Fig. 7. (Color online) Poissons ratio as function of the radius for Si (black circles),
Ge (open red circles), SiC (solid green squares), and GeC (open blue squares) NWs,
grown along the crystalographic directions (a) [001] and (b) [111].
F. Salazar, L.A. Prez / Computational Materials Science 63 (2012) 4751 51

You might also like