formation of bound states in the continuum for a quantum dot with variable width

9
PHYSICAL REVIEW B 83, 235321 (2011) Formation of bound states in the continuum for a quantum dot with variable width Evgeny Bulgakov 1,2 and Almas Sadreev 1 1 Kirensky Institute of Physics, RU-660036, Krasnoyarsk, Russia 2 Siberian State Aerospace University, Krasnoyarsk Rabochii, 31, Krasnoyarsk, Russia (Received 7 December 2010; revised manuscript received 28 February 2011; published 14 June 2011) We consider mechanisms of formation of the bound states in the continuum in open rectangular quantum dots with variable width. Because of symmetry there might be bound states in the continuum (BSCs) embedded into one, two, and three continua because of the symmetry of system. These BSCs arise for selected values of the width. We show numerically that the BSCs can be excited for transmission of wave packets if the quantum dot width is varied in time and reaches these selected values of dot width. Moreover, we consider numerically a decay process of different eigenstates in the closed quantum dot to show that some of them trap in the BSC after the quantum dot is opened. DOI: 10.1103/PhysRevB.83.235321 PACS number(s): 03.65.Nk, 05.60.Gg, 73.23.Ad, 73.21.La I. INTRODUCTION In 1929, von Neumann and Wigner 1 claimed that the single-particle Schr¨ odinger equation could possess localized solutions that correspond to isolated discrete eigenvalues embedded in the continuum of positive energy states for some artificial oscillation bounded potential. Extension and some correction of this work was done by Stillinger and Herrick 2 who presented a few examples of spherically symmetric attractive local potentials with bound states embedded in the continuum (BSCs) of scattering states in the context of possible BSCs in atoms and molecules (e.g., Refs. 36). Friedrich and Wintgen 7 discussed the realization of BSCs in a hydrogen atom in a magnetic field. Afterward, it was suggested that BSCs might be found in certain two-electron systems. 2,8 Later in 1973 Herrik 9 and Stillinger 10 predicted a BSC in semiconductor heterostructure superlattices with specific choices of each layer in the superlattice in analogy with the von Neumann and Wigner oscillating potential. 1 This BSC was observed by Capasso et al. as a very narrow absorption peak. 11 Examples of BSCs can be more easily found if one goes beyond the one-dimensional Schr¨ odinger equation; for example, in model systems 1220 or in real two-dimensional (2D) systems. 2131 In the hard-wall approximation they are the solutions of the two-dimensional equation −∇ 2 ψ (x,y ) = (2 )ψ (x,y ). (1) Here, E is measured in units of the characteristic electron energy E 0 = π 2 ¯ h 2 /(2md 2 ) with d being the width of the leads. A typical view of the open system is presented in Fig. 1. Note that the Helmholz equation (1) is applied not only for electron transmission through the quantum dot (QD) but also for microwave and acoustic transmission through billiards 32 and optical transmission in photonic crystals. 31,33 It is widely accepted that, below the well-defined threshold for the propagating states equal to 1, the eigenenergies of the Schr¨ odinger equation are discrete and the corresponding eigenfunctions are square-integrable bound states. Above the threshold, the eigenenergies are distributed continuously and the corresponding eigenfunctions are extended. However, there might be bound states with discrete energies above the threshold (i.e., BSCs). Numerous considerations in model and real 2D open systems have demonstrated the BSC by a vanishing of the resonance width. 1230 The main feature of the BSC is that an incident wave passing through the QD has no coupling with the BSC; that is, the BSC can not be excited by this wave. We can say that the BSC is a “true dark state” as was proposed by Muraguchi et al. 34 Therefore, there is a question of excitement of the BSC by the incident wave. If even to transmit a wave packet (WP) through the scattering system the BSC cannot be excited. However, we can violate the orthogonality of the BSC to the scattering state by varying the QD’s shape for transmission of the WP. We show numerically that this strategy is able to excite the BSC. Another strategy to achieve the BSC is to excite an eigenstate of the closed QD and open the QD afterward. However, before a study of these processes we consider mechanisms by which BSCs arise in open rectangular QDs. II. MECHANISMS OF BOUND STATES IN CONTINUUM IN OPEN QUANTUM DOTS The first, most obvious mechanism for the BSC is related to a symmetry of the system. Let us write the total Hamiltonian of the system of QD plus two leads (see Fig. 1) as follows: 35,36 H = C=L,R p C p 2 C dE|E,C,p C Ep C ,E,C|+ b |b b b| + C=L,R b p C p 2 C dE|b V bp C p C ,E,C|+ H.c., (2) where the index b enumerates the eigenstates of the closed QD, the index p C = 1,2,3,... does the continua (the channels of propagation) of the right and left lead, and V bp C are the coupling matrix elements between the eigenstates of closed QD and mode p C of semi-infinite lead C. We consider that the leads are identical. Then the coupling matrix elements do not depend on the lead C = L,R and are as follows: 35,36 V bp = 1 2πk p 1/2 1/2 dy sin[πp(y 1/2)] ∂φ b (0,y ) ∂x . (3) Here, E = p 2 + k 2 p 2 , and k p is the dimensionless wave number of channel p in units of the lead’s width, 1/d . We consider that the system is symmetric relative to an 235321-1 1098-0121/2011/83(23)/235321(9) ©2011 American Physical Society

Upload: almas

Post on 04-Dec-2016

214 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Formation of bound states in the continuum for a quantum dot with variable width

PHYSICAL REVIEW B 83, 235321 (2011)

Formation of bound states in the continuum for a quantum dot with variable width

Evgeny Bulgakov1,2 and Almas Sadreev1

1Kirensky Institute of Physics, RU-660036, Krasnoyarsk, Russia2Siberian State Aerospace University, Krasnoyarsk Rabochii, 31, Krasnoyarsk, Russia

(Received 7 December 2010; revised manuscript received 28 February 2011; published 14 June 2011)

We consider mechanisms of formation of the bound states in the continuum in open rectangular quantum dotswith variable width. Because of symmetry there might be bound states in the continuum (BSCs) embedded intoone, two, and three continua because of the symmetry of system. These BSCs arise for selected values of thewidth. We show numerically that the BSCs can be excited for transmission of wave packets if the quantum dotwidth is varied in time and reaches these selected values of dot width. Moreover, we consider numerically adecay process of different eigenstates in the closed quantum dot to show that some of them trap in the BSC afterthe quantum dot is opened.

DOI: 10.1103/PhysRevB.83.235321 PACS number(s): 03.65.Nk, 05.60.Gg, 73.23.Ad, 73.21.La

I. INTRODUCTION

In 1929, von Neumann and Wigner1 claimed that thesingle-particle Schrodinger equation could possess localizedsolutions that correspond to isolated discrete eigenvaluesembedded in the continuum of positive energy states for someartificial oscillation bounded potential. Extension and somecorrection of this work was done by Stillinger and Herrick2

who presented a few examples of spherically symmetricattractive local potentials with bound states embedded inthe continuum (BSCs) of scattering states in the context ofpossible BSCs in atoms and molecules (e.g., Refs. 3–6).Friedrich and Wintgen7 discussed the realization of BSCsin a hydrogen atom in a magnetic field. Afterward, it wassuggested that BSCs might be found in certain two-electronsystems.2,8 Later in 1973 Herrik9 and Stillinger10 predicteda BSC in semiconductor heterostructure superlattices withspecific choices of each layer in the superlattice in analogy withthe von Neumann and Wigner oscillating potential.1 This BSCwas observed by Capasso et al. as a very narrow absorptionpeak.11

Examples of BSCs can be more easily found if onegoes beyond the one-dimensional Schrodinger equation; forexample, in model systems12–20 or in real two-dimensional(2D) systems.21–31 In the hard-wall approximation they are thesolutions of the two-dimensional equation

−∇2ψ(x,y) = (Eπ2)ψ(x,y). (1)

Here, E is measured in units of the characteristic electronenergy E0 = π2h2/(2md2) with d being the width of theleads. A typical view of the open system is presented inFig. 1. Note that the Helmholz equation (1) is applied notonly for electron transmission through the quantum dot (QD)but also for microwave and acoustic transmission throughbilliards32 and optical transmission in photonic crystals.31,33

It is widely accepted that, below the well-defined thresholdfor the propagating states equal to 1, the eigenenergies ofthe Schrodinger equation are discrete and the correspondingeigenfunctions are square-integrable bound states. Above thethreshold, the eigenenergies are distributed continuously andthe corresponding eigenfunctions are extended. However,there might be bound states with discrete energies abovethe threshold (i.e., BSCs). Numerous considerations in model

and real 2D open systems have demonstrated the BSC by avanishing of the resonance width.12–30

The main feature of the BSC is that an incident wave passingthrough the QD has no coupling with the BSC; that is, the BSCcan not be excited by this wave. We can say that the BSC isa “true dark state” as was proposed by Muraguchi et al.34

Therefore, there is a question of excitement of the BSC by theincident wave. If even to transmit a wave packet (WP) throughthe scattering system the BSC cannot be excited. However, wecan violate the orthogonality of the BSC to the scattering stateby varying the QD’s shape for transmission of the WP. Weshow numerically that this strategy is able to excite the BSC.Another strategy to achieve the BSC is to excite an eigenstateof the closed QD and open the QD afterward. However, beforea study of these processes we consider mechanisms by whichBSCs arise in open rectangular QDs.

II. MECHANISMS OF BOUND STATES IN CONTINUUM INOPEN QUANTUM DOTS

The first, most obvious mechanism for the BSC is related toa symmetry of the system. Let us write the total Hamiltonianof the system of QD plus two leads (see Fig. 1) as follows:35,36

H =∑

C=L,R

∑pC

∫ ∞

p2C

dE|E,C,pC〉E〈pC,E,C| +∑

b

|b〉εb〈b|

+∑

C=L,R

∑b

∑pC

∫ ∞

p2C

dE|b〉 VbpC〈pC,E,C| + H.c., (2)

where the index b enumerates the eigenstates of the closedQD, the index pC = 1,2,3, . . . does the continua (the channelsof propagation) of the right and left lead, and VbpC

are thecoupling matrix elements between the eigenstates of closedQD and mode pC of semi-infinite lead C. We consider that theleads are identical. Then the coupling matrix elements do notdepend on the lead C = L,R and are as follows:35,36

Vbp =√

1

2πkp

∫ 1/2

−1/2dy sin[πp(y − 1/2)]

∂φb(0,y)

∂x. (3)

Here, E = p2 + k2p/π2, and kp is the dimensionless wave

number of channel p in units of the lead’s width, 1/d.We consider that the system is symmetric relative to an

235321-11098-0121/2011/83(23)/235321(9) ©2011 American Physical Society

Page 2: Formation of bound states in the continuum for a quantum dot with variable width

EVGENY BULGAKOV AND ALMAS SADREEV PHYSICAL REVIEW B 83, 235321 (2011)

d d x

yW

L

FIG. 1. Schematic view of rectangular QD with attached waveg-uides in hard-wall approximation.

inversion of the axis x, which is perpendicular to the transportaxis x, as shown in Fig. 1. Then the eigenstates of the closedQD and the eigenmodes of the semi-infinite leads can beclassified as even and odd states. In particular, the first, third,fifth, . . . modes which define channels of the straight leads areeven modes, while the second, fourth, . . . modes are odd ones.

Let us consider the odd eigenstate φb of the closed QDwith εb < 4. Then the matrix element coupling that statewith the first even-channel mode is Vb1 = 0. Respectively,the state will not leak into the leads and could be definedas the BSC. However, the state φb is not eigen for the totalHamiltonian because of coupling with closed (evanescent)channels with p = 3,5, . . .. As a result the last will slightlydistort the odd eigenstate φb, transforming it into the BSC thatis the eigenstate of the total Hamiltonian (2). We refer to thepaper by Moiseyev37 where an analysis of these BSCs basedon symmetry arguments was presented.

The second mechanism of the BSC comes from a couplingof the eigenstate with the channel mode, which might go tozero upon varying the QD’s width. For numerical purposes wetake the model potential along the y axis in the following form:

V (y) = V0{1 + 0.5[tanh[C(y − W/2)]

− tanh[C(y + W/2)]]}, (4)

which is shown in Fig. 2(a) where W is the effective widthof the QD. This potential makes a QD with soft walls alongthe y axis. Let the eigenstate parity of the QD and the parity

of the channel mode coincide so that Vb,p �= 0. We start withthe first even channel. For variation of the QD’s width, thecoupling matrix element might go to zero. It is clear that thismight happen only for high eigenenergy to provide enoughoscillations of the derivative ∂φb(0,y)

∂xin the integral over y.

Numerics shows that Vb,1 indeed goes to zero for εb > 4.Assume that, owing to the second mechanism, there is

an odd BSC with the eigenenergy Ec uncoupled from theodd mode of the second-channel continuum at some fixedwidth W . Then, because of the parity of the first- and third-channel modes, this odd BSC is decoupled from the first- andthird-channel continua. Therefore, by varying the QD’s width,we can obtain a BSC embedded into two continua (p = 1,2)if Ec < 9 or even three continua (p = 1,2,3) if EC < 16. Wedefine these BSCs as second or third order, respectively. Thiscontrasts with the quasibound states in a two-channel quantumwire that have been previously reported.20

The third mechanism of the BSC is the more sophisticated.Let each eigenstate of the QD be coupled with the continuum.Upon opening of the QD, they become resonances given bypoles of the S-matrix in the complex energy plane. Then theBSC is the result of destructive interference of the resonances,which might be full with variation of the QD’s width. Thatmechanism for the BSC was shown analytically in a two-levelapproximation for QDs coupled with one continuum38–42 andwas realized in two-dimensional systems. In particular, thethird mechanism for the BSC has been realized in a quasi-one-dimensional constriction with an attractive, finite-size impurityby varying the size of the impurity,21 in a bend waveguide forvariation of bending,22 and in an open QD for variation of theQD’s width.24 In Fig. 2(a) we present the pattern of a BSC witheigenenergy Ec = 1.5845 embedded into the continuum of thefirst channel, which is the result of the full destructive inter-ference of two resonance states at W = Wc = 2.7498,L = 4.For the varying of the QD’s width the conductance displayesavoided crossing of zero conductance with unit one.24 Thisphenomenon was recently observed by Lepetit et al. inhigh-permettivity dielectric metamaterial resonator.43

Again, similar to the second mechanism, the BSC haszero coupling with the even first- and third-channel modesbecause of the destructive interference of odd-resonance states.

FIG. 2. (Color online) Patterns of BSCs with eigenenergy attributable to the third mechanism. (a) BSC with eigenenergy Ec = 1.5845 isthe result of interference of two resonances that become full for the QD width W = Wc = 2.7498,L = 4. The width is governed by the softpotential (4), which squeezes the QD as schematically shown at left. (b) Extremely excited BSC with eigenenergy Ec = 13.329 embedded intothree continua p = 1,2,3 for W = 2.8731,L = 4.

235321-2

Page 3: Formation of bound states in the continuum for a quantum dot with variable width

FORMATION OF BOUND STATES IN THE CONTINUUM . . . PHYSICAL REVIEW B 83, 235321 (2011)

Therefore, such a BSC might be embedded simultaneouslyinto the first-, second-, and third-continuum channels. Wepresent in Fig. 2(b) a pattern of this extremely excited third-order BSC with eigenenergy Ec = 13.329. However, the mostinteresting thing is that this BSC is the result of full destructiveinterference of, at least, three resonance states. Thus, wereport the numerical solution of the problem formulated byMiyamoto:44 could the BSC exist because of the multilevelinterference effect b = 1,2,3, . . . for Vb,p �= 0?

The concept of a Fano mirror is the fourth mechanismfor BSCs.13,45,46 Let the closed system be composed oftwo identical scattering centers connected by a wire. If thetransmission through each scattering center becomes zero, theycan serve as ideal mirrors as in a Fabry-Perot interferometer.Then BSCs arise if an integer number of half waves arespaced between the Fano mirrors. This mechanism, exclusivelytransparent, was applied to a typical one-dimensional double-barrier structure with a periodically driven potential barrier,45

to a photonic crystal structure with a waveguide coupled withtwo single-mode cavities,13,46 and to two coupled identicalquantum dots.15,25,27,47–49

Irrespective of the physical mechanism of the BSC, ageneral theory of BSCs can be formulated as follows: Thescattering wave function in the interior of the QD obeys theLippmann-Schwinger equation36

(Heff − E)|ψ〉 = V |E,p〉, (5)

where |E,p〉 is the state of channel p, and V is the couplingmatrix given in Eq. (3). This formulation of the Lippmann-Schwinger equation implies that the wave is incident at channelp of the left lead and excites the state in the interior of the QD.The effective non-Hermitian Hamiltonian Heff is the result ofthe Feshbach projection of the total Hamiltonian into the innerspace of the closed QD:35,36,42,50–52

〈b|Heff|b′〉 = εbδbb′ + 2∞∑

p=1

Vbp

1

E+ − Hp

Vpb′ . (6)

Here the factor 2 takes into account that the leads areidentical. After some algebra the effective Hamiltonian takesthe form36,53

〈b|Heff|b′〉 = εbδbb′ − 2∞∑

p=1

eikpVbpVb′p. (7)

The BSC that occurs for the inversion of matrix Heff − E

does not exist [i.e., det(Heff − E) = 0].18,24,31 That equation isa special case of the equation for the complex eigenvalues ofthe effective Hamiltonian

Heff|λ) = zλ|λ), (8)

provided that one of the complex zλ becomes real. Here, thecomplex eigenvalues zλ = Eλ − iλ/2 where Eλ define theenergy positions of resonances with resonance widths λ

51,52

if the resonance width is sufficiently small compared to thedistance between resonances. The right eigenstates |λ) withthe left eigenstates (λ| form the biorthogonal basis.36,51

Therefore, the condition for a BSC det(Heff − E) = 0 isequivalent to the often-used definition of a BSC as a resonancestate |λ) with vanishing width,38,39,44,54 which is an eigenstateof the effective Hamiltonian

Heff(Ec,Wc)|BSC〉 = Ec|BSC〉. (9)

〈BSC|V |E,p〉 = 0 for those channels p to which continuumthe BSC is embedded. Comparison with Eq. (5) shows that theBSC is a homogeneous solution of the Lippmann-Schwingerequation. The solutions of Eq. (9) are shown in Fig. 2.

Let the energy be below the bottom of the secondpropagation band: E < 4 [i.e., p = 1 in Eq. (5)]. Then allkp are imaginary except the first-channel wave number k1.In particular, for the BSC energy we obtain k2 = 2.3926i.Assume we can disregard all channels except the first becauseof the small value of exp(−|kp|),p > 1 and write the effectiveHamiltonian as follows:36,43

〈b|Heff|b′〉 = εbδbb′ − 2eik1Vb1Vb′1. (10)

2.73 2.74 2.75 2.76 2.770

1

2

3

4

5

6

7

8x 10

−4

W

(b)(a)

FIG. 3. (Color online) (a) Eigenfunctions φ33 and φ5,1 (in insets) and eigenenergies ε3,3 and ε5,1 of the closed QD versus the QD’s width W

are shown by dashed lines. The resonance energies Er are shown by solid lines. The lower line is the resonance position whose width goes tozero at W = Wc = 2.7498, as shown in (b), and therefore gives rise to the BSC marked by the open star. The upper solid line corresponds tothe superradiant solution24,41 shown in inset. The corresponding resonance width with maximal width is not shown. Inset in bottom copies theBSC in Fig. 2.

235321-3

Page 4: Formation of bound states in the continuum for a quantum dot with variable width

EVGENY BULGAKOV AND ALMAS SADREEV PHYSICAL REVIEW B 83, 235321 (2011)

0 2 4 62.7

2.8

2.9

3

3.1

3.2

3.3

t

W

Wc

FIG. 4. Time evolution of the QD width (solid line). QD widthstarts with W = 3.2, then linearly decreases and reaches the widthWc = 2.7498 at which the BSC occurs.

For the effective Hamiltonian (10) the equation for the BSCtakes the following form:

det(Heff − E) =∏b

(E − εb)

(1 + 2eik1

∑b

V 2b1

E − εb

)= 0.

(11)

The equation is satisfied if the closed QD becomes degenerate(i.e., εb = εb′ ).43 Mathematically, this result is based on thecontribution of the continuum in the effective Hamiltonianbeing given by products of the matrix elements.

However, there is in fact an avoided crossing of the nearestresonance positions Eλ, as shown in Fig. 3(a) by solid lines.The first resonance width reaches zero at W = Wc as shownin Fig. 3(b). Respectively, with correspondence to Eq. (9), theresonance state becomes a BSC as shown in bottom inset ofFig. 3(a). While the second resonance takes maximal width38,39

[not shown in Fig. 3(b)] with the corresponding eigenstate (thesuperradiant state41) shown in Fig. 3(a) by the top inset. TheBSC with eigenenergy Ec = 1.5845 is marked by the star inFig. 3(a) and does not coincide with the crossing point ofthe nearest eigenenergies of the closed QD, shown by dashed

FIG. 5. (Color online) Snapshots of transmission of the wavepacket by absolute value through the QD for the QD temporalsqueezing shown in Fig. 4. Dark blue in snapshots corresponds tozero of wave function. (a) t = 0, (b) t = 1, and (c) t = 5.

0 5 10 15 20 250

0.1

0.2

0.3

0.4

0.5

t

p(t)

FIG. 6. Time evolution of the probability to be a quantum particlein the interior of QD for the wave-packet propagation shown in Fig. 5.

lines. The channels with p > 1 are the reason for this slightdiscrepancy. In fact, the effective Hamiltonian has the thefollowing form:

〈b|Heff|b′〉 = εbδbb′ − 2eik1Vb1Vb′1 − 2∞∑

p=2

e−|kp |VbpVb′p,

(12)

for E < 4. Then the determinant of the matrix (Heff − E) can-not be rigorously reduced to Eq. (11), and equation (9) for theBSC can be solved only numerically with the results presentedin Fig. 3. Numerics give that the BSC has energy Ec = 1.5845and the resonance width goes to zero at W = Wc = 2.7498.These values at the BSC point are slightly different fromthe degeneracy point Wd = 2.7405,εd = 1.5921. Thus, theevanescent modes of leads shift the BSC point from thedegeneracy point. These results are in full agreement with theBSC occurring at those points where the resonance width goesto zero while the resonances undergo avoided crossings.24,39

The effect of evanescent modes might be more profound forBSCs of higher eigenenergies, as shown later. In fact, even forthe bound state with energy below the propagation band ofclosed channels, their contribution might be important.55

0 2 4 6 8 101

1.5

2

2.5

3

3.5

t

<E

(t)>

FIG. 7. Time evolution of the mean energy of the wave packet.

235321-4

Page 5: Formation of bound states in the continuum for a quantum dot with variable width

FORMATION OF BOUND STATES IN THE CONTINUUM . . . PHYSICAL REVIEW B 83, 235321 (2011)

FIG. 8. (Color online) (a) Even eigenstate φ43 (a) with the eigenenergy ε43 = 1.9975. (b) The state after t = 100. Sizes of the rectangularsoft quantum dot are L = 4,Wc = 2.7498.

III. PROPAGATION OF WAVE PACKET THROUGHQUANTUM DOT WITH TIME-DEPENDENT SHAPE

The BSC is the square-integrable; that is, a localized eigen-function of the total Hamiltonian (2) with the correspondingdiscrete eigenenergy embedded into the propagation band ofscattering states. The BSC is orthogonal to other extendedeigenstates of the total Hamiltonian, which are plane wavesfar from the scattering region presented by the QD underconsideration. In other words, the BSC is “invisible” forwaves transmited through the QD. Therefore, the BSC isorthogonal to the WP that is an expansion over the waves.For the propagating WP that could “see” the BSC, we violatetemporarily this orthogonality for the transmission. One canapply a laser field of dipole type e�r · �E0 sin ωt that breaksthe parity of bound states and therefore mixes the boundstate with the transport solution.56 A similar approach wasused recently by Muraguchi et al.34 It was shown that, fortransmission through the lozenge QD, the WP populates theeven eigenstate of the QD that was transited into the oddbound state by absorbing a photon quantum. This last statehas zero coupling with the first channel of leads and couldtherefore be the BSC if its energy were below 4. However,in Ref. 34 the energy of the stimulated odd bound stateexceeded 4 to give rise to a decay into the second channel.In order to trap this state in the interior of the QD, sufficientlylarge potential barriers between the QD and leads wereapplied. That state was defined in Ref. 34 as a “quasi darkstate.”

We consider propagation of the WP through the symmetricrectangular QD and show that the BSC (true dark state)can be excited. In order to violate the orthogonality of theBSC to the WP we apply a slow time-dependent variationof the width of the QD. For that we apply the potential(4), which squeezes the QD’s width W linearly in time asshown in Fig. 4. The parameters of the potential are thefollowing: V0 = 10, C = 17. For the QD, time dependenceof the QD’s shape can be easily achieved by an application oftime-dependent gate voltage.

The final width of the billiard exactly corresponds to theBSC point shown in Fig. 3(a) by the star. For t = 0 the WPwas taken as follows:

ψWP(x,y,t = 0)

= sin[π (y − 1/2)] exp[−(x + x0)2/(2σ 2) + ik0x], (13)

with x0 = 5, σ = 1.5, and k0 = 2.5. It is located in the leftinput lead as shown in Fig. 5(a). In the next snapshots[Figs. 5(b) and 5(c)] one can see that, after the process ofthe WP transmission and the squeezing of the QD’s width W ,some part of the wave packet is trapped in the interior of theQD roughly after t = 4. Quantitatively, the process of trappingis shown in Fig. 6 as the time evolution of the probability tofind the particle in the interior of the QD. Around time t = 2the probability reaches a sharp maximum and afterward fallsdown and keeps constant for t � 4. The corresponding trappedwave function is shown in Fig. 5(c). Based on these figures wecan conclude that these two processes of WP propagation andsqueezing of the QD’s width are able to excite the bound statewith discrete energy within the first-channel continuum.

The BSC pattern shown in Fig. 2 is very close to the trappedwave function shown in Fig. 5 (c). Moreover, in Figure 7 weplotted the time evolution of the mean energy

〈E(t)〉 =∫A

dxdyψ∗WP(x,y,t)HψWP(x,y,t)∫

Adxdy|ψWP(x,y,t)|2 ,

for the transmission of the wave packet through the QD. Here A

is the area of the QD. The mean energy converges into the BSCeigenenergy Ec = 1.5844. Thus, these results unambiguouslyshow that a part of the wave packet trapped in the QD is theBSC.

0 5 10 15 200

0.2

0.4

0.6

0.8

1

t

p(t)

FIG. 9. Time evolution of the probability to find the quantumparticle in the interior of the QD for the even initial state φ4,3 (solidline) and for the odd initial state φ3,4 (dashed line).

235321-5

Page 6: Formation of bound states in the continuum for a quantum dot with variable width

EVGENY BULGAKOV AND ALMAS SADREEV PHYSICAL REVIEW B 83, 235321 (2011)

FIG. 10. (Color online) (a) Initial state is taken as the odd eigenstate φ3,4 with eigenenergy ε3,4 = 2.5735. (b) Temporary state after t = 100that is the BSC. QD sizes are the same as in Fig. 8.

IV. DECAY OF EIGENSTATES OF CLOSED QDAFTER OPENING

It is possible to achieve the BSC by decay of the eigenstatesof the closed QD after opening an access to the attachingleads. The coupling can be governed by a voltage applied tothe quantum point contact gates between the QD and electronwaveguide. In microwave billiards the coupling can be variedvia an electrooptic shutter as a diaphragm between the billiardand the waveguide.32

As given by Eq. (9), the BSC is the eigenstate of the effectivenon-Hermitian Hamiltonian. Therefore, it is reasonable touse the biorthogonal basis of the eigenstates of the effectiveHamiltonian35,36,51,57

Heff|λ) = zλ|λ), (λ|Heff = zλ(λ|, zλ = Eλ − iλ/2,

(λ|λ′) = δλ,λ′ . (14)

From the biorthogonality condition it follows that∑λ

|λ)(λ| = 1. (15)

Moreover, by Hermitian conjugation of Eq. (14) we have

(λ|∗H †eff = z∗

λ(λ|∗. (16)

These conjugated eigenstates form a biorthogonal basis, too.Then, for the probability, we have to find the particle in the

interior of the QD:

p(t) = 〈ψB(t)|ψB(t)〉 =∑λ,λ′

〈ψB(0)|λ∗)(λ∗| exp(iH †eff t)

× exp(−iHeff t)|λ′)(λ′|ψB(0)〉=∑λ,λ′

〈ψB(0)|λ∗)(λ∗|λ′)

× exp[i(Eλ − Eλ′)t − (λ + λ′)t](λ′|ψB(0)〉. (17)

FIG. 11. (Color online) (a) Even eigenstate φ51 with the eigenenergy ε51 = 1.59 for W = 3. (b) That state fast decays into the quasi BSCclose to the exact BSC shown in Fig. 2 after t = 1. (c) After this, the quasi BSC decays over a long time and evolves into the truly bound statebelow the propagation band of the first channel after t = 100.

235321-6

Page 7: Formation of bound states in the continuum for a quantum dot with variable width

FORMATION OF BOUND STATES IN THE CONTINUUM . . . PHYSICAL REVIEW B 83, 235321 (2011)

0 5 10 15 200

0.2

0.4

0.6

0.8

1

t

p(t)

FIG. 12. Time evolution of the probability to find a quantumparticle in the interior of the QD for the initial state φ5,1 andeigenenergy ε5,1 = 1.59. The parameters of the QD are W = 3,

L = 4.

where a subscript “B” means the state projected into theinterior of the QD. Therefore, for the limit t → ∞ we obtainthat only the state with λ = 0 will survive in the interior of theQD. This result is in full correspondence with the definition ofthe BSC given by Eq. (9).

At first, we consider the decay of the even eigenstate φb ofthe closed QD with the eigenenergy εb < 4. The eigenstatesof the closed QD with soft walls along the y axis are

φm,n(x,y) =√

2

Lsin

(mπ (x − L/2)

L

)φn(y; W ). (18)

We assume that, for the initial time, the pure state ψm,n

with odd integer n is populated, and after an access intothe leads is arisen. We can expect the probability to decaybecause of leakage into the leads with the characteristicdecay time τb,1 = 1

b,1with b,1 = 2π |Vb1|2.35 The factor 2 in

the decay time is because two identical leads are attached tothe QD. The subscript “1” means the first-channel mode of theleads.

Let us take, for example, the even eigenfunction as b =(4,3) with the corresponding eigenenergy ε4,3 = 1.9975 forW = Wc = 2.7498,L = 4. We evaluated the resonance widthb1 ≈ 0.22 in units of E0/h where we used Eq. (3) to obtaink1 ≈ π

√εb − 1. The initial state φ4,3 evolves after long time

t = 100 into the state shown in Fig. 8(b) that is close to theeigenstate φ4,1. However, because of evanescent modes ofthe waveguides it blows up into the bound state below thepropagation band of the first channel, as seen from Fig. 8(b).The corresponding probability p(t) = ∫

A|ψ(x,y,t)|2dxdy to

find a quantum particle in the interior of the QD has typicalexponential behavior p(t) ≈ e−b1t , as shown in Fig. 9. Thisresult allows us to evaluate b1 ≈ 0.223, which agrees withthe earlier evaluated result for b1.

Next, we consider the odd eigenstate φ3,4 that has zerocoupling with the first-channel mode of leads: b,1 = 0.Therefore, the particle in this state will not leak into the leads.In other words, the odd eigenstates of the closed QD with theeigenenergies below 4 are expected to be the BSCs. In fact,they are not truly the eigenstates of the total Hamiltonian (2)because of their coupling with the next evanescent modes ofthe leads with p > 1. As a result, the initial state ψ(0) = φ3,4

will slightly blow up from the QD and distort as shown inFigs. 10(a) and 10(b) to finally transfer into the true BSC. Thecorresponding time evolution of the probability is shown inFig. 9 by the dashed line.

As for the third initial state, we take the eigenstate φ5,1, theparity of which coincides with the parity of the first-channelcontinuum. We take the width W = 3 close to the BSCpoint Wc = 2.7405. The eigenenergy ε5,1 = 1.59 for W = 3 isslightly above the BSC energy Ec = 1.5844. Therefore, for anexpansion of φ5,1 over the eigenstates of the effective Hamil-tonian we may expect that the contribution of that eigenstate|λc(W )) is important, which transfers into the BSC for W →Wc. Indeed, the snapshot in Fig. 11(b) reveal a fast recovery ofthe state close to the BSC shown in Fig. 2. Nevertheless, thestate |λc(W )) is not the true BSC for W �= Wc and has thereforesmall but finite λ(W ). Then, in correspondence with Eq. (17),we have a slow decay of that state into some truly boundstate with energy below the propagation band, as shown inFig. 11(c). Indeed, the time evolution of the probability to finda particle in the interior of the QD reveals two decay times, asFig. 12 shows.

Finally, we consider the decay of the highly excited eigen-states whose eigenenergies are embedded into the propagationband of the second- and third-continuum channels. In order forthe soft potential V (y) to confine the particle, we take V0 = 10if εb < 9 and V0 = 25 if 9 < εb < 16. Let us take, first, theeven eigenstate φ6,7 with the eigenenergy ε6,7 = 6.9253 thatis below the third-channel propagation band. The state isuncoupled with the second channel, but is coupled with the first

FIG. 13. (Color online) (a) Initial state is taken as the even eigenstate φ6,7 with eigenenergy ε6,7 = 7.02. (b) Temporary state after t = 100.QD sizes are L = 4,Wc = 2.9728. This width corresponds to the second-order BSC with energy Ec = 6.9253.

235321-7

Page 8: Formation of bound states in the continuum for a quantum dot with variable width

EVGENY BULGAKOV AND ALMAS SADREEV PHYSICAL REVIEW B 83, 235321 (2011)

FIG. 14. (Color online) (a) Pattern of eigenstate φ4,10 with eigenenergy ε4,10 = 11.753 that undergoes noticeable distortion upon decay tobecome the BSC with eigenenergy Ec = 11.7105 embedded into three continua p = 1,2,3 for Wc = 2.9183,L = 4 (b).

channel. Therefore, the state will decay into the first-channelcontinuum. However, for variation of the QD width, the statemight become uncoupled with the first channel for W =2.9728. Thereby we can obtain the second-order BSC. Indeed,Fig. 13 shows that, after t = 100, the eigenstate φ6,7 undergoesnoticeable changes but finally is trapped in the interior ofthe QD. An effect of evanescent modes is substantially moreprofound for the BSC with high eigenenergies, as seen fromFigs. 13 and 14. The strong effect of evanescent modes is seenalso in Fig. 2(b).

The time evolution of the probability agrees with that result.If we took the QD’s width W different from Wc = 2.9728 theeigenstate φ6,7 would decay. The closer W is to Wc the longeris the decay process.

Next, let us take the odd eigenstate φ4,10 with eigenenergyε4,10 = 11.753 that is uncoupled with the first-, second-,and third-channel modes for W = 2.9183. It also undergoesnoticeable change (e.g., in energy and in wave-functionpattern) to become a third-order BSC, as shown in Fig. 14.

V. CONCLUDING REMARKS

The bound state in continuum (BSC) has a similaritywith the bound states below the propagation band of thecontinuum. They are both eigenstates of the total Hamiltonian(2), which describes the QD with leads attached. They arealso both localized mainly in the QD, but have exponentialtails in the leads because of coupling of the QD with theleads, and they are orthogonal to other extended eigenstatesof the total Hamiltonian. The difference between them is inthe positions of their discrete eigenenergies. The BSC hasthe energy within the propagation band of the continuum.Moreover, because of the symmetry of the QD, the BSC maybe embedded simultaneously into two or three continua. Wedefined these BSCs as being of second or third order. As wasconsidered in Sec. II, there might be at least four mechanismsfor the BSC. The first two are similar mechanisms that arerelated to orthogonality of the eigenstate of the closed QD tothe continuum channel mode. The third mechanism is relatedto the full destructive interference of at least two nearestresonance states.38,39 Moreover, we present in Fig. 2(b) apattern of the BSC that is the result of full destructive

interference of at least three states. Finally, the Fabry-Perotmechanism is the fourth mechanism.

Irrespective of the origin of the BSC, it does not decay.This obvious result is a direct consequence of the resonancewidth going to zero upon approaching the BSC point (inthe present case W → Wc). The property of orthogonalityof the BSC to the extended states of the system means that,if a plane wave is incident from the lead it has no couplingwith the bound states. In other words, the bound states are“invisible” for incident waves. We can say that they are “truedark states,” as was proposed by Muraguchi et al.34 Thatstatement is correct even if a quantum particle is incident inthe form of a WP that is a linear superposition of the planewaves.

There is a simple way to “observe” the bound states belowthe continuum. One can apply a time-periodic perturbationwhose frequency exceeds the distance between the bottomof propagation band and the energy of the bound state.56 Asimilar approach was used in Ref. 34 where a laser fieldwas applied to transfer the even bound state of the closedQD into the odd quasi BSC. Then the propagating wavepacket is able to “see” this quasi BSC via the laser fieldmixing.

For the BSC we used two ways to “observe” the BSC. Thefirst is to shrink the WD’s width in time in order to reach thepoint of the BSC. If, for this time process, the wave packet ispassing through the QD, it is able to excite the BSC, as ournumerics show. Another way is the process of time decay of theproper bound states of the closed QD when at the initial timeconnection between the QD and leads is opened. This can bedone if the gate voltage applied to quantum point contactsbetween the QD and the electron waveguide is decreasedtoo fast. To use the microwave analog of the open QD32

one can use an electrooptic shutter between the leads andmicrowave billiard in the case of microwave open billiards.Consideration shows that the decay processes are not simpleexponential ones. For the proper choice of QD width, someeigenstates successfully evolve into true BSCs. One can uselinear superposition of the eigenstates of the closed QD insteadof the bare eigenstate at the initial time. This choice will yieldBSCs for the time evolution. This process will be successfulprovided the QD width is chosen so that the equation for theBSC (9) has a solution.

235321-8

Page 9: Formation of bound states in the continuum for a quantum dot with variable width

FORMATION OF BOUND STATES IN THE CONTINUUM . . . PHYSICAL REVIEW B 83, 235321 (2011)

1J. von Neumann and E. Wigner, Phys. Z. 30, 465 (1929).2F. H. Stillinger and D. R. Herrick, Phys. Rev. A 11, 446 (1975).3R. G. Newton, Scattering Theory of Waves and Particles. (Springer,Berlin, 1982).

4T. A. Weber and D. L. Pursey, Phys. Rev. A 50, 4478 (1994).5D. L. Pursey and T. A. Weber, Phys. Rev. A 52, 3932 (1995).6L. S. Cederbaum, R. S. Friedman, V. M. Ryaboy, and N. Moiseyev,Phys. Rev. Lett. 90, 13001 (2003).

7H. Friedrich and D. Wintgen, Phys. Rev. A 31, 3964 (1985).8F. H. Stillinger and T. A. Weber, Phys. Rev. A 10, 1122 (1975).9D. R. Herrik, Physica B 85, 44 (1977).

10F. H. Stillinger, Physica B 85, 270 (1977).11F. Capasso, C. Sirtori, J. Faist, D. L. Sivco, S.-N. G. Chu, and A. Y.

Cho, Nature (London) 385, 565 (1992).12T. V. Shahbazyan and M. E. Raikh, Phys. Rev. B 49, 17123 (1994).13S. Fan, P. R. Villeneuve, J. D. Joannopoulos, M. J. Khan,

C. Manolatou, and H. A. Haus, Phys. Rev. B 59, 15882 (1999).14C. Texier, J. Phys. A: Math. Gen. 35, 3389 (2002).15I. Rotter and A. F. Sadreev, Phys. Rev. E 69, 66201 (2004); 71,

046204 (2005).16M. L. Ladron de Guevara, F. Claro, and P. A. Orellana, Phys. Rev.

B 67, 195335 (2003).17B. Wunsch and A. Chudnovskiy, Phys. Rev. B 68, 245317 (2003).18E. N. Bulgakov, K. N. Pichugin, A. F. Sadreev, and I. Rotter, JETP

Lett. 84, 508 (2006).19M. L. Ladron de Guevara and P. A. Orellana, Phys. Rev. B 73,

205303 (2006).20H. Nakamura, N. Hatano, S. Garmon, and T. Petrosky, Phys. Rev.

Lett. 99, 210404 (2007).21C. S. Kim, A. M. Satanin, Y. S. Joe, and R. M. Cosby, Phys. Rev.

B 60, 10962 (1999).22O. Olendski and L. Mikhailovska, Phys. Rev. B 66, 35331 (2002).23V. Vargiamidis and H. M. Polatoglou, Phys. Rev. B 72, 195333

(2005).24A. F. Sadreev, E. N. Bulgakov, and I. Rotter, Phys. Rev. B 73,

235342 (2006).25A. F. Sadreev, E. N. Bulgakov, and I. Rotter, J. Phys. A 38, 10647

(2005).26K. D. Rowe and P. J. Siemens, J. Phys. A 38, 9821 (2005).27G. Ordonez, K. Na, and S. Kim, Phys. Rev. A 73, 022113 (2006).28K.-K. Voo and C. S. Chu, Phys. Rev. B 74, 155306 (2006).29G. Cattapan and P. Lotti, Eur. Phys. J. B 60, 51 (2007); 66, 517

(2008).30O. Valsson, C.-S. Tang, and V. Gudmundsson, Phys. Rev. B 78,

165318 (2008).

31E. N. Bulgakov and A. F. Sadreev, Phys. Rev. B 78, 075105(2008).

32H.-J. Stockmann, Quantum Chaos: An Introduction (CambridgeUniversity Press, Cambridge, 1999).

33J. Joannopoulos, R. D. Meade, and J. Winn, Photonic Crystals(Princeton University Press, Princeton, 1995).

34M. Muraguchi, K. Shiraishi, T. Okunishi, and K. Takeda, J. Phys.Condens. Matter 21, 064230 (2009).

35F. M. Dittes, Phys. Rep. 339, 215 (2000).36A. F. Sadreev and I. Rotter, J. Phys. A 36, 11413 (2003).37N. Moiseyev, Phys. Rev. Lett. 102, 167404 (2009).38A. Z. Devdariani, V. N. Ostrovsky, and Yu. N. Sebyakin, Sov. Phys.

JETP 44, 477 (1976).39H. Friedrich and D. Wintgen, Phys. Rev. A 32, 3231 (1985).40V. B. Pavlov-Verevkin, Phys. Lett. A 129, 168 (1988).41A. Volya and V. Zelevinsky, Phys. Rev. C 67, 054322 (2003).42A. F. Sadreev, E. N. Bulgakov, K. N. Pichugin, I. Rotter, and

T. V. Babushkina, in Quantum Dots, Research, Technology andApplications, edited by R. W. Knoss (Nova Sciencers Publ.,Hauppauge), pp. 547–577.

43T. Lepetit, E. Akmansoy, J.-P. Ganne, and J.-M. Lourtioz, Phys.Rev. B 82, 195307 (2010).

44M. Miyamoto, Phys. Rev. A 72, 063405 (2005).45C. S. Kim and A. M. Satanin, Phys. Rev. B 58, 15389 (1998).46E. N. Bulgakov and A. F. Sadreev, Phys. Rev. B 81, 115128

(2010).47A. F. Sadreev, E. N. Bulgakov, and I. Rotter, JETP Lett. 82, 498

(2005).48H. Lee and L. E. Reichl, Phys. Rev. B 79, 193305 (2009).49J. W. Gonzalez, M. Pacheco, L. Rosales, and P. A. Orellana,

Europhys. Lett. 91, 66001 (2010).50H. Feshbach, Ann. Phys. (NY) 5, 357 (1958); 19, 287 (1962).51J. Okołowicz, M. Płoszajczak, and I. Rotter, Phys. Rep. 374, 271

(2003).52D. V. Savin, V. V. Sokolov, and H.-J. Sommers, Phys. Rev. E 67,

026215 (2003).53S. Datta, Electronic Transport in Mesoscopic Systems (Cambridge

University Press, 1995).54E. N. Bulgakov, I. Rotter, and A. F. Sadreev, Phys. Rev. A 75,

067401 (2007).55G. B. Akguc and L. E. Reichl, Phys. Rev. E 64, 056221 (2001); 67,

046202 (2003).56E. N. Bulgakov and A. F. Sadreev, JETP Lett. 66, 431 (1997); JETP

87, 1058 (1998).57I. Rotter, J. Phys. A 42, 153001 (2009).

235321-9