You are on page 1of 11

Early View publication on www.interscience.wiley.

com
(issue and page numbers not yet assigned;
citable using Digital Object Identifier DOI)

Laser & Photon. Rev., 111 (2008) / DOI 10.1002/lpor.200710032

Abstract We describe recent progress in photonic crystal


nanocavity lasers with an emphasis on our recent results on
ultrafast pulse generation. These lasers produce pulses on the
picosecond scale, corresponding to only hundreds of optical
cycles. We describe laser dynamics in optically pumped single
cavities and in coupled cavity arrays, at low and room temperature. Such ultrafast, efficient, and compact lasers show great
promise for applications in high-speed communications, information processing, and on-chip optical interconnects.

2008 by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Ultrafast photonic crystal lasers


Dirk Englund 1,* , Hatice Altug 2 , Bryan Ellis 1 , and Jelena Vuckovic 1
1
2

Ginzton Laboratory, Stanford University, Stanford CA 94305


Electrical and Computer Engineering Department, Boston University, Boston MA 02215

Received: 13 November 2007, Revised: 17 March 2008, Accepted: 13 May 2008


Published online: 2 July 2008

Key words: photonic crystal; laser; Purcell effect; ultrafast


PACS: 42.55.Sa,42.55.Tv,42.50.Ct, 42.70.Qs

1. Introduction
The field of photonics is transitioning towards highly integrated nanoscale devices. For the first time, researchers
are able to integrate fast low-power optical components on
semiconductor chips, causing some to draw analogies to the
semiconductor electronics revolution. The most commercially significant application of such integrated photonics
is optical communications. Because of low output powers, devices would first find applications in short-distance
communication, including high-speed local networks, and
board-to-board and chip-to-chip interconnects. Additional
applications lie in biochemical sensing and data storage.
One of the most promising architectures for integrated
nanoscale devices is the planar photonic crystal (PC). Inplane confinement is achieved by distributed Bragg reflection using periodic arrangements of holes, while out-ofplane confinement results from total internal reflection. Cavities, defined by defects in the PC, can confine light near
the ultimate volume limit /2n in all dimensions. With
optimized local geometry, extremely high confinement is
possible, with quality factors on the order of 104 in ac-

tive [13] and 106 in passive structures [4]. Through the


combination of a high quality factor (Q) and small mode
volume Vm , such cavities can dramatically increase the
vacuum Rabi energy, enabling cavity quantum electrodynamic effects such as enhanced spontaneous emission (SE)
rate of embedded emitters [5]. This cavity Purcell effect
lowers the lasing threshold through higher SE coupling
efficiency , far in excess of 50% for even modest mean
Purcell factor [3]. By contrast, in Vertical Cavity Surface Emitting Lasers (VCSELs) is typically less than 0.1%.
The Purcell effect can also increase the direct modulation
speed [6, 7]. The photonic on-chip design is ideally suited
for the integration of different optical components; e.g.,
light sources/detectors and multiplexers/demultiplexers for
optical communications.
In this paper, we describe recent progress on photonic
crystal nanocavity lasers. We will focus mainly on ultrafast
laser dynamics reported in recent work from our group,
but will attempt to mention the major works in the field
where appropriate. Previous studies investigated the lasing dynamics indirectly in the frequency domain [7]; here,
we instead focus on direct time-domain measurements. In

Corresponding author: e-mail: englund@stanford.edu

2008 by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Sect. 2.1, we begin with the designs for single-cavity and


coupled-cavity devices, then in Sect. 2.2 introduce a model
that describes lasing action in our devices. In Sect. 3, we
discuss our recent work on optically pumped PC lasers,
driven by multiple quantum wells (QWs). These structures
show extremely fast lasing action due to fast relaxation
dynamics. In the near-IR range, we demonstrate room temperature lasing with large-signal modulation response on
picosecond time-scales, i.e., with only several hundreds of
electric-field oscillations per pulse. A surface passivation
technique greatly improves the practicality of the PC laser
by limiting nonradiative (NR) losses. Using a similar design in the InP/InGaAsP material system, we demonstrate
large-signal modulation with pulse widths near 10ps in the
telecom band. In Sect. 4, we shift our attention from the
QW to quantum dot (QD) gain medium. In typical cavities
with Q 1000 and QD density 100/m2 , where threshold is determined by material properties, the QD active
material reduces lasing threshold because of lower gain
area and surface recombination losses. We conclude with a
brief discussion of electrical pumping.

2. Small-volume PC lasers
The small-volume, high-Q cavities enabled by photonic
crystals can decrease turn-on time and lasing threshold [6].
This improvement results when the gain spectrum overlaps
with the cavity resonance so that spontaneous emission
into the cavity mode exhibits higher spontaneous emission
rate and spontaneous emission coupling efficiency . These
effects can decrease turn-on time and lasing threshold [6].
In addition, microcavity lasers can be designed with very
broad modulation bandwidth because the relaxation oscillation can be shifted beyond the cavity cutoff frequency [8].
Above threshold, higher pump powers lead to faster
decay due to increased stimulated emission rates. Small
mode volume PC cavities can be used to achieve large photon densities and speed up this process. Compared to other
types of lasers such as VCSELs, PC lasers offer lower driving power (Sec. 3), higher relaxation oscillation frequency
(see Sec. 2.4), and potentially faster electrical modulation
speed because of the potential for lower device capacitance
and resistance.

2.1. PC nanocavity laser design


In designing the PC structure for fast lasing action, two
considerations are weighed: the Q value must be relatively
large to achieve SE Purcell enhancement and hence high
SE coupling efficiency; at the same time, the mode energy
ring-down time c = Q/ should be small as it limits
the lasers response time. We choose Q 2 103 , corresponding to = 1 ps at laser wavelength 1 m. This
value of Q is easily achieved with the the single-defect
cavity shown in Fig. 1c, defined in a square-lattice photonic

2008 by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

D. Englund, H. Altug, et al.: Ultrafast photonic crystal lasers

(a) Bz

(b) Bz

y
x

(c)

x
|E|2

(d)

Figure 1 (online color at: www.lpr-journal.org) Square-lattice


photonic crystal laser structures. (a) x-dipole-mode field pattern
(out-of-plane magnetic field Bz ). ydipole is rotated by 90 .
(b) Quadrupole mode. (c) Single-defect cavity with electric field
intensity (inset). (d) Coupled cavity array structure in GaAs.

crystal. The quadrupole mode, shown in the inset, has a


Q 2000 as predicted by Finite Difference Time Domain
(FDTD) simulations [9].
Because of its small size, the single-defect PC laser has
the disadvantage that output power is low on the order of
a few W and much of this power is lost due to a wideangle emission profile. On the other hand, band edge PC
lasers, which operate in slow-group velocity regions of the
PC dispersion, comprise a greater gain area, as was shown
in the first PC laser demonstrations [10, 11]. They also offer greatly improved emission directionality [12]. However,
they entail other drawbacks, such as reduced lateral confinement [13]. A good compromise appears to be combining
the strengths of the nanocavity and band edge lasers by arranging single-cavity lasers into an array [9]. If the cavities
are sufficiently close, lasing can be achieved in a common
mode. This PC nanocavity array laser has far more directional emission and larger active material than the single
laser, while providing better lateral confinement than the
band edge laser. The nanocavity array achieves very low
group velocity in any photonic crystal direction and a very
high density of electromagnetic states; in effect, the structure is the two-dimensional analog of coupled resonator
optical waveguides (CROWs) in photonic crystals [14, 15].
Though coupled arrays of small numbers of VCSELs were
previously investigated [16], coupling between individual
lasers is difficult and requires a rather complicated fabrication procedure. Photonic crystal nanocavity arrays allow
precise control of both the uniformity and the coupling.

www.lpr-journal.org

Laser & Photon. Rev. (2008)

(a)

(b)

Figure 2 (online color at: www.lpr-journal.org) Three-hole defect cavity. (a) FDTD design with electric field intensity. (b) Fabricated structure in GaAs with central InAs QD layer.

We investigated two-dimensional cavity arrays with


modes that couple equally in the different crystal directions
(monopole and quadrupole modes in the square lattice or
monopole and hexapole modes in the triangular lattice). We
primarily consider PC cavity array structures in a square
lattice [9]; this choice was arbitrary and we would expect
similar results in the triangular lattice. The coupled cavity array and its dipole and quadrupole field patterns are
shown in Fig. 1a,b. In these modes, the in-plane electric
field components Ex and Ey , as well as out-of-plane Bz ,
are maximized in the center of the slab. This is commonly
called the transverse electric (TE)-like mode.
For QD lasers, we explored higher Purcell factors with a
three-hole defect cavity [17]. Though theoretically limited
to Q 120, 000 in this design, the fabricated structure
shown in Fig. 2 has Q 3000, reduced by fabrication
imperfections and material loss [18].

2.2. Rate equations


A simple rate equations model describes the lasing dynamics well. Material gain is averaged over the full mode as the
QW or QD layers span the full structure. The mode holds
p photons in a volume Vm . The laser dynamics are modeled with three carrier levels: the excitonic ground state,
the pump level carrier number nE (populated above the
GaAs-bandgap using a laser with power Lin ), and the QW
lasing level carrier number nG (resonant with the lasing
mode frequency). We then have [19]
dp
Fm nG
p
= g(nG )p +

(1)
dt
r
p


dnG
nE
Fm + FP C
1
=
nG
+
g(nG )p
dt
E,f
r
P C,nr


dnE
Lin
1
1
1
=
nE
+
+
dt
~p
E,r
E,nr
E,f
In the top equation, the cavity photon number is driven
by the QW through stimulated emission (gain term g(nG )p
see [19]) and SE (at the resonant modes Purcell-enhanced
rate Fm /r ). The cavity loses photons at the cavity loss
rate 1/p . The carrier number nG in the center equation is
pumped by carrier relaxation from the pump level population nE at rate 1/E,f . Besides pumping the cavity, nG

www.lpr-journal.org

decays through NR channels at rate 1/P C,nr and PC leaky


modes at rate FP C /r , where FP C 0.2 expresses SE
rate quenching inside the PC bandgap compared to the SE
rate 1/r in the bulk QW (following simulations in [5]).
In the bottom equation, the nE level is pumped through
above-band optical excitation with power Lin at frequency
p (the first term) and decays through carrier relaxation to
nG , NR recombination, and SE (second term).
In the following text, we will use these rate equations
to model the lasing action of single and coupled PC lasers,
at both room and low temperature ( 10K) and containing
QWs or QDs as gain material. We use a logarithmic gain
model for QWs and a linear gain model for QDs. Under
high pump power, the rate equations model would require
modification to account for QD saturation [20].

2.3. Threshold
Solving Eqs. (1) in steady state gives the lasing threshold
power, defined here as the power where the average photon
number p = 1 [8]. Assuming that most pump-level population drops into the lasing level (E,f  E,r , E,nr ), the
threshold is given by




~p
p
p
Lin,th =
nG,th FP C +
+1 .
p
r
P C,nr
In our QW- and QD-driven cavities, the threshold carrier
number in the active volume Va is approximated using the
materials transparency concentration, nG,th Ntr Va
1018 cm3 Vm , where the gain confinement factor
0.16 approximates the cavity mode overlap with the QW
region. Furthermore, with p 1 ps, r 600 ps, and
P C,nr 100 ps, it is easy to see that for our laser structures
the first term in the brackets dominates, giving


~p
FP C
1
Lin,th
Va Nth
+
(2)

r
P C,nr
The threshold is thus determined by the gain materials
transparency concentration Ntr , radiative loss into non-laser
modes, and nonradiative recombination. In the QD-driven
devices, nonradiative recombination is reduced and the
term FP C /r dominates. The factor FP C indicates that
threshold is reduced by suppression of SE into non-lasing
modes [5]. On the other hand, in QW-driven devices, the
nonradiative term 1/P C,nr determines threshold. Although
Eq. (2.3) was derived for steady-state, we find that it is also
a good approximation for pulsed excitation.

2.4. Laser intensity modulation


Two modulation schemes are used in telecommunications:
small-signal and large-signal modulation [21, 22]. In smallsignal modulation, the laser is driven at a constant abovethreshold pump power Lin,0 and modulated with a small signal Lin , resulting in differential changes P = (p/Vm )

2008 by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

D. Englund, H. Altug, et al.: Ultrafast photonic crystal lasers

and NG = (nG /Va ) to the steady-state photon density


P0 and lasing-level carrier densities NG,0 . At low driving
power above threshold, the differential output power P
about the steady-state output P0 is limited by the relaxation
oscillation frequency [6],
2
R
=

avg P0

N0
+
+
p
p r /Fm
P0 P C,nr r /Fm

(3)

Here = Fm /(FP C + Fm ), as described in [5], a is the


differential gain, and vg the group velocity. In conventional
lasers, only the first term in Eq. (3) is considered as is
small [21]. Therefore, to increase bandwidth, P0 and hence
the driving power is raised. The higher power can result
in thermal problems [23], though injection locking may
help in VCSELs [24]. Eq. (3) shows that in the high-
laser, strong cavity effects help to increase R without the
need to increase pump power, opening a new pathway for
increasing laser modulation bandwidth [8].
In large-signal modulation, the rate equations predict
that the modulation rate is limited by the pump-level relaxation time E,f and cavity response time p = Q/.
An additional turn-on delay arises as spontaneous emission
builds the cavity field to the point when stimulated emission
becomes dominant. This delay time is reduced in the highPurcell regime through faster SE rate and higher . This
is seen from the turn-on behavior for laser cavities with
different Q in Fig. 3. Here the Purcell factor is calculated as
Fm =

3
Q
,
4 2 Vm /(/n)3

(4)

where the factor accounts for spatial averaging and is estimated at 0.18 from the measured Purcell rate enhancement of the coupled cavity array in Sect. 3.2. The figure
shows that as Fm is increased, the turn-on delay asymptotically decreases to a value determined by the carrier
relaxation time and the pump power. The delay decreases

Pout(mW)
0
0
0
0

pump Pin

10

Q=1600, Fm= 81

10

Q=800, Fm= 41

10

Q=400, Fm= 20

10

0
0
0

with pulse energy, which is set here to excite a carrier concentration of 3 Ntr . As Q is increased from 100 to 1600,
the lasing duration first decreases with faster SE rate, then
extends as it approaches the cavity ring-down time. Depending on the driving conditions, the modulation rate can
be optimized with a Q that provides high Purcell factor but
does not excessively slow the cavity response.

2.5. Rate equations model in FDTD


The three-level rate equations model does not account for
spatial variations in the carrier concentration across the photonic crystal device. However, we have found that spatial
effects such as carrier transport from the pump spot to the
gain region, or spatial hole burning effects [25], are important in understanding lasing efficiency and time response.
For that reason, we have developed a finite-difference time
domain model that includes carrier dynamics.
Material gain is implemented in FDTD by an effective
conductivity , as in references [26, 27]. An auxiliary differential equation is used to describe the evolution of the
current density J. In turn, J is related to the carrier density
NG (assumed to be equal for holes and electrons in the
intrinsic semiconductors considered here). The set of equations obtained when J = E is substituted into Maxwells
equations and is then expressed in the time-domain and
discretized, as described in [28]. The resulting nonlinear
FDTD model allows calculation of the carrier drift into the
lasing mode, and is important for explaining the laser time
response measurements covered in Sect. 3.

3. Quantum well photonic crystal lasers


Quantum wells provide large gain when embedded in the
center of the PC membrane, where the resonant TE-like
mode has the maximum electric field energy density. A
single QW in the PC slab center would see the highest
electric field and hence the highest gain overlap; however,
to optimize the laser current, it is often better to distribute
carriers (or current) across several quantum wells [29]. A
QW-driven PC nanocavity laser was first demonstrated with
four InGaAsP quantum wells [30] and was soon followed
by other demonstrations employing between three and six
quantum wells [31, 32], all operating in the telecommunications band.

Q=200, Fm= 10

10

Q=100, Fm= 5

3 t(ps) 4

Figure 3 (online color at: www.lpr-journal.org) Calculated lasing


power P (t) (Vm ~G /p ) in response to a 3-ps pump pulse (top),
for a range of Q. The turn-on delay drops with increasing Q. The
excitation carrier density is 3Ntr per pulse for all plots, and pump
efficiency = 1 in this idealized model.

2008 by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

3.1. GaAs/InGaAs structures


We first investigated time-domain characteristics of PC
nanocavity lasers using a streak camera with a Hamamatsu N5716-03 streak tube. Since the detector response
is limited to wavelengths below 1 m, we fabricated PC
lasers emitting between 900980 nm. These employ four
8-nm In0.2 Ga0.8 As QWs separated by 8-nm GaAs barriers (see illustration in Fig. 4). The top and bottom QWs

www.lpr-journal.org

cryostat

12

streak camera

8
4

400

HWP

PBS

500

600

700

800

900

1000

spectrometer
or OSA

(b) 10K,pulsed

(a) 10K,pulsed

Q=1520 coupled cavity

930

950
L (nm)

intensity (a.u.)

Ti:Saph
pump
pulses

counts/s 10 4

MQWs

15

coupled cavity

are about 32 nm from the center and still see 89% of the
central maximum field intensity. We use compressively
strained QWs which have higher differential gain, lower
transparency carrier density Ntr , and higher coupling to the
TE-like-polarized cavity mode than unstrained QWs [21].
We first consider lasers consisting of 172 nm-thick GaAs
slabs patterned with 9 9 arrays of coupled PC cavities
in a square-lattice PC (Fig. 1d). The structures are fabricated by electron beam lithography in polymethyl methacrylate (PMMA), followed by plasma-etch mask transfer and
wet-etch removal of a sacrificial layer beneath the membrane. To reduce nonradiative (NR) surface recombination
on the large QW area exposed through PC patterning, the
sample was passivated in a (NH4 )S solution, which resulted in a 3.7-fold reduction in the lasing threshold [33].
We found that surface passivation was critical in our samples for room-temperature and continuous-wave (CW) lowtemperature operation.
The structures are pumped optically with 3-ps short
pulses at an 80 MHz repetition rate and a wavelength centered at 750 nm using the confocal microscope as described
in [34] and shown in Fig. 4. High-resolution lasing spectra are measured with the spectrometer, while time response is obtained using a streak camera with 3-ps resolution. At room temperature, the photoluminescence of the
In0.2 Ga0.8 As quantum wells peaks at 980 nm. For higher
gain and heat dissipation, we first evaluated cooled structures [6].
The PC array laser in Fig. 1d supports a lasing mode at
mode = 950 nm at low temperature (LT) of 10K (Fig. 5a).
Because of fabrication imperfections, PC holes near the
edges of the structure were slightly smaller and cavities

www.lpr-journal.org

80
60
40
20

counts/s 10 4

Figure 4 (online color at: www.lpr-journal.org) Confocal microscope setup (lens numerical aperture=0.65). The laser structures
are mounted in a cryostat, which is cooled for some measurements.
Emission is directed to either the streak camera, spectrometer with
cooled Si detector, or optical spectrum analyzer for IR spectra.
Inset: Illustration of coupled cavity array membrane.

25
15

L in (MW)

substrate

10

20

30

(d) 293K,pulsed

model
coupled cavity

L in (MW)

100 200 300 400

15

(e) 10K,CW
model

10

L in (MW)

model

unpassivated

0
0

970

(c) 10K,pulsed

25 .

passivated

counts/s 10 2

objective
lens

Intensity (1000 counts)

Laser & Photon. Rev. (2008)

cts.10 (f)10K, CW (struct 2)


14
10

coupled cavity

model
single cavity

6
L in(MW)

10

30

50

L in(MW)

10

15

Figure 5 (online color at: www.lpr-journal.org) QW-driven PC


lasing characteristics (passivated structures). (a) Coupled-cavity
array spectrum below threshold and at low temperature (10K).
The lasing mode consists of an estimated 79 cavities. (b) Lowtemperature lasing curve shows threshold reduction after passivation. (c,d) Low-and room-temperature lasing curves with pulsed
excitation (3.5-pulses at 80 MHz repetition, passivated structure).
(e,f) Continuous excitation lasing curves for coupled and single
cavity. Horizontal axes show average pump power. The fits are
by Eqs. (1).

showed a higher resonance wavelength. As a result, we


observed that coupled cavity modes existed only in a subset of the full array. From optical microscope images, we
estimate that the lasing mode comprises only 79 cavities;
the pump beam diameter was adjusted to this size. Fig. 5c
shows the lasing curve for pulsed excitation (3.5 ps at 13 ns
repetition), with an averaged threshold of 6.5 W(measured
in front of the objective lens). This corresponds to a large
peak pump power of 21mW.
The threshold power is much lower under continuous
pumping at low temperature. Fig. 5e displays the lasing
curve of the passivated structure, indicating onset of lasing
at only 9 W. For a single cavity, threshold is even lower,
near 2 W, shown in Fig. 5f. This threshold and a similarly
low value recently reported with GaInAsP/InP QWs [35]
are lower than in previous low-threshold QW lasers [36,37].
We believe that three main factors reduce threshold in
CW operation. First, carrier radiative efficiency is higher in
steady-state lasing as stimulated emission outpaces other
nonradiative recombination processes, which are more sig-

2008 by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

D. Englund, H. Altug, et al.: Ultrafast photonic crystal lasers

counts (normalized)

concentration (1/cm3)

RT

experiment
theory

LT

(a)

pump

t(ps)
0

10

. 10
10
8
6
4
2

17

20

30

40

50

pump

NE

60

(b)
NG

laser response:
P(t) . 5
efficient carrier
conversion
threshold
t(ps)

10

20

30

40

Figure 6 (online color at: www.lpr-journal.org) Laser time response. (a) Experimental data shows response nearly following
the excitation pulse at room temperature; data at both temperatures
are acquired at 2 lasing threshold. (b) Illustration of pump inefficiency in pulsed operation. Pump energy is efficiently channeled
into the cavity mode only during lasing (shaded area under P (t)
curve, amplified here 5 for visibility); much of the remaining
pump energy is wasted to SE and NR losses.

nificant in pulsed operation (see illustration in Fig. 6b).


Second, for the same average pump power, the peak power
of the pulsed beam is thousands of times larger and results
in a higher temperature of photoexcited carriers. The faster
diffusion of the high-temperature carriers results in a larger
effective pump spot (observed in photoluminescence) with
lower gain overlap. Third, we estimate that CW operation
is made even more efficient by carrier drift into cavity. This
drift results from a carrier density gradient caused by spatial
hole burning in the cavity mode (see Fig. 10a). It is expected
to be insignificant for the higher-temperature pumping in
pulsed operation [34].
These contributions are quantified by applying the rate
model of Eqs. (1)(see fits in Fig. 5). All recombination rates
are estimated from time-resolved measurements on nonlasing structures. The model indicates that pump CW absorption efficiency 0.055 is far better than in pulsed
operation, where = 1.3 103 [34]. The comparison
of CW and pulsed excitation regimes indicates that there
is significant room for improving pumping efficiency in
pulsed mode at low temperature.
At room temperature (RT), threshold is higher. The lasing curve in Fig. 5d indicates a lasing threshold of 68 W
average power. The larger threshold results in part from a
higher transparency concentration, smaller optical gain [21],
and larger NR surface recombination rate [33]. These ef-

2008 by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

fects are furthermore exaggerated by heating due to higher


threshold pump power. Because of larger thermal velocity
and diffusion, the above-mentioned carrier drift into the
lasing cavity will be reduced. The larger threshold causes
heating in the suspended membrane structures that limits
the maximum output power, as can be seen in the falloff in Fig. 5d at 350 W. Because of this heating, we
achieved only quasi-CW operation at RT. This required a
chopper wheel that provided 1 ms-long pulses at a 17 Hz
repetition rate. Heat dissipation can be greatly improved
in RT-operation by fabricating the PC laser structures on
top of low-index substrates such as sapphire or silicon oxide [31, 32, 3840], or by replacing QWs with QDs which
have lower nonradiative loss and carrier transparency [41].
We have also found that capping the photonic crystal membrane in PMMA improves heat dissipation by up to 20,
based on measurements of the maximum pump power before the structure is damaged. The capping method also
helps prevent re-oxidation of passivated structures.
Because of faster carrier dynamics, RT operation results
in faster modulation speed. This is seen in Fig. 6a comparing RT and LT lasing response to 3.4-ps-long pump pulses
(13 ns repetition). Both measurements were obtained with
pump powers roughly 2 above threshold, corresponding
to averaged pump powers of 13 W and 136 W at low-and
room temperature, respectively. We measured significantly
faster lasing response at room temperature, with the lasing
pulses roughly following the 3.4-ps pump duration. Frequency chirp was less than the cavity linewidth up to 2
threshold pump power.
The speed-up results primarily because the intraband
relaxation time is shorter at RT; we measured E,f < 1 ps
[34], which agrees with previous reports for III-V quantum wells [25, 42, 43]. This behavior is captured well by
the three-level rate equations model (fits in Fig. 6a) whose
calculated response is convolved with a filter that takes into
account the 3-ps response time (FWHM) of the streak camera [33]. Based on the model, lasing response approaches
FWHM = 1.2 ps at 2 threshold pump power when
pumped with shorter 1-ps laser pulses, implying that modulation rates in the THz regime would be possible. The delay
can be decreased with higher pump power, but is ultimately
limited by the carrier relaxation time E,f .

3.2. Spontaneous emission rate modification


We measure the Purcell factor Fm directly from lifetime
measurements of the cavity array pumped below threshold, compared to emission lifetime in the unpatterned
(bulk) sample [6]. Decay times for the passivated cavity array structure are estimated from Fig. 7b,c, indicating
uncoupled 142 ps and coupled 19 ps. Then the center
equation of (1) is used to calculate the underlying recombination rates (with p = 0). For the bulk and PC regions
at times far after the excitation pulse (when nE 0), the
laser level carrier number nG and its measured photolumi-

www.lpr-journal.org

intensity

Laser & Photon. Rev. (2008)

(a) bulk

passivated, !=605ps

0.8

(a)

0.6

th
ng
e
l
ve
wa

unpassivated, !=618ps

0.4
0.2

1
0.8

0.2

intensity

2.0
1.5
1.0

unpassivated
""""""""""!=34 ps

0.4

25

50

75

100

125

150

175

0.4
0.2

t(ps)
10

20

30

40

50

Figure 7 (online color at: www.lpr-journal.org) Microphotoluminescence from bulk quantum well, PC (uncoupled to cavity array
mode), and non-lasing PC cavity array at 1/2 threshold power
(Pin = 12 W before objective lens for original and 12 W for
passivated structures, pulse length 3.5 ps with 80 MHz repetition).
Measurements at 10K. Solid fits are by Eqs. (1); dashed fits show
exponential decay approximations.

nescence signal decay according to


1
coupled

Fm + FP C
1
+
r
P C,nr

1
FP C
1
=
+
uncoupled
r
P C,nr
1
bulk

(5)

1
1
+
r
bulk,nr

From bulk measurements, we estimate the natural radiative lifetime r 605 ps, assuming bulk,nr  r . Eqs. (5)
then give Fm 28. Repeating these measurements for an
unpassivated single-defect cavity gives a spatially averaged
Fm 81 [6]. The high Purcell factor for single cavities
is not surprising as they are expected to have a maximum

www.lpr-journal.org

10 15 20 25
Time (ps)

Figure 8 (online color at: www.lpr-journal.org) Large-signal


lasing response in QW-driven PC laser. (a) Response to excitation
pulses at (i) 9 0.5 and (ii) 15 ps. (b) Excitation pulse train
created by etalon setup. Imperfect mirror arrangement causes
an exponential decrease in pulse power and only the first three
pulses exceed the photonic crystal lasing threshold. (c) Lasing
response delay.

passivated
"""""!=19ps

0.6

t(ps)
200

(c) cavity

0.8

T delay ~ 1.5 ps

0.5
00

2.5

passivated
""""""""""!=142ps

0.6

(c)

(b)

(b) PC, uncoupled

PhC laser

t(ps)
750 1000 1250 1500 1750 2000

500

Intensity (a.u.)

250

Pump laser

intensity

F of 165 for the cavity with this set of Q and Vm [5], indicating that spatial averaging over the mode reduced Fm
by 2. Baba et al. previously estimated SE lifetime enhancement exceeding 16 (detector response limited) for
similar structures in GaInAsP PC nanocavities [25].

3.3. Delay time


As we indicated above, an important parameter in the largesignal modulation scheme is the delay time, which decreases in high Purcell-factor cavities. We measured the delay time at 100K (with 890-nm pump wavelength) as 1.5 ps
(Fig. 8c). This delay time is nearly two orders of magnitude
shorter than in previous measurements for VCSELs [44].

3.4. Large-signal modulation


To further demonstrate high-speed characteristics, we directly modulate single-defect cavity lasers at high speeds
by pumping with a series of 170-fs pulses generated using
a Fabry-Perot etalon [6]. Fig. 8a,b shows the results for
direct modulation of a nanocavity at low temperature. This
measurement shows that in principle, large-signal modulation well in excess of 100 GHz is indeed possible. Faster
operation at room temperature is expected, but the etalon
measurements were not repeated in the passivated structure.

2008 by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

D. Englund, H. Altug, et al.: Ultrafast photonic crystal lasers

0.4
0.2

0.6
0.4
0.2
0

single cavity

(b)

L in (#W)
20

40

60

"!=0.25nm

"5

(a)

carrier concentration

NG

0.5

0
1

coupled cavity
array

(c)

2.2
1.8
1.4

1510

1520

!(nm)

1530

carrier concentration

"4

0.5

0.6
0

1500

(b)

18

x 10

!y /a

1
0.8

Intensity (normalized)

0.6

(a)

Intensity (normd)

0.8

Intensity (normalized)

!y /a

-20 -10

0.2

10

20

30

40

t(ps)

5
"5

Quantum wells have a major drawback as gain media in


photonic crystal devices. Nonradiative surface recombination rate is very large because the QW has a large surface
area exposed to air at the hole walls. We have reduced the
NR recombination problem, for the first time in a PC laser
structure, by applying a surface passivation treatment [33].
The (NH4 )S-mediated treatment reduced the NR recombination rate by more than four times (Fig. 7b) and led
to a fourfold reduction of lasing threshold, as shown in
Fig. 5(b). The increased efficiency extends the operating
range from cryogenic to practical regimes, enabling the
above-mentioned room-temperature and ultra-low threshold CW operation.

3.6. Telecom wavelength laser


Lasers operating near 1500 nm are particularly interesting
for applications in optical telecommunications [3032]. For
this reason, we investigated the time-domain characteristics
of PC nanocavity lasers in InP with InGaAsP quantum
well gain [45, 46]. The structure was tested with the setup
in Fig. 4 (configured for 1550 nm with detector N571602). The coupled cavity designs are identical to the GaAs
structures presented earlier, though scaled up for the longer
operating wavelength.
The single-cavity InP structure is shown in Fig. 10c. Its
lasing behavior is shown in Fig. 9a and indicates a threshold of 22 W when pumped with 3-ps above-band (750 nm)
excitation pulses. We measure large-signal response with
FWHM 10 ps, at 2 above threshold.
In a 9 9 photonic crystal cavity array, we measured
FWHM 19 ps, also at 2 above threshold. We believe
that the longer pulse duration results from inhomogeneity
in the pump beam, leading to different gain in different cavities.
To analyze the pulse duration more closely, we modeled
the lasing action using the nonlinear FDTD simulations

2008 by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

"4

(c) SEM

!x/a

!x/a

(d) Bz
"4

!y /a

Figure 9 (online color at: www.lpr-journal.org) PC laser in InP


with InGaAsP multiple QWs. (a) Lasing mode (2 threshold)
and light-in, light-out curve (inset) for single-cavity structure.
(b) Single-cavity lasing response (Lin = 45 W) (c) Coupledcavity array time response (Lin = 2 threshold).

3.5. Surface passivation

!x/a

2"m

4
"4

Figure 10 (online color at: www.lpr-journal.org) (a) Lasinglevel carrier concentration (1 ps after injection) showing density
gradient towards lasing cavity. Spatial hole burning results from
the fast stimulated recombination during the lasing pulse. Pump
power is 2 Ntr at the center of the gaussian spot with radius
2a (lattice periods). (b) Carrier concentration in PC array, 1 ps
after injection. Just beyond threshold, small inhomogeneities in
the pump spot (radius 6a) and coupled cavity mode can result in a
spreading of lasing onset times, contributing to longer total pulse
duration. (c) SEM of single-cavity InP laser structure. (d) Out-ofplane magnetic field of lasing mode, 1 ps after carrier injection.

discussed in Sect. 2.5. Optical pumping with a finite-sized


beam results in inhomogeneous gain and uneven lasing
action, spreading out the total pulse duration. This is seen
from the lasing level concentration NG and lasing fields
in Fig. 10b,d, recorded here 1 ps after injecting carriers
at a two-fold transparency concentration. The cavities are
in different stages of lasing action. As a result, the total
response time is extended. Experimentally we find that at
higher powers, the pulse response becomes shorter. This
observation supports our model, as all cavities would be
further above threshold and lase in a closer timeframe. This
result shows that phase-locking on the ultrafast time scale
requires homogeneous pumping across the array.

4. Quantum dot photonic crystal lasers


We now consider PC lasers that use QDs for gain. These
permit lower threshold due to lower carrier transparency
and nonradiative surface recombination, and greater temperature stability and differential gain. Their speed is set
by the smaller of relaxation rate 1/E,f and relaxation
oscillation rate R . In high-Q photonic crystal nanocavities, the lasing response is sped up through the Purcell

www.lpr-journal.org

Laser & Photon. Rev. (2008)

pump

* experiment
theory

Intensity (a.u.)

Intensity (a.u.)

t(ps)

response

t(ps)

L in (!W)
(c)

(b)

Intensity (a.u.)

Intensity (a.u.)

(a)

(d)

t(ps)

Figure 11 (online color at: www.lpr-journal.org) GaAs PC laser


with InAs QD gain. (a) Measured lasing curve and fit by Eq. (1).
(b) The measured turn-on delay between pump (first peak) and
laser response (second peak) is limited by carrier relaxation.
(c) Measured large-signal modulation speed increases with pump
power. (d) Corresponding fit by rate equations.

effect, while large and thus higher efficiency and lower


threshold are achieved. We investigated these aspects in
the 135 nm thick GaAs PC membrane shown in Fig. 2(a),
containing a high-density (600 m2 ) of InAs QDs. These
self-assembled dots have shallow confinement and operate
only at cryogenic temperature with an emission wavelength
of 940 nm and inhomogeneous linewidth of 20 nm. Lowerenergy QDs allow laser operation at room temperature in
pulsed [47] and CW modes [41].
The y-polarized fundamental mode (Fig. 2(b)) is resonant in the structure near 920 nm, with cold-cavity Q
3000(p 1.5 ps). We measure a gradual onset of lasing
near 1 W, as shown in Fig. 11a. From fits to the lasing
curve, we estimate a SE coupling factor 0.2. Streak
camera measurements of the rise time of photoluminescence from quantum dots in bulk GaAs indicate that the carrier relaxation time E,f 10 ps for a wide range of pump
powers. We also find that resonant pumping of higher-order
confined states of the QDs (such as p-level states) does
not appreciably lower E,f . Because the carrier capture
time is longer than the cavity photon lifetime, it ultimately
determines the maximum modulation bandwidth. This is
what we observe in Fig. 11b which shows a delay of 13.5 ps
(at five times threshold) and does not drop below 12 ps for
higher powers. Simulations with Eqs. (1) support this observation as rise time is limited by the carrier capture time.
In our cavity-QED-enhanced structure, the relaxation-time
limit is rapidly reached in the high- case. In contrast, in
non-PC quantum dot lasers not employing strong cavity
effects, far higher pump power is needed to reach this limit.
Once lasing is reached, stimulated emission causes fast
carrier recombination. We measured a decay time of 8.5 ps
at pump powers around five times the threshold (Fig. 11c).

www.lpr-journal.org

For higher pump powers the laser response appears largely


unchanged, presumably due to carrier saturation. We again
model the system with Eqs. (1), employing a linear gain
model and parameters given in [48]. Fig. 11d shows the
simulated laser response at various pump powers, demonstrating good agreement between theory and experiment.
The present work predicts that large-signal modulation
in present PC lasers employing conventional self-assembled
In(Ga)As QDs is limited to 30 GHz due to relaxation
dynamics. While there is also evidence to suggest that
carrier relaxation and hence maximum modulation rate
actually further slows at increased temperature [49], recent advances in QD growth can open the way to higher
performance. QDs driven through phonon-assisted tunneling show very short relaxation time, with E,f 2 ps at
room temperature [50], and were recently demonstrated in
ridge waveguide lasers with 25 GHz small-signal modulation bandwidth [50]. This bandwidth may be significantly
improved using a PC laser cavity. In addition, reduction in
the inhomogenous linewidth broadening and reduction in
hot-carrier effects and associated gain compression [51],
will improve PC QD laser efficiency and speed. P-type
doping of quantum dots also promises to speed up carrier dynamics [52].

5. Conclusions and future directions


Photonic crystal lasers provide unprecedented speed, reaching pulses on picosecond scales. They also show very
low threshold, lasing at only several microwatts of pump
power. Their planar design makes them ideal on-chip integration. But for practical applications, PC lasers will
need to be pumped electrically. Many groups are currently
pursuing this goal, and electrically driven single-cavity
PC lasers have been demonstrated in free-standing membranes [53, 54] and band-edge laser structures [55]. For
high-speed electrical modulation in extended structures
such as band edge and nanocavity array lasers, it will be
important that the structure be uniformly pumped, as the
spatial modeling of carrier dynamics in Sect. 3.6 suggests.
A further challenge for any PC laser will be keeping RC
time constants small, where C and R are the capacitance
and resistance of the laser. Compared to VCSELs, PC lasers
promise far lower capacitance due to small a footprint and
lower resistance because of thin intrinsic material between
electrodes. A promising recent step demonstrated time constants below 10 ps using micron-scale contacts with sub-fF
capacitance [56]. With recent advances in integration, electrical pumping, and ultrafast operation, PC crystal lasers
promise to fill a growing need for integrated, ultrafast optical communication.
Acknowledgements This work was supported by the MARCO
IFC Center, NSF Grants ECS-0424080 and ECS-0421483, the
MURI Center (ARO/DTO Program No. DAAD19-03-1-0199), as
well as NDSEG & NSF Fellowships (D.E.) and Stanford Graduate
Fellowship (B.E.).

2008 by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

10

Dirk Englund is a graduate student


in applied physics at Stanford University. He received his Bachelors
of Science degree in physics from
the California Institute of Technology in 2002. He is a recipient of
the NSF, NDSEG, Stanford Graduate
Mayfield, and U.S. Fulbright fellowships. His research focuses on quantum photonic devices.
Hatice Altug received the B.S. degree in Physics from Bilkent University, Turkey in 2000. She received
the M.S. and Ph.D. degree in electrical engineering and applied physics
from Stanford University in 2006.
Currently, she is a Peter Paul Career
Development Professor in Electrical
and Computer Engineering Department at Boston University. Her research involves design and implementation of high performance and ultracompact nano-photonic devices and sensors including
lasers and all-photonic switches and their large-scale onchip integration for communication and bio-sensing applications.
Bryan Ellis was born in Denver, Colorado in 1983. He received the B.S.E
degree in electrical engineering from
Princeton University in 2005. He is
currently working towards a Ph.D.
degree in electrical engineering from
Stanford University. His research interests include nanophotonic devices
employing optical microcavities for
use in optical communications and optical interconnect technologies.
Jelena Vuckovic received the PhD degree from Caltech in 2002, and has
been working at Stanford University
as a faculty since 2003. Her research
focuses on nano- and quantum photonic devices and circuits. She is an
author of more than 60 publications
in refereed journals, more than 70
invited and plenary talks, five book
chapters, five issued and several pending U.S. patents,
and a recipient of numerous awards, including the Office of Naval Research Young Investigator Award and
the Frederick Terman Fellowship, given to the most
promising young faculty in sciences and engineering
at Stanford.

2008 by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

D. Englund, H. Altug, et al.: Ultrafast photonic crystal lasers

References
[1] K. Nozaki, S. Kita, and T. Baba, Opt. Express 15(12), 7506
7514 (2007).
[2] D. Englund, A. Faraon, I. Fushman, N. Stoltz, P. Petroff,
and J. Vuckovic, Nature 450(6), 85761 (2007).
[3] S. Strauf, K. Hennessy, M. T. Rakher, Y. S. Choi,
A. Badolato, L. C. Andreani, E. L. Hu, P. M. Petroff, and
D. Bouwmeester, Physical Review Letters 96(12), 127404
(2006).
[4] B. S. Song, S. Noda, T. Asano, and Y. Akahane, Nature
Materials 4, 207 210 (2005).
[5] D. Englund, D. Fattal, E. Waks, G. Solomon, B. Zhang,
T. Nakaoka, Y. Arakawa, Y. Yamamoto, and J. Vuckovic,
Physical Review Letters 95(July), 013904 (2005).
[6] H. Altug, D. Englund, and J. Vuckovic, Nature Physics 2,
484488 (2006).
[7] T. Yoshie, M. Loncar, A. Scherer, and Y. Qiu, Applied
Physics Letters 84(18), 35433545 (2004).
[8] G. Bjork and Y. Yamamoto, IEEE Journal of Quantum
Electonics 27(11), 238696 (1991).
[9] H. Altug and J. Vuckovic, Applied Physics Letters 86(11),
111102 (2005).
[10] K. Inoue, M. Sasada, J. Kawamata, K. Sakoda, and J. Haus,
Japanese Journal of Applied Physics 38, L15759 (1999).
[11] M. Meier, A. Mekis, A. Dodabalapur, A. Timko, R. E.
Slusher, J. D. Joannopoulos, and O. Nalamasu, Applied
Physics Letters 74(1), 79 (1999).
[12] J. Mouette, C. Seassal, X. Letartre, P. Rojo-Romeo, J. L.
Leclereq, P. Regreny, P. Viktorovitch, E. Jalaguier, R. Perreau, and H. Moriceau, Electronics Letters 39(6), 526528
(20 March 2003).
[13] X. Letartre, C. Monat, C. Seassal, and P. Viktorovitch, J.
Opt. Soc. Am. B 22(12), 25812595 (2005).
[14] K. Sakoda, K. Ohtaka, and T. Ueta, Opt. Express 4(12),
481489 (1999).
[15] T. D. Happ, M. Kamp, A. Forchel, J. L. Gentner, and
L. Goldstein, Applied Physics Letters 82(1), 46 (2003).
[16] D. G. Deppe, J. P. van der Ziel, N. Chand, G. J. Zydzik, and
S. N. G. Chu, Applied Physics Letters 56(21), 20892091
(1990).
[17] Y. Akahane, T. Asano, B. S. Song, and S. Noda, Nature
425(October), 944947 (2003).
[18] D. Englund, A. Faraon, B. Zhang, Y. Yamamoto, and
J. Vuckovic, Optics Express 15(April), 55508 (2007).
[19] For InAs MQW, g(nG ) = g0 vg ln(nG /Va Ntr ), Va =
active material volume, g0 2400/cm, group velocity
vg = c/n, Ntr 3 1018 cm3 (from [21]); gain confinement factor 0.16; : pump power absorption ratio;
r : SE lifetime in unpatterned QW; P C,nr : NR lifetime in
PC; E,f , E,r , E,nr : lifetimes of pump-level relaxation,
SE, and NR transitions (resp). At RT, we model SE rate
2
as Rsp = BNG
in the intrinsic QW; the ratio in B for
cavity and bulk emissions gives SE rate modification. For
InAs QDs at 10K, g(nG ) = g0 (nG /Va Ntr ), g0 =
8.2 106 cm3 /s, Ntr = 3.2 1017 cm3 , = 0.028.
[20] M. Grundmann and D. Bimberg, Phys. Rev. B 55(15), 9740
9745 (1997).
[21] L. A. Coldren and S. W. Corzine, Diode Lasers and Photonic Integrated Circuits (New York: Wiley, 1995).

www.lpr-journal.org

Laser & Photon. Rev. (2008)

[22] C. H. Chang, L. Chrostowski, and C. J. Chang-Hasnain, J.


Sel. Top. Quantum Electron. 9, 13861393 (2003).
[23] K. L. Lear and al et, Advances in Vertical Cavity Surface
Emitting Lasers in Trends in Optics and Photonics Series
15, 6974 (1997).
[24] L. Chrostowski, X. Zhao, and C. Chang-Hasnain, Microwave Theory and Techniques, IEEE Transactions on
54(2), 788796 (Feb. 2006).
[25] T. Baba, D. Sano, K. Nozaki, K. Inoshita, Y. Kuroki, and
F. Koyama, Applied Physics Letters 85(18), 39893991
(2004).
[26] M. Kretschmann and A. A. Maradudin, J. Opt. Soc. Amer.
B, Opt. Phys. 21(January), 150158 (2004).
[27] S. C. Hagness, R. M. Joseph, and A. Taflove, Radio Sci
31(4), 931941 (1996).
[28] W. H. Pernice, F. P. Payne, and D. F. Gallagher, Opt. Express 15(18), 1143311443 (2007).
[29] L. A. Coldren and S. W. Corzine, Diode Lasers and Photonic Integrated Circuits (New York: Wiley, 1995).
[30] O. Painter, R. Lee, A. Scherer, A. Yariv, J. OBrien, P. Dapkus, and I. Kim, Science 284(June), 18191821 (1999).
[31] J. K. Hwang, H. Y. Ryu, D. S. Song, I. Y. Han, H. W. Song,
H. K. Park, Y. H. Lee, and D. H. Jang, Applied Physics
Letters 76(21), 29822984 (2000).
[32] C. Monat, C. Seassal, X. Letartre, P. Viktorovitch,
P. Regreny, M. Gendry, P. Rojo-Romeo, G. Hollinger,
E. Jalaguier, S. Pocas, and B. Aspar, Electronics Letters
37(12), 764766 (7 Jun 2001).
[33] D. Englund, H. Altug, and J. Vuckovic, Appl. Phys. Lett.
91(July), 071124 (2007).
[34] D. Englund, H. Altug, I. Fushman, and J. Vuckovic, Appl.
Phys. Lett. 91(July), 071126 (2007).
[35] K. Nozaki, S. Kita, and T. Baba, Opt. Express 15(12), 7506
7514 (2007).
[36] M. Nomura, S. Iwamoto, and M. Nishioka, Appl. Phys.
Lett. 89(161111) (2006).
[37] M. H. Shih, W. Kuang, A. Mock, M. Bagheri, E. H. Hwang,
J. OBrien, and P. Dapkus, Appl. Phys. Lett. 89(101104)
(2006).
[38] C. Monat, C. Seassal, X. Letartre, P. Regreny, M. Gendry,
P. R. Romeo, P. Viktorovitch, M. L. V. dYerville, D. Cassagne, J. P. Albert, E. Jalaguier, S. Pocas, and B. Aspar,
Journal of Applied Physics 93(1), 2331 (2003).
[39] G. Vecchi, F. Raineri, I. Sagnes, A. Yacomotti, P. Monnier,
T. J. Karle, K. H. Lee, R. Braive, L. L. Gratiet, S. Guilet,
G. Beaudoin, A. Taneau, S. Bouchoule, A. Levenson, and
R. Raj, Opt. Express 15(12), 75517556 (2007).

www.lpr-journal.org

11

[40] B. B. Bakir, C. Seassal, X. Letartre, P. Regreny, M. Gendry,


P. Viktorovitch, M. Zussy, L. D. Cioccio, and J. M. Fedeli,
Opt. Express 14(20), 92699276 (2006).
[41] M. Nomura, S. Iwamoto, K. Watanabe, N. Kumagai,
Y. Nakata, S. Ishida, and Y. Arakawa, Opt. Express 14(13),
63086315 (2006).
[42] M. Asada, IEEE J. of Quant. Electron. 25(9), 20192026
(1989).
[43] J. Y. Duboz, E. Costard, E. Rosencher, P. Bois, J. Nagle,
J. M. Berset, D. Jaroszynski, and J. M. Ortega, Journal of
Applied Physics 77(12), 64926495 (1995).
[44] H. Deng, Q. Deng, and D. G. Deppe, Applied Physics Letters 70(6), 741743 (1997).
[45] H. Altug and J. Vuckovic, Opt. Express 13(22), 88198828
(2005).
[46] D. Englund, H. Altug, and J. Vuckovic, to be published
(2008).
[47] T. Yoshie, O. Shchekin, H. Chen, D. Deppe, and A. Scherer,
Electronics Letters 38(17), 967968 (2002).
[48] B. Ellis, I. Fushman, D. Englund, B. Zhang, Y. Yamamoto,
and J. Vuckovic, Appl. Phys. Lett. 90(June), 151102 (2007).
[49] J. Urayama, T. B. Norris, H. Jiang, J. Singh, and P. Bhattacharya, Applied Physics Letters 80(12), 21622164
(2002).
[50] Z. M. S Fathpour and P. Bhattacharya, Journal of Physics
D: Applied Physics 38(13), 21032111 (2005).
[51] D. R. Matthews, H. D. Summers, P. M. Smowton, and
M. Hopkinson, Applied Physics Letters 81(26), 49044906
(2002).
[52] D. Deppe, H. Huang, and O. Shchekin, Quantum Electronics, IEEE Journal of 38(12), 15871593 (Dec 2002).
[53] H. Park, S. Kim, S. Kwon, Y. Ju, J. Yang, J. Baek, S. Kim,
and Y. Lee, Science 305, p. 14441447 (2004).
[54] M. K. Seo, K. Y. Jeong, J. K. Yang, Y. H. Lee, H. G. Park,
and S. B. Kim, Applied Physics Letters 90(17), 171122
(2007).
[55] H. Matsubara, S. Yoshimoto, H. Saito, Y. Jianglin,
Y. Tanaka, and S. Noda, Science 319(5862), 445447
(2008).
[56] R. Schmidt, U. Scholz, M. Vitzethum, R. Fix, C. Metzner,
P. Kailuweit, D. Reuter, A. Wieck, M. C. Hubner, S. Stufler,
A. Zrenner, S. Malzer, and G. H. Dohler, Applied Physics
Letters 88(12), 121115 (2006).

2008 by WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like