Quantum Phononics: Kazutaka Nakamura
Quantum Phononics: Kazutaka Nakamura
Kazutaka Nakamura
Quantum
Phononics
Introduction to Ultrafast Dynamics of
Optical Phonons
Springer Tracts in Modern Physics
Volume 282
Series editors
Yan Chen, Department of Physics, Fudan University, Shanghai, China
Atsushi Fujimori, Department of Physics, University of Tokyo, Tokyo, Japan
Thomas Müller, Institut für Experimentelle Kernphysik, Universität Karlsruhe,
Karlsruhe, Germany
William C. Stwalley, Department of Physics, University of Connecticut, Storrs,
USA
Jianke Yang, Department of Mathematics and Statistics, University of Vermont,
Burlington, VT, USA
Springer Tracts in Modern Physics provides comprehensive and critical reviews of
topics of current interest in physics. The following fields are emphasized:
– Elementary Particle Physics
– Condensed Matter Physics
– Light Matter Interaction
– Atomic and Molecular Physics
– Complex Systems
– Fundamental Astrophysics
Suitable reviews of other fields can also be accepted. The Editors encourage
prospective authors to correspond with them in advance of submitting a manuscript.
For reviews of topics belonging to the above mentioned fields, they should address
the responsible Editor as listed in “Contact the Editors”.
Quantum Phononics
Introduction to Ultrafast Dynamics of Optical
Phonons
123
Kazutaka Nakamura
Laboratory for Materials and Structures,
Institute of Innovative Research
Tokyo Institute of Technology
Yokohama, Japan
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface
Phonon are quantums of lattice vibrations and are bosons as well as photon. The
phonons are nowadays coherently generated and controlled by using an ultrashort
optical pulse. Quantum phononics is a new field in condensed-matter physics and
materials science, which includes the generation, detection, and the engineering and
coherent control of quantum states of lattice vibrations. This book describes the
fundamentals of generation, detection, and coherent control of optical phonons
using ultrashort laser pulses. While they are important for heat transfer and
capacity, acoustic phonons are not included in this book. The required basic
knowledge of quantum mechanics, quantum optics, solid state physics, and non-
linear spectroscopy is presented in Chaps. 1–4, which were developed during a
one-quarter course for graduate students in the department of materials science at
Tokyo Institute of Technology. Chapters 5–7 are based on the author’s recent
journal papers on coherent phonons and their coherent control. I have not tried to be
encyclopedic in the treatment, either in terms of complete references or a discussion
of every type of experiment and theory. Reference are included for more detailed
background or historical interest. Instead, I tried to explain detailed derivations for
important equations, which are often difficult for experimental physicists and
materials scientists.
Chapter 1 reviews the basics of quantum mechanics, state vectors, time evolu-
tion of a quantum state, and perturbation expansion, which are used throughout the
text. In Chap. 2, the density operator and a double-sided Feynman diagram are
introduced, which are used to describe nonlinear optical processes. Chapter 3
explains the quantum mechanics of the harmonic oscillator, which is of essential
importance in field quantization and phonon dynamics. Coherent and squeezed
states, which are important in quantum optics, are introduced. Chapter 4 reviews
lattice vibrations and field quantization and introduces the “phonon”. In Chaps. 5
and 6, we discuss coherent phonons using ultrafast optical measurements and the
quantum mechanical description of the generation and detection mechanism.
Section 5.3 shows squeezed phonons. In Chap. 7, we discuss the coherent control of
optical phonons using an ultrashort pulse train and its selected applications.
v
vi Preface
I especially thank Prof. Yosuke Kayanuma, my co-worker for several years, for
his guidance and advice on theory. I thank my colleagues and students participated
in my research projects on coherent phonons, especially Prof. Masahiro Kitajima,
Prof. Oleg V. Misochko, Prof. Fujio Minami, Prof. Yutaka Shikano, Dr. Yasuaki
Okano, Dr. Hiroshi Takahashi, Dr. Jianbo Hu, and Dr. Katsura Norimatsu. I also
thank Gordon Han Ying Li for checking English and equations in this book.
Immeasurable thanks are due to my wife Mitsuko, without her constant compan-
ionship and support this book would never have been possible. My thanks also go to
Dr. Claus Ascheron, Physics Editor of Springer-Verlag at Heidelberg (now retired),
for his kind support.
vii
viii Contents
where C1 and C2 are complex numbers. When we set another ket vector |ϕ
C1
|ϕ = , (1.2)
C2
1 We consider a dual vector space using ket and bra vector spaces.
© Springer Nature Switzerland AG 2019 1
K. Nakamura, Quantum Phononics, Springer Tracts in Modern Physics 282,
https://doi.org/10.1007/978-3-030-11924-9_1
2 1 Time Evolution of Quantum State
where C1∗ and C2∗ are complex conjugates of C1 and C2 , respectively. The inner
product of their own is
1.2 Operator
An operator  acts on a ket vector from the left side Â|ϕ and the resulting product
is another ket vector. When the ket vector is represented by a N -dimensional vector,
the operator  corresponds to the N × N matrix. When a vector |ϕ satisfies the
following equation
Â|ϕ = a|ϕ, (1.5)
where a is the constant, |ϕ and a are called eigenvector and eigenvalue for Â.
Setting an operator P̂(a) = |aa|, in which |a is normalized, the operator selects
the component of a ket vector (|ϕ) parallel to |a:
where a|ϕ is a classical number. The operator is known as the projection operator
along the base ket vector |a [1]. Using the normalized orthogonal system {|ai }, the
identity operator Iˆ is expressed using the projection operator
Iˆ = P̂(ai ) = |ai ai |. (1.8)
i i
When we operate the identity operator on any operator B̂, the operator does not
change:
B̂ = Iˆ B̂ Iˆ = |ai ai | B̂|a j a j | = ai | B̂|a j |ai a j |. (1.9)
i, j i, j
Thus, any operator can be represented as linear combinations of |ai a j |. In partic-
ular, the Hermitian operator  can be expressed by a linear combination of its own
eigenvectors {|αi }, where Â|αi = αi |αi :
 = Ci |αi |ai αi | = Ci P̂(αi ). (1.10)
i i
2 Dirac wrote in his book “In this way we see that a measurement always causes the system to jump
onto an eigenstate of the dynamical variable that is being measured, the eigenvalue this eigenstate
belongs to being equal to the result of the measurement” [2]. Dirac’s words are also referred to in [1].
4 1 Time Evolution of Quantum State
There are three types of representation for time evolution of a quantum system (equa-
tion of motion in quantum mechanics): Schrödinger, Heisenberg, and Interaction
pictures [3].
where Ĥ is the Hermitian operator representing the total energy of the system. Ĥ is
called the Hamiltonian and is the time-independent operator for the closed system.
When there is no degeneracy, the energy eigenstate |n for the eigenvalue E n satisfies
which is called time-independent Schrödinger equation. When the state |ϕ(t) starts
from the energy eigenstate |n, |ϕ(0) = |n, the state at time t is expressed as
Only the phase factor is different between |ϕ(t) and |ϕ(0). The phase factor rotates
with an angular frequency ωn = E n /. In the closed system, the probability of detect-
ing the energy eigenstate does not change and the energy eigenstate is a steady state.
Since an arbitrary state in the closed system is expanded
using energy eigenstates,
any initial state |ψ(0) is represented by |ψ(0) = n Cn |n. Then the state |ψ(t)
at time t is expressed by
|ψ(t) = Cn e−i En t/ |n = Cn e−iωn t |n. (1.16)
n n
and is constant throughout the time evolution. Thus, the transformation from |ψ(0)
to |ψ(t) is a unitary transformation. We consider the unitary operator Û (t) for the
time evolution,
|ψ(t) = Û (t)|ψ(0). (1.18)
This unitary operator is called the time-evolution operator. Using the Schrödinger
equation, the time-evolution operator satisfies
d
i Û (t) = Ĥ Û (t), (1.19)
dt
with an initial condition Û (0) = Iˆ. If the Hamiltonian is time independent, the time-
evolution operator is expressed by using an exponential form3
−i Ĥ t
Û (t) = exp . (1.20)
d
i Û (t) = Ĥ (t)Û (t). (1.21)
dt
This can be understood as the Hamiltonian is constant during the small period Δt.
1.4 Time Evolution of a Quantum System 7
The linear operator of a physical quantity is time dependent in the Heisenberg picture,
while in the Schrödinger picture, the operator is time independent and the state vector
is time dependent. The expectation value of physical observable  at time t is obtained
in the Schrödinger representation
If we set the time-dependent operator ÂH (t) and the time-independent state vector
|ϕH as
Therefore, the expectation value at time t can be expressed by using the time-
dependent linear operator and the time-independent state vector. In the Heisenberg
picture, the time evolution of ÂH (t) is expressed by4
d
i ÂH (t) = ÂH (t), Ĥ . (1.32)
dt
d
i |ϕI (t) = ĤI (t)|ϕI (t). (1.35)
dt
The solution is formally
i −i
|ϕI (t) = exp Ĥ0 t exp Ĥ t |ϕ(0). (1.36)
d i
ÂI (t) = Ĥ0 , ÂI (t) . (1.37)
dt
The interaction picture can be understood as Ĥ0 being treated in the Heisenberg
picture and the interaction part ĤI in the Schrödinger picture.
ĤI (tn ) ĤI (tn−1 ) · · · ĤI (t1 )|ϕI (t0 ), (1.38)
where the initial state is |ϕI (t0 ). The time-evolution operator from t0 to t is defined
as Û (t, t0 ) ≡ exp(−i Ĥ0 (t − t0 )/). The ket vector in the Schrödinger representation
is expressed as |ϕ(t) = Û (t, t0 )|ϕI (t) and |ϕ(t0 ) = |ϕI (t0 ). Using these relations,
we get
∞
−i n t tn t2
Û (t, t0 )|ϕ(t) = |ϕ(t0 ) +
†
dtn dtn−1 · · · dt1
n=1
t0 t0 t0
ĤI (tn ) ĤI (tn−1 ) · · · ĤI (t1 )|ϕI (t0 ), (1.39)
1.4 Time Evolution of a Quantum System 9
and then
∞
−i n t tn t2
|ϕ(t) = Û (t, t0 )|ϕ(t0 ) + dtn dtn−1 · · · dt1
n=1
t0 t0 t0
By using the relation: Û (tn , tn−1 ) = Û (tn , t0 )Û (t0 , tn−1 ) = Û (tn , t0 )Û † (tn−1 , t0 ), the
above equation can be expressed as
∞
−i n t tn t2
|ϕ(t) = Û (t, t0 )|ϕ(t0 ) + dtn dtn−1 · · · dt1
n=1
t0 t0 t0
This indicates that the system propagates with Û (t1 , t0 ) from t0 under the time-
independent Hamiltonian Ĥ0 until time t1 , when the system interacts with the inter-
action Hamiltonian (perturbation) Ĥ (t1 ). After the interaction, the system propagates
with Û (t2 , t1 ) until time t2 and this process repeats until time t. Û (t, t0 )|ϕ(t0 ) is the
zeroth-order perturbation ket vector, and each term in the summation is the nth-order
one. For example, the second-order perturbation is
2
−i t t2
|ϕ (2) (t) = dt2 dt1
t0 t0
Û (t, t2 ) Ĥ (t2 )Û (t2 , t1 ) Ĥ (t1 )Û (t1 , t0 )|ϕ(t0 ). (1.42)
This is shown graphically in a single-sided Feynman diagram (Fig. 1.1) [4], where
time flows from the left side to the right side.
1.5 Summary
The quantum state is expressed using the ket and bra vectors. Dynamical variables
are represented by linear operators that act on ket vectors. Time evolution of the
quantum state is expressed by using three ways: the Schrödinger, Heisenberg, and
interaction pictures. The time-evolution operator Û (t), which is a unitary operator, is
introduced, and the ket at time t is expressed by |Ψ (t) = Û (t)|Ψ (0). The perturba-
tive expansion is used to express the time evolution of the ket vector in the interaction
picture and is shown graphically using the single-sided Feynman diagram [5].
References
1. Sakurai, J.J.: Modern Quantum Mechanics, Revised edn. Addison-Wesley Publishing Company
Inc., Reading (1994)
2. Dirac, P.A.M.: The Principles of Quantum Mechanics, 4th edn. Oxford University Press, Oxford
(1958)
3. Tannor, D.J.: Introduction to Quantum Mechanics, A Time-Dependent Perspective. University
Science Books, California (2007)
4. Hamm, P.: Principles of Nonlinear Optical Spectroscopy: A Practical Approach. http://www.
mitr.p.lodz.pl/evu/lectures/Hamm.pdf
5. I wrote this chapter referencing Refs. [1–4] and following books: Takahiro Sunagawa and
Masahiro Ueda, Ryoshi Sokutei to Ryoshi Seigyo (in Japanese, Quantum Measurement and
Quantum Control), Science Sya (2016); Yoshio Kuramoto and Junichi Ezawa, Ryoushi Rikigaku
(in Japanese, Quantum Mechanics), Asakura Shyoten (2008); Keiji Igi and Hikari Kawai, Kiso
Ryousi Rikigaku (in Japanese, Fundamental Quantum Mechanics), Kodansya (2007); Masahito
Ueda, Gendai Ryoushi Buturigaku (in Japanese, Modern Quantum Physics), Baifukan (2004);
Akira Shimizu, Shin-han Rryoushi Ron no Kiso (in Japanese, New edition Fundamental of
Quantum Physics), Science Shya (2004)
Chapter 2
Density Operator
This chapter describes the density operator, which is used to calculate light-matter
interaction and the generation of coherent phonons in Chap. 6. The density operator
can represent both a pure quantum state and a mixed state. The von Neumann equation
for the time evolution of the density operator is derived. The perturbative expansion of
its time evolution is explained and represented with double-sided Feynman diagrams.
Time evolution in a two-level system is shown as an example.
A pure state is a state in which a system is defined by a state vector |ϕ with a unit
probability. On the other hand, when the system includes several quantum states |ϕi
with a ratio of pi , thestate is called a mixed state. This state is quite different from a
√
superposition state ( i pi |ϕi ) and cannot be expressed by a ket vector. A mixed
state is expressed by a density operator. The mixed state is important when we treat
a quantum system consisting of many particles. In this system, some particles are
often in different quantum states and the system is expressed with a density operator.
On the other hand, when all particles are in a single state, the system is in a pure state.
When a system consists of state vectors |ϕi with probabilities of pi , the density
operator ρ̂ of this system is defined by
where i pi = 1 [1, 2]. The density operator is expressed by an outer product of ket
and bra vectors, which corresponds to a matrix. The probability that we detect the
system in the state |ϕi is expressed by ϕi |ρ̂|ϕi using the density operator. When a
system is in a pure state and has only one state |ψ, the density operator is
ρ̂ = |ψψ|. (2.2)
The density operator is a Hermitian operator because the probability pi is real and
can be expanded with a orthonormal basis set {|ϕi }. Using an orthonormal basis set,
the trace of an arbitrary operator  is defined by
T r [ Â] ≡ ϕi | Â|ϕi . (2.3)
i
Let us consider a superposition state |φ = i ci |ϕi , where |φ is normalized
and {|ϕi } is a orthonormal basis set. The density operator of this superposition state
is calculated as
ρ̂ = |φφ| = ci c∗j |ϕi ϕ j | = ρ̂ + ci c∗j |ϕi ϕ j |, (2.4)
ij i= j
where ρ̂ is the density operator of the mixed state composed of |ϕi . Then the density
operator of the superposition state has additional off-diagonal terms of ci c∗j |ϕi ϕ j |
compared to that of the mixed state. These off-diagonal terms are due to interfer-
ence between the quantum states. A state, which has such a interference term, has
“coherence”.
Here, we consider, for example, a system including N -particles, in which each
particle has only two levels (a ground state |1 and an excited state |2), and N1 and
N2 particles are in the |1 and |2 states, respectively. Figure 2.1 shows its typical
example. When there is no interaction between each particle, the density operator is
expressed by
N1 N2
ρ̂ = |11| + |22|, (2.5)
N N
where N = N1 + N2 . This is an example of the mixed state. On the other √ hand, in
a√pure state, all particles are expressed by the single ket vector |ψ = N1 /N |1 +
N2 /N |2 and the density operator is expressed by
N1 N2 N1 N2
ρ̂ = |ψψ| = |11| + |22| + (|12| + |12|) . (2.6)
N N N2
The last term shows an interference between two states and the coherence.
2.2 Density Operator 13
Since |ϕi (t) obeys the Schrödinger equation, the time evolution of the density oper-
ator is calculated by
d d
ρ̂(t) = pi |ϕi (t)ϕi (t)|
dt i
dt
d|ϕi (t) dϕi (t)|
= pi ϕi (t)| + |ϕi (t)
i
dt dt
i i
= pi − Ĥ |ϕi (t)ϕi (t)| + |ϕi (t)ϕi (t)| Ĥ
i
1
= Ĥ , ρ̂(t) , (2.8)
i
1 This equation corresponds to the Liouville equation for classical mechanics. Then, this is also
called the quantum Liouville equation or Liouville–von Neumann equation [2].
14 2 Density Operator
When the Hamiltonian is separated into two parts as Ĥ = Ĥ0 + Ĥ (t), where Ĥ0
is time independent, the density operator can be treated with the interaction picture.
The density operator ρ̂I (t) and the interaction ĤI (t) in the interaction picture are
given by
i Ĥ0 t −i Ĥ0 t
ρ̂I (t) = exp ρ̂(t) exp
i Ĥ0 t −i Ĥ0 t
ĤI (t) = exp Ĥ (t) exp , (2.9)
d 1
ρ̂I (t) = ĤI , ρ̂I (t) . (2.10)
dt i
The von Neumann equation in the interaction picture is formally the same as that in
the Schrödinger picture.
1 t
ρ̂(t) = ρ̂(0) + Ĥ (t), ρ̂(0) dt. (2.11)
i 0
1 t
ρ̂ (1) (t) = Ĥ (t1 ), ρ̂(0) dt1 . (2.12)
i 0
This corresponds to that the interaction occurs once at time t1 . Inserting ρ̂ (1) (t) into
the integral, we get the second-order approximation
2 t t2
1
ρ̂ (2) (t) = dt2 dt1 Ĥ (t2 ), Ĥ (t1 ), ρ̂(0) . (2.13)
i 0 0
Repeating this procedure one after another, we get the nth-order approximation
2.2 Density Operator 15
n t tn t2
1
ρ̂ (n) (t) = dtn dtn−1 dt1 · · ·
i 0 0 0
Ĥ (tn ), Ĥ (tn−1 ), . . . Ĥ (t1 ), ρ̂(0) . . . . (2.14)
Finally, we get
∞
ρ̂(t) = ρ̂ (n) (t). (2.15)
n=0
which indicates that |ϕ(0) and ϕ(0)| get interaction once at time t1 . In a similar
way, the second-order approximation includes
Ĥ (t2 ), Ĥ (t1 ), ρ̂(0) = Ĥ (t2 ) Ĥ (t1 )|ϕ(0)ϕ(0)| − Ĥ (t2 )|ϕ(0)ϕ(0)| Ĥ (t1 )
which consists of the ket having interaction at t1 and t2 , the ket having interaction at
t2 with the bra having interaction at t1 , the ket having interaction at t2 with the bra
having interaction at t1 , and the bra having interaction at t1 and t2 .
The von Neumann equation in the interaction picture is formally equivalent to that in
the Schrödinger picture. Then, the density operator in the interaction picture is also
expressed as
∞
−i n t tn t2
ρ̂I (t) = ρ̂I (t0 ) + dtn dtn−1 · · · dt1
n=1
t0 t0 t0
ĤI (tn ), ĤI (tn−1 ), . . . , ĤI (t1 ), ρ̂I (t0 ) . . . . (2.18)
From this equation, the density operator in the Schrödinger picture is expressed as
∞
−i n t tn t2
ρ̂(t) = Û (t, t0 )ρ̂(t0 )Û (t, t0 ) +
†
dtn dtn−1 · · · dt1
n=1
t0 t0 t0
Û (t, t0 ) ĤI (tn ), ĤI (tn−1 ), . . . , ĤI (t1 ), ρ̂(t0 ) . . . Û † (t, t0 ), (2.19)
16 2 Density Operator
where we used ρ̂I (t0 ) = ρ̂(t0 ). The first term is the zeroth-order perturbation and the
time-evolution operator Û (t, t0 ) and its Hermitian conjugate operator act on ket and
bra vectors, respectively. The first-order perturbation is obtained as
−i t
ρ̂ (1) (t) = dt1 Û (t, t0 ) ĤI (t1 ), ρ̂(t0 ) Û † (t, t0 ). (2.20)
t0
= Û (t, t0 )Û † (t1 , t0 ) Ĥ (t1 )Û (t1 , t0 )|ϕ(t0 )ϕ(t0 )|Û † (t, t0 )
−Û (t, t0 )|ϕ(t0 )ϕ(t0 )|Û † (t1 , t0 ) Ĥ (t1 )Û (t1 , t0 )Û † (t, t0 )
= Û (t, t1 ) Ĥ (t1 )Û (t1 , t0 )|ϕ(t0 )ϕ(t0 )|Û † (t, t0 )
−Û (t, t0 )|ϕ(t0 )ϕ(t0 )|Û † (t1 , t0 ) Ĥ (t1 )Û † (t, t1 ). (2.21)
The first term shows that the ket propagates under Ĥ0 until time t1 , interacts with
perturbation Ĥ (t1 ) at time t1 , and propagates again under Ĥ0 and the bra propagates
under Ĥ0 without perturbation. In the second term, the perturbative interaction occurs
for the bra. This interpretation is expressed by using double-sided Feynman diagrams
(Fig. 2.2). The upper and lower lines show propagation of the ket and bra vectors,
respectively. Time is running from the left to the right.
Similar calculation gives the second-order perturbation as
2
−i t t2
ρ̂ (2) (t) = dt2 dt1 Û (t, t0 ) ĤI (t2 ), ĤI (t1 ), ρ̂(t0 ) Û † (t, t0 )
t0 t0
2.2 Density Operator 17
2
−i t t2
= dt2 dt1
t0 t0
(Û (t, t2 ) Ĥ (t2 )Û (t2 , t1 ) Ĥ (t1 )Û (t1 , t0 )|ϕ(t0 )ϕ(t0 )|Û † (t, t0 )
− Û (t, t2 ) Ĥ (t2 )Û (t2 , t0 )|ϕ(t0 )ϕ(t0 )|Û † (t1 , t0 ) Ĥ (t1 )Û † (t, t1 )
− Û (t, t1 ) Ĥ (t1 )Û (t1 , t0 )|ϕ(t0 )ϕ(t0 )|Û † (t2 , t0 ) Ĥ (t2 )Û † (t, t2 )
+ Û (t, t0 )|ϕ(t0 )ϕ(t0 )|Û † (t1 , t0 ) Ĥ (t1 )Û † (t2 , t1 ) Ĥ (t2 )Û † (t, t2 )).
(2.22)
The perturbative interaction occurs two times for the ket or the bra in the first or
the fourth term, respectively, and occurs one time for both the ket and bra in the
second and third terms. The double-sided Feynman diagrams for the second-order
perturbation is shown in Fig. 2.3, in which (a)–(d) correspond each term in the (2.22).
18 2 Density Operator
The perturbative expansion of the time evolution of the density operator is expressed
using double-sided Feynman diagrams. Here, we list several rules for the double-
sided Feynman diagram [3].
• Time is running from the left to the right.
• The double-sided Feynman diagram consists of two lines which expressed time
evolution of ket and bra vectors.
• Interaction with the interaction Hamiltonian Ĥ (t) is represented by arrows.
• Each diagram has a sign (−1)n , where n is the number of interaction to the bra
vector (arrows to the bottom line).
where ε1 and ε2 are the energy and ε1 < ε2 . The Hamiltonian is also expressed using
the matrix:
ε1 0
Ĥ = . (2.24)
0 ε2
The time evolution of the density operator, which is shown in the matrix form, is
obtained from the von Neumann equation
d ρ11 ρ12 −i 0 (ε1 − ε2 )ρ12
= , (2.25)
dt ρ21 ρ22 (ε2 − ε1 )ρ21 0
for the interaction-free case. The diagonal components, which correspond to popu-
lations, are constant (ρ11 (t) = ρ11 (0) and ρ22 (t) = ρ22 (0)) in time. The off-diagonal
components, which correspond to polarizations, oscillate
2.3 Density Operator for a Two-Level System 19
is applied to the two-level system, the total Hamiltonian ( Ĥ (t) = Ĥ0 + V̂ (t)) is
expressed by
ε1 γ eiωt
Ĥ (t) = . (2.30)
γ e−iωt ε2
2 This interaction corresponds to an dipole interaction between the two-level system and light with
an angular frequency of ω. When we use μ as a dipole moment and E 0 as an amplitude of elec-
tric field, γ = μE 0 . γ e−iωt |21| and γ eiωt |12| correspond to the excitation process with light
absorption and the de-excitation process with emission. In addition, the rotating wave approximation
is assumed.
20 2 Density Operator
The sum of the population in |1 and |2 is time independent, because ρ˙11 + ρ˙22 = 0.
From the results of the interaction-free case, the off-diagonal terms have e−iω21 t or
eiω21 t. Then, we use a rotating frame defined as ρ̃12 = e−iω21 t ρ12 , ρ̃21 = eiω21 t ρ21 ,
ρ̃11 = ρ11 , and ρ̃22 = ρ22 . In the rotating frame, the density matrix elements are
For the resonant condition (ω − ω21 = 0), the equations are greatly simplified, and
we get
and
If we put ρ̃11 − ρ̃22 = Ceat , we get a = ±i2Ω. For the initial condition of ρ11 (0) = 1
and ρ22 (0) = 0, the coefficient C is obtained to be 1/2. Then, we get
ei2Ωt + e−i2Ωt
ρ11 − ρ22 = = cos 2Ωt. (2.36)
2
By using ρ11 + ρ22 = 1, we finally get
1 − cos 2Ωt
ρ11 = = cos2 Ωt. (2.37)
2
The result shows that the system absorbs energy from the interaction potential V̂ (t)
in the time range between 0 and Ω/2 and emits in the time range between Ω/2 and
Ω (as shown in Fig. 2.4). This absorption and emission process periodically repeat
with a frequency of Ω, which is called Rabi frequency.
2.3 Density Operator for a Two-Level System 21
Population
0.6
0.4
0.2
0.0
0 1 2 3 4
time (π/Ω)
When we introduce the phase relaxation (Γ ) in this two-level system, ρ˙˜12 and ρ˙˜21
are obtained by
If we represent the relaxation rate in units of the Rabi oscillation (Γ = nΩ), the
population ρ11 is a damped oscillation
1
ρ11 = 1 + e−nΩt cos( 4 − n 2 Ωt) , (2.40)
2
if 0 ≤ n ≤ 2 and
1
ρ11 = 1 + e−nΩt cosh( n 2 − 4Ωt) , (2.41)
2
if n > 2.
The time evolution of the population (ρ11 and ρ22 ) is shown in Figs. 2.5 and
2.6 for Γ = 0.5 and Γ = 2.5 , respectively. Both ρ11 and ρ22 approach to the
steady-state value (1/2).
22 2 Density Operator
Population
dotted curves show ρ11 and 0.6
ρ22 , respectively
0.4
0.2
0.0
0 1 2 3 4
time (π/Ω)
0.2
0.0
0 1 2 3 4
time (π/Ω)
2.4 Summary
The density operator, which can describe a pure state and mixed state, is described.
The time evolution of the density matrix is obtained by the von Neumann equation.
The perturbative expansion of the time evolution of the density operator is explained
and represented by using a double-sided Feynman diagram. The time evolution of
the populations in the excited and ground states in the two-level model is calculated
[4].
References
1. Schatz, G.C., Ratner, M.A.: Quantum Mechanics in Chemistry. Dover Publication Inc., Mineola
(2002)
2. Tannor, D.J.: Introduction to Quantum Mechanics, A Time-Dependent Perspective. University
Science Books, California (2007)
References 23
3. Mukamel, S.: Principles of Nonlinear Optical Spectroscopy. Oxford University Press, New York
(1995)
4. I wrote this chapter referencing Refs. [1–3] and following books: Yoshio Kuramoto and Junichi
Ezawa, Ryoushi Rikigaku (in Japanese, Quantum Mechanics), Asakura Shyoten (2008); Keiji
Igi and Hikari Kawai, Kiso Ryousi Rikigaku (in Japanese, Fundamental Quantum Mechanics),
Kodansya (2007); Takahiro Sunagawa and Masahiro Ueda, Ryoshi Sokutei to Ryoshi Seigyo (in
Japanese, Quantum Measurement and Quantum Control), Science Sya (2016), Masahito Ueda
Gendai Ryoushi Buturigaku (in Japanese, Modern Quantum Physics), Baifukan (2004); Akira
Shimizu, Shin-han Rryoushi Ron no Kiso (in Japanese, New edition Fundamental of Quantum
Physics), Science Shya (2004); Kyo Inoue, Kogaku Kei no Tameno Ryoushi Kogaku (in Japanese,
Quantum Optics for Engineer), Morikita Shyoten (2015); Masahiro Matsuoka, Ryoushi Kogaku
(in Japanese, Quantum Optics), Shokabou (2000)
Chapter 3
Harmonic Oscillator and Coherent
and Squeezed States
3.1.1 Hamiltonian
p2 mω2 x 2
H= + , (3.1)
2m 2
where x and p are position and momentum, respectively (Fig. 3.1).
Replacing
x and p to linear operators x̂ and p̂, which satisfy commutation relation
x̂, p̂ = i, we get the quantum mechanical Hamiltonian Ĥ by
p̂ 2 mω2 x̂ 2
Ĥ = + . (3.2)
2m 2
Here, we introduce new operators â and â † [1], which are defined using x̂ and p̂ by
1
â ≡ √ mω x̂ + i p̂ ,
2mω
1
â † ≡ √ mω x̂ − i p̂ . (3.3)
2mω
The position and momentum operators (x̂ and p̂) are expressed by
x̂ = â + â † ,
2mω
mω
p̂ = i −â + â † , (3.4)
2
using
the new operators. Inserting this relationship into the commutation relation
x̂, p̂ , we get
x̂, p̂ = x̂ p̂ − p̂ x̂
−i
= (â + â † )(â − â † ) − (â − â † )(â + â † )
2
= −i â † â − â â †
= i â, â † . (3.5)
n̂ ≡ â † â, (3.8)
1
Ĥ = ω n̂ + , (3.9)
2
where n and |n are eigenvalue and eigenstate, respectively. Operating â on (3.10)
from the left side, the left-hand side of the equation is
â n̂|n = â â † â|n = â † â + 1 |n = n̂ + 1 |n, (3.11)
Then, we get
n̂ + 1 |n = n â|n, (3.13)
and
This equation means that â|n is also an eigenstate of the operator n̂ and its eigenvalue
is n − 1. We named an eigenstate belonging to the eigenvalue n |n, then an eigenstate
belonging to the eigenvalue n − 1 is |n − 1. The state â|n is expressed by
where Cn is a constant. When we calculate the inner product with itself, the left-hand
side is
and
m 0 +1)
â |n = 0. (3.21)
1 1
Ĥ |n = ω n̂ + |n = ω n + |n, (3.24)
2 2
The harmonic oscillator has energy of ω/2 even at the lowest state |0, which is
called the zero-point energy, and a constant energy spacing of ω. The eigenstate
|n is called an occupation number state,3 because well-defined numbers of quanta
(ω) are included in the state. |0 is the lowest energy state and called the ground
state or the vacuum state.
The operator â shifts a number state |n to the lower state |n − 1, and is called the
annihilation operator [1]. On the other hand, its conjugated operator â † is the creation
operator, which shifts the state to the upper state |n + 1. Operating â † to (3.10) from
the left side, the left-hand side is
â † n̂|n = â † â † â|n = â † â â † − 1 |n = n̂ − 1 â † |n (3.26)
Then, we get
2m is, of course, a nonnegative integer, because it is number of times for the operation.
3 The occupation number state is also called the Fock state for photons in quantum optics.
30 3 Harmonic Oscillator and Coherent and Squeezed States
where Cn is a normalized constant. When we calculate the inner product with itself,
the left-hand side is
Then, we get
√
â † |n = n + 1|n + 1. (3.32)
â|0 = 0. (3.33)
Operating â † n times to the lowest state |0, we get upper number states one by one
as [1]
|1 = â † |0
1 1
|2 = √ â † |1 = √ (â † )2 |0
2 2·1
1 1
|3 = √ â † |2 = √ (â † )3 |0
3 3·2·1
..
.
1 1
|n = √ â † |n − 1 = √ (â † )n |0. (3.34)
n n!
The annihilation
and creation
operators
obey commutation relations with the number
operator as â, n̂ = â and â † , n̂ = â † .
The wave function ϕ0 (x) of the lowest state |0 is obtained in a position space by
solving (3.33).4 In the one-dimensional position space, the position and momentum
operators are expressed as
4 The wave function ϕ( x) is obtained by projection of the ket vector |ϕ into position space and has
the relationship ϕ( x) = x|ϕ.
3.1 Hamiltonian and Energy Eigenstate 31
x̂ = x
d
p̂ = −i , (3.35)
dx
and the annihilation operator is represented by
1 d
â = √ mωx + . (3.36)
2mω dx
d
mωx + ϕ0 (x) = 0, (3.37)
dx
−mωx 2
ϕ0 (x) = C exp . (3.38)
2
and obtained as
mω
C= . (3.40)
π
Then, the wave function ϕ0 (x) is a Gaussian function centered at x = 0. The wave
functions for the upper number states are obtained by operating the creation operator
n=1
n=0
x
32 3 Harmonic Oscillator and Coherent and Squeezed States
â † one by one to ϕ0 (x). For example, the wave function ϕ1 (x) for |1 is
1 d mω −mωx 2
ϕ1 (x) = â ϕ0 (x) = √
†
mωx − exp
2mω dx π 2
2m −mωx 2
= √ x exp . (3.41)
2π 2
The wave function ϕ1 (x) is a odd function and has a node at x = 0. The ϕ2 (x) is
also calculated by â † ϕ1 (x). The wave functions ϕ0 (x), ϕ1 (x), and ϕ1 (x) are shown
in Fig. 3.2.
1
|n(t) = exp −i n + ωt |n, (3.42)
2
because its energy is ω(n + 1/2). Of course, the probability distribution function
of the number state does not change in time n(t)|n(t) = n|n, because it is a
stationary state. The expectation value of position for the number state is calculated
to be zero:
x̂ = n(t)|x̂|n(t) = n| (â + â ) |n = 0.
†
(3.43)
2mω
When the position of the harmonic oscillator is measured, the observed value is scat-
tered at every observation and its standard deviation is σ , with an average value of
zero. This feature is different from the characteristics of classical harmonic oscil-
lators, in which the position changes in time with its angular frequency ω. Such
oscillation of position can be realized in a superposition state of adjoining number
states. Suppose the superposition state of |0 and |1 as
1 ω
|ϕ(t) = √ |0 + e−iωt |1 e−i 2 t . (3.47)
2
1
x 2 = 0| + eiωt 1| 2n̂ + 1 |0 + e−iωt |1
2 2mω
= , (3.49)
mω
and the standard deviation is
σ = − cos2 (ωt)
mω 2mω
= (1 + sin2 (ωt)). (3.50)
2mω
Then, the standard deviation for position also oscillates, which means the shape of
the probability distribution function changes in time.
On the other hand, a superposition state composed of number states, which differ
more than two, shows no oscillation in expectation value of position, because the
position operator includes only â and â † . Suppose the superposition state of |0
and |2 are
1 ω
|ϕ(t) = √ |0 + e−i2ωt |2 e−i 2 t . (3.51)
2
34 3 Harmonic Oscillator and Coherent and Squeezed States
1 2
x 2 = 0| + ei2ωt 2| â + (â † )2 + 2n̂ + 1 |0 + e−i2ωt |2
2 2mω
√
= 3 + 2 cos(2ωt) , (3.53)
2mω
and the standard deviation is the same value
√
σ = 3 + 2 cos(2ωt) . (3.54)
2mω
Therefore, in the superposition state, the mean value does not change and only the
standard deviation oscillates with 2ω.
A coherent state5 is one of the most important states, which is used in quantum
optics and quantum information technologies [4]. The coherent state is defined by
an eigenstate of the annihilation operator â. Consider an eigenvalue and eigenstate
of â as α and |α
5 The concept of the coherent state is first proposed by Schrödinger [3]. Much work on the coherent
state in quantum optics was performed by R. J. Glauber. The coherent state is also called the Glauber
state in quantum optics.
3.3 Coherent States 35
then, we get
√
n|â|α = n + 1n + 1|α. (3.59)
Then, we get
αn
|α = 0|α √ |n, (3.62)
n n!
|α|2n
= |0|α|2 e|α| = 1
2
|α = |0|α|2 (3.63)
n
n!
requires
|0|α|2 = e−|α| .
2
(3.64)
where the global phase is set to be 1. The coherent state is the superposition state of
an infinite number of the number states. When we measure the harmonic oscillator
in the coherent state, the probability to find it in the eigenstate |n is
36 3 Harmonic Oscillator and Coherent and Squeezed States
Population Pn
0.6
0.4
0.2
0.0
0 4 8 12
n
(b) 1.0
0.8
Population Pn
0.6
0.4
0.2
0.0
0 4 8 12
n
|α|2n
P(n) = |n|α|2 = e−|α|
2
, (3.66)
n!
which is a Poisson distribution.6 Then, the the coherent state is the superposition state
of energy eigenstates of the harmonic oscillator with a Poisson distribution. Typical
examples of the population of the number states in the coherent state is shown in
Fig. 3.3.
Here, we describe the uncertainty relation between position and momentum in the
coherent state of the harmonic oscillator. The expectation values for position and
6 ThePoisson distribution P(n) = e−λ λk /k! represents the probability of events occurring k times
per unit time, when the events occur randomly λ per unit time.
3.3 Coherent States 37
momentum are
x = α|x̂|α = α|(â + â )|α =
†
(α + α ∗ )
2mω 2mω
mω mω
p = α| p̂|α = −i α|(â − â )|α = −i
†
(α − α ∗ ), (3.67)
2 2
x 2 = α|x̂ 2 |α = α|(â + â † )(â + â † )|α
2mω
= α|(â 2 + (â † )2 + 2â † â + 1)|α = (α + α ∗ )2 + 1
2mω 2mω
mω
p 2 = α| p̂ 2 |α = − α|(â − â † )(â − â † )|α
2
mω
=− (α − α ∗ )2 − 1 . (3.68)
2
Then, we get the variance
( x)2 =
2mω
mω
( p) =
2
, (3.69)
2
and the uncertainty relation is given by
( x)2 ( p)2 = . (3.70)
2
It shows the coherent state is a minimum uncertainty state for position and
momentum.
Here, we describe that the coherent state is also derived using a displaced harmonic
oscillator. Consider the initial state as the ground state |0 of the harmonic oscillator,
in which the potential minimum is at x = 0 position (schematically shown in Fig. 3.4).
When this harmonic potential is suddenly shifted α along the x-coordinate, the initial
state transfers to a new state. This operation is described by a displacement operator
D̂(α):
This operator is an unitary operator and satisfies D̂ † (α) = D̂(−α). The displacement
operator shifts the annihilation operator by α:
∞
αn
|α(t) = e−i Ĥ t/ |α = 0|α √ e−i Ĥ t/ |n
n=0 n!
3.3 Coherent States 39
∞
αn
= 0|α √ e−i(n+1/2)ωt/ |n
n=0 n!
∞
(αe−iωt â † )n
= 0|αe−iωt/2 |0
n=0
n!
= e−|α| /2 −iωt/2
exp αe−iωt â † |0.
2
e (3.74)
Comparing this equation with the time-independent coherent state expanded in the
number states
|α = e−|α| /2
2
exp α â † |0, (3.75)
we know that α is replaced by αe−iωt and the phase factor e−iωt/2 is multiplied in
the time-dependent coherent state. Then, we can rewrite as
for a real α value. Then, the expectation value of position oscillates with the angular
frequency ω. As shown in (3.69), the variances of both position and momentum do
not depend on α and are time independent. This implies that the shape of the wave
packet of the coherent state does not change in time. Then, the coherent state is
similar to a classical harmonic oscillator.
The coherent state is a minimum uncertainty state for position and momentum
(ΔxΔp = /2). In a squeezed state, one of these two variances is reduced while
maintaining the minimum uncertainty state [4]. We define new operators (b̂ and
b̂† ) using Bogoliubov transformation to the annihilation and creation operators (â
and â † ) as
Using eγ defined by
γ mω
e ≡ , (3.81)
mω
Therefore, b̂ and b̂† are considered to be annihilation and creation operators for the
harmonic oscillator ( Ĥb = ω (b̂† b̂ + 1/2)) defined by the mass m and the angular
frequency of ω [5]. A sharpened potential for the harmonic oscillator is schematically
shown in Fig. 3.5. We set the ground state (vacuum state) of this Hamiltonian as |0b ,
which satisfies b̂|0b = 0. At this state, Δx and Δp are obtained as
−γ
Δx = = e
2m ω 2mω
m ω mω γ
Δp = = e . (3.83)
2 2
This state is the squeezed vacuum state, which is also a minimum uncertainty state
(ΔxΔp = /2) and Δx is reduced compared to that of the ground state in the original
harmonic oscillator with (m, ω) for a positive γ value.
|0b is expanded with number states of Ĥ :
∞
|0b = Cn |n. (3.84)
n=0
3.4 Squeezed States 41
Then, the coefficient of |n should be zero and we get the recursion relation of
ν n
Cn+1 =− Cn−1 . (3.86)
μ n+1
The |0b contains only even or odd number states. We choose the even solution
because the ground state |0 is included. When we start from C0 , C2m is obtained by
−ν 1
C2 = C0 ,
μ 2
2
−ν 3 −ν 3·1
C4 = C2 = C0
μ 4 μ 4·2
3
−ν 5 −ν 5·3·1
C6 = C4 = C0
μ 6 μ 6·4·2
..
.
m
−ν 2m − 1 −ν (2m − 1)!!
C2m = C2(m−1) = C0
μ 2m μ 2m!!
(2m − 1)!!
= (−1)m (tanh γ )2m C0 , (3.87)
(2m)!!
42 3 Harmonic Oscillator and Coherent and Squeezed States
which leads to
∞
(tanh γ )2m (2m − 1)!!
|C0 | 2
1+ = 1. (3.89)
m=1
(2m)!!
we get
∞
(2m − 1)!! − 1
1+ (tanh2 γ )m = 1 − tanh2 γ 2 , (3.91)
m=1
(2m)!!
where (2m − 1)!! = (2m)!/(2m m!). The population of the number state is
Population Pn
0.15
0.10
0.05
0.00
0 5 10 15
n
(b)
-3
4x10
Population Pn
0
0 5 10 15
n
The squeezed vacuum state |0b consists of only even number states. A typical
example of the population of the number states in the squeezed states is shown
in Fig. 3.6.
Next, we consider time evolution of the squeezed vacuum state. Since the time
evolution of the even number states are expressed by exp(−iω(2m + 1/2)t)|2m,
the time-dependent squeezed vacuum state |0(t)b is
∞ √
1 (2m)!
|0(t)b = √ (−1) m
(tanh γ e−i2ωt )m |2m, (3.95)
cosh γ m=1 2m m!
where the phase exp(−iωt/2) is omitted. This means that tanh γ should be replaced
by tanh γ e−i2ωt . Then, we need to set sinh γ → sinh γ e−i2ωt , which corresponds to
ν = e−i2ωt sinh γ . Here, we introduce a squeeze operator Ŝ(ξ ) as
1 ∗ 2
Ŝ(ξ ) ≡ exp ξ â − ξ â †2 , (3.96)
2
44 3 Harmonic Oscillator and Coherent and Squeezed States
(iλ)2
eiλ Â B̂e−iλ Â = B̂ + iλ[ Â, B̂] + [ Â, [ Â, B̂]] + · · · , (3.98)
2!
i 1 ∗ 2
 = ξ â − ξ â †2
λ2
B̂ = â, (3.99)
we get
i
[ Â, B̂] = [(ξ ∗ â 2 − ξ â †2 ), â]
2λ
i i
= (−ξ )[â †2 , â] = (2ξ )â † . (3.100)
2λ 2λ
position (x̂) and momentum ( p̂). At first we calculate the expectation values for the
creation and annihilation operators.
(b)
−2γ
(Δx)2 = (cosh γ − sinh γ )2 = e
2mω 2mω
mω mω 2γ
(Δp)2 = (cosh γ + sinh γ )2 = e , (3.113)
2 2
and squeezing exists in the x̂ quadrature for γ > 0. On the other hand, for θ =
π , squeezing exists in the p̂ quadrature. Figure 3.7 shows the error ellipse for the
squeezed states.
For an arbitrary angle θ , we use a rotational transformation to operators (x̂ and
p̂) and new operators (x̂2 and p̂2 ):
Then, we get x̂2 = e−iθ/2 x̂ and p̂2 = e−iθ/2 p̂, and the creation and annihilation
operators for the new coordinates (â2 and â2† ) are defined by â2 = e−iθ/2 â and
â2† = e−iθ/2 â † .
The squeeze operator is expressed using â2 and â2† as
1 −iθ 2
Ŝ(ξ ) = Ŝ(r eiθ ) = exp r e â − r eiθ â †2
2
1 −iθ/2 2
= exp r (e â) − r (eiθ â † )2
2
1
= exp r â22 − r â2†2 . (3.116)
2
This means that the squeeze operator acts with θ = 0 to x̂2 and p̂2 . Then, the variance
is obtained as
−2γ
(Δx2 )2 = e
2mω
mω 2γ
(Δp2 )2 = e , (3.117)
2
2
Δx)
π/ω)
and the squeeze occurs in the x2 direction. Figure 3.8 shows the error ellipse at
θ = π/4.
Time evolution of variance is calculated using (3.113) and replacing θ by 2ωt:
(Δx)2 = ξ |x̂ 2 |ξ = cosh2 γ + sinh2 γ − 2 cos(2ωt) cosh γ sinh γ
2mω
mω
(Δp) = ξ | p̂ |ξ =
2 2
cosh2 γ + sinh2 γ + 2 cos(2ωt) cosh γ sinh γ .
2
(3.118)
Then, the variance in the squeezed vacuum state oscillates with the angular frequency
of 2ω as shown in Fig. 3.9, though the variance is constant for the coherent state.
The expectation value of the annihilation operator for the state is calculated as
Using these equations, we get the expectation value of the position and the momen-
tum as
x̂ = α, ξ |x̂|α, ξ = α, ξ |(â + â † )|α, ξ
2mω
= (α + α ∗ )
2mω
mω
p̂ = α, ξ | p̂|α, ξ = −i α, ξ |(â − â † )|α, ξ
2
mω
= −i (α − α ∗ ), (3.122)
2
which are the same as those of the coherent states (3.67). Similarly, we get
x̂ 2 = α, ξ |(â â + â † â + â â † + â † â † )|α, ξ
2mω
= ξ |x̂ 2 |ξ + (α 2 + α ∗2 + 2|α|2 )
2mω
mω
p̂ 2 = − α, ξ |(â â + â † â − â â † − â † â † )|α, ξ
2
mω 2
= ξ | p̂ 2 |ξ − (α + α ∗2 − 2|α|2 ). (3.123)
2
The variances are obtained as
and the same as those of the squeezed vacuum state (3.113). Therefore, in the general
squeezed state |α, ξ , the mean value of the position is the same as that of the coherent
state |α and the variance is the same as that of the squeezed vacuum state |ξ . The
time evolution of the position x̂ and the variance (Δx)2 oscillates with the frequency
of ω and 2ω, respectively.
3.5 Summary
|n. The coherent state, which is a minimum uncertainty state, is introduced. The
squeezed state, in which the variance of the position and the momentum is reduced
while maintaining the minimum uncertainty state, is also introduced. In the coherent
state, the mean value of the position oscillates with the angular frequency ω with
keeping its variance constant. In the vacuum squeezed state, the mean value of the
position is constant and its variance oscillates with 2ω [6].
References
1. Sakurai, J.J.: Modern Quantum Mechanics, Revised edn. Addison-Wesley Publishing Company
Inc., Reading (1994)
2. Takahashi, Y.: Ba no Ryoshiron I (in Japanese: Quantum Field Theory I). Baifukan, Tokyo
(1974)
3. Schrödinger, E.: Der steige Ubergang von der Mikro- zur Makromechanik. Die Nautureis-
senschaften 14, 664–666 (1926)
4. Gerry, C.C., Knight, P.L.: Introductory Quantum Optics. Cambridge University Press, New York
(2008)
5. Zwiebach, B.: Quantum dynamics (2013). https://ocw.mit.edu/courses/physics/8-05-quantum-
physics-ii-fall-2013/lecture-notes/MIT8_05F13_Chap_06.pdf
6. I wrote this chapter referencing Refs. [1–5] and following books: Yoshio Kuramoto and Junichi
Ezawa, Ryoushi Rikigaku (in Japanese, Quantum Mechanics), Asakura Shyoten (2008); Keiji
Igi and Hikari Kawai, Kiso Ryousi Rikigaku (in Japanese, Fundamental Quantum Mechanics),
Kodansya (2007); Masahito Ueda Gendai Ryoushi Buturigaku (in Japanese, Modern Quantum
Physics), Baifukan (2004); Akira Shimizu, Shin-han Rryoushi Ron no Kiso (in Japanese, New
edition Fundamental of Quantum Physics), Science Shya (2004); Kyo Inoue, Kogaku Kei no
Tameno Ryoushi Kogaku (in Japanese, Quantum Optics for Engineer), Morikita Shyoten (2015);
Masahiro Matsuoka, Ryoushi Kogaku (in Japanese, Quantum Optics), Shokabou (2000)
Chapter 4
Lattice Vibration and Phonon
Suppose atoms with mass of m stay in line with spacing a.1 The displacement of the
lth atom from its equilibrium position is set to be q l as shown in Fig. 4.1. The spring
constant between atoms is f . The equation of motion of this atom is
1 Here,we referred to the method described in [1] for classical and quantum mechanics of a linear
atomic chain.
© Springer Nature Switzerland AG 2019 51
K. Nakamura, Quantum Phononics, Springer Tracts in Modern Physics 282,
https://doi.org/10.1007/978-3-030-11924-9_4
52 4 Lattice Vibration and Phonon
Fig. 4.1 Schematic of the linear atomic chain. The unit cell with the lattice constant a contains one
atom with mass of m. The force constant is f , and q l is atomic displacement in the lth unit cell
1
q l (t) = √ eikla Q k (t), (4.2)
N
N
1 ikla ∗ 1 ik la
√ e √ e = δk,k . (4.4)
l=1
N N
where D = f /m. This is the equation of simple harmonic oscillation for Q k , and
then we express Q k as
Then we get a dispersion relation between the frequency (ω) and the wave vector (k)
as
√
ωk = 2 D| sin(ka/2)|. (4.8)
The dispersion curve is shown in Fig. 4.2. The frequency ωk is equal to ω−k .
4.1 Linear Atomic Chain 53
1.5
1.0
0.5
0.0
-1.0 -0.5 0.0 0.5 1.0
Wave number (π/a)
1 1
q l (t) = √ eikla e−iωk t Ak + √ e−ikla eiωk t A∗k
k
N N
1 1
= √ eikla Q k (t) + √ e−ikla Q ∗k (t) . (4.9)
k
N N
pl (t)2 2
f l
H= + q (t) − q l+1 (t) , (4.10)
l
2m 2 l
1
pl (t) = m q̇ l (t) = m √ eikla Q̇ k (t) + c.c.
k
N
1
= −iωk m √ eikla Q k (t) + c.c. . (4.11)
k
N
Here c.c. means the complex conjugate. The kinetic energy T is obtained by
pl (t)2 1 1 ikla
T = = −iωk m √ e Q k (t) + c.c.
2m 2m l k
N
1 ik la
× −iωk m √ e Q k (t) + c.c. . (4.12)
k
N
54 4 Lattice Vibration and Phonon
1 ikla+ik la
e = δk,−k , (4.13)
N l
m 2
= ω − Q k (t)Q −k (t) − Q ∗k (t)Q ∗−k (t)
2 k k
∗ ∗
+Q k (t)Q k (t) + Q k (t)Q k (t) . (4.14)
by using different pairs of displacements. Each term of the right-hand side in the
equation is expressed as V1 , V2 , and V3 . V1 is obtained by
1 1 −ikla ∗
V1 = f √ e Q k (t) + √ e
ikla
Q k (t)
l k
N N
1 1 −ik la ∗
ik la
× √ e Q k (t) + √ e Q k (t)
k
N N
= f Q k (t)Q −k (t) + Q k (t)Q ∗k (t)
k
+Q ∗k (t)Q k (t) + Q ∗k (t)Q ∗−k (t) . (4.16)
f −ika f mωk2
f − e − eika = , (4.18)
2 2 2
we get
mωk2
V = Q k (t)Q −k (t) + Q k (t)Q ∗k (t) + Q ∗k (t)Q k (t) + Q ∗k (t)Q ∗−k (t) .
2
(4.19)
The lattice vibration is expressed as the summation of the harmonic oscillators with
their coordinates Q k (t), which are not the same as the atomic displacements q l (t).2
For the canonical quantization, the coordinate ql and momentum pl are replaced by
linear operators q̂l and p̂l , which satisfy the commutation relations
√
2 At the Γ -point (k = 0), the atomic displacement in the lth unit cell is the same of Q k (t)/ N .
56 4 Lattice Vibration and Phonon
l n
q̂ , p̂ = iδl,n ,
l n
q̂ , q̂ = p̂l , p̂ n = 0. (4.23)
1 1 −ikla †
q̂ =
l
√ e Q̂ k + √ e
ikla
Q̂ k
k
N N
1
= √ eikla Q̂ k + Q̂ †−k , (4.24)
k
N
because the summations of k and −k are the same for the periodic boundary
conditions. The Fourier transformation of q̂ l is expressed by
1 −ikla l
N
√ e q̂ = Q̂ k + Q̂ †−k . (4.25)
N l=1
1 ikla 1 −ikla †
p̂ =
l
−iωk m √ e Q̂ k + iωk m √ e Q̂ k ,
k
N N
m ikla
= √ e −iωk Q̂ k + iω−k Q̂ †−k . (4.27)
N k
1 −ikla l
N
√ e p̂ = −iωk m Q̂ k + iω−k m Q̂ †−k
N l=1
= β̂k mωk , (4.28)
1
Q̂ k = α̂k + i β̂k ,
2
1
Q̂ †−k = α̂k − i β̂k . (4.30)
2
[ Q̂ k , Q̂ †k ] = Q̂ k Q̂ †k − Q̂ †k Q̂ k
1
= ˆ
(α̂k + i βk )(α̂−k − i β̂−k ) − (α̂−k − i β̂−k )(α̂k + i βk )
ˆ
4
1
= [α̂k , α̂−k ] + [β̂k , β̂−k ] + i[β̂k , α̂−k ] − i[α̂k , β̂−k ] . (4.31)
4
The operators α̂k and α̂k† and β̂k and β̂k† are commutative, because α̂k and β̂k include
only q̂ l or p̂l , respectively: α̂k , α̂−k = 0, β̂k , β̂−k = 0. Then we get
1
[ Q̂ k , Q̂ †k ] = i[β̂−k , α̂k ] + i[β̂k , α̂−k ] . (4.32)
4
Using the commutation relation [ p̂l , q̂ l ] = −iδl,l we get
N
1 1
ia(k −k)
[ Q̂ k , Q̂ †k ] = + e = δk,k . (4.34)
4N m l=1
ωk ωk 2mωk
† 1
Ĥ = ωk b̂k b̂k + . (4.37)
k
2
This Hamiltonian has the same form as the harmonic oscillator, and the operators
b̂k† and b̂k are creation and annihilation operators, respectively. This quantum of the
lattice vibration is called a phonon.3 The phonon is a boson like the photon. In the
stationary condition, the lattice wave with the wave number k is explained to be
occupied by n k phonons.
In the linear atomic chain, there is only one phonon mode (acoustic phonon). Next,
we consider a linear diatomic chain, which includes two atoms (atom 1 and atom 2)
in a unit cell.4 The atom 1 and atom 2 have masses of m 1 and m 2 , respectively. a is
a length of the unit cell. In the linear diatomic chain, another phonon mode (optical
phonon) exists in addition to the acoustic phonon. The atomic displacement of each
atom in the lth unit cell is expressed by qκl for atom κ, κ = 1, 2. By adapting the
periodic boundary condition for N unit cells, qκl+N = qκl (Fig. 4.3).
The kinetic energy T and the potential energy V are obtained by
( p l )2
κ
T =
l,κ
2m κ
1 2 2
V = f q2l − q1l + g q1l+1 − q2l , (4.38)
2 l
d l
pκl = m κ q̇κl = m κ q . (4.39)
dt κ
3A phonon is a collective oscillation, and is sometimes referred to as a quasi particle. The quasi
particle is a fictitious body consisting of the original real individual particle pulse train of disturbed
neighbors [2].
4 We referred to the method described in [3] for classical and quantum mechanics of the linear
diatomic chain.
4.2 Linear Diatomic Chain 59
Fig. 4.3 Schematic of the linear diatomic chain. The unit cell contains two atoms (atom 1 and
atom 2 represented by black and white balls) with masses of m 1 and m 2 , respectively. The force
constants between two atom 1 and atom 2 are f and g. x(l) represents the position of the lth unit
cell. x1 and x2 represent the equilibrium position of the atom 1 and atom 2 from the edge of the
unit cell, respectively
For the atom 1 in the lth unit cell, the equation of motion is
d2 l dV
m1 2
q1 = − l
dt dq1
1 d 2 2
=− l
f q2l − q1l + g q1l+1 − q2l
2 dq
l 1 l
= f q2 − q1 + g q2l−1 − q1l , (4.40)
where −dV /dq1l is the force acting on the atom 1. The equation for atom 2 is calcu-
lated in a similar way:
d2 l dV
m2 2
q1 = − l = f q1l − q2l + g q1l+1 − q2l . (4.41)
dt dq2
1 1
A0 (k)aκ (k)eikxκ −iω(k)t + c.c.,
l
qκl = √ (4.42)
2 N mκ
where k is the wave number, ω(k) is the vibration frequency, and xκl is the equilibrium
position of the atom κ in the lth unit cell. Inserting (4.42) into (4.40), we get
√ l 1 l 1 l
− (ω(k))2 m 1 a1 (k)eikx1 = f √ a2 (k)eikx2 − √ a1 (k)eikx1
m2 m1
1 l 1 l
+ g √ a2 (k)eikx2 − √ a1 (k)eikx1 , (4.43)
m2 m1
where we neglected the complex conjugate term. The equilibrium position of each
atom has relationships as follows: x1l = x(l) + x1, x2l = x(l) + x2, and x2l−1 =
60 4 Lattice Vibration and Phonon
f +g f + ge−ika ik(x2−x1)
(ω(k))2 a1 (k) = a1 (k) − √ e a2 (k). (4.45)
m1 m1m2
f +g f + geika −ik(x2−x1)
(ω(k))2 a2 (k) = a2 (k) − √ e a1 (k). (4.46)
m2 m1m2
where
f +g
D11 =
m1
f + ge−ika ik(x2−x1)
D12 = √ e
m1m2
f +g
D22 =
m2
f + geika −ik(x2−x1)
D21 = √ e . (4.48)
m1m2
then
Here we calculate
f +g f +g m1 + m2
D11 + D22 = + = ( f + g)
m1 m2 m1m2
f +g
= ,
μ
2 fg 4 fg ka
D11 D22 − D12 D21 = (1 − cos(ka)) = sin2 , (4.51)
m1m2 m1m2 2
Here, we consider a special case for the linear diatomic chain: f = g for the force
constant, m 2 > m 1 for the atomic mass and x1 = 0, x2 = a/2 for the equilibrium
atomic position. From (4.54), the eigenfrequency is
3f
8 2 ka
ω j (k) = 1 ± 1 − sin . (4.55)
2m 1 9 2
62 4 Lattice Vibration and Phonon
0.0
-1.0 -0.5 0.0 0.5 1.0
Wave number (π /a)
The frequency is plotted as a function of the wave number in Fig. 4.4. There is
a gap, in which the frequency does not exist, between the high- and low-frequency
branches.
At ka 0 condition (long-wavelength
√ approximation), the eigenfrequencies were
obtained to be ω1 = 0 and ω2 = 3 f /m 1 , where we set the low and high frequencies
ω1 and ω2 , respectively. The atomic displacements are calculated by using (4.42) as
q1l
=1 (4.56)
q2l
q1l m2
=− = −2 (4.57)
q2l m1
for the j = 2 branch. All atoms move in the same direction with the same displace-
ment, and the whole chain moves in the j = 1 branch, which is called an acoustic
mode. In the j = 2 branch, two neighboring atoms move in opposite directions to
each other and the center of mass does not move. This oscillation is called an optical
mode, because it can induce electronic polarization and interacts with the electronic
√ light for ions. At√around the ka = π condition, the eigen frequencies are
field of
ω1 = f /m 1 and ω2 = 2 f /m 1 for the acoustic and optical modes, respectively.
The optical phonons with a wave number of k ∼ 0 are quite important because they
are excited by light, for example, in Raman spectroscopy (Fig. 4.5).
Fig. 4.5 Directions of atomic displacement for the acoustic mode (a) and the optical mode (b) at
ka 0
⎛ ⎞
1
qκl = √ ⎝ l
aκ, j (k)eikxκ Q j (k) + aκ, j (−k)e−ikxκ Q j (−k)⎠ , (4.58)
l
N m κ j,k≥0 j,k>0
where
1
Q j (k) = A0, j (k)e−iω j (k)t + A0, j (k)eiω j (−k)t
2
1
= A0, j (k)e−iω j (k)t + A∗0, j (−k)eiω j (k)t . (4.59)
2
The most important merit to use normal coordinates is that the Hamiltonian is diag-
onalized:
1
H =T +V = P j (k)P j∗ (k) + Q j (k)Q ∗j (k)ω2j . (4.60)
2 k, j
then we get
Using the same method as the linear atomic chain, we replace Q j (k) and P j (k) by
Q̂ j (k) and P̂ j (k), which satisfy the commutation relation:
Q̂ ∗j (k), P̂ j (k ) = Q̂ j (k), P̂ j∗ (k ) = iΔ(k − k )δ j, j
Q̂ ∗j (k), Q̂ j (k ) = P̂ j∗ (k), P̂ j (k ) = 0, (4.63)
1 ∗
H = P̂ j (k) P̂ j (k) + ω2j Q̂ ∗j (k) Q̂ j (k)
2 k, j
1 1
= ω j b̂†j (k)b̂ j (k) + . (4.65)
2 k, j 2
4.3 Summary
References
1. Haken, H.: Quantum Filed Theory of Solids. North-Holland Publishing Company, Amsterdam
(1976)
2. Bruesch, P.: Phonons: Theory and Experiments I. Springer Series in Solid-State Sciences 34.
Springer, NewYork (1982)
3. Mattuck, R.D.: A Guide to Feynman Diagrams in the Many-Body Problems, 2nd edn. Dover
Publications Inc., New York (1992)
Chapter 5
Coherent Phonons: Experiment
Coherent phonons are coherently excited by an ultrashort optical pulse and used to
study dynamics of optical phonons. In this chapter, we present a brief history of
experiments and phenomenological theories of coherent phonons. A typical exam-
ple of the experimental technique and results of coherent phonons using a pump
and probe protocol is explained. Several selected examples of coherent phonons in
semiconductors and semimetals are presented. In addition, the squeezed phonons,
which are essential nonclassical states, are described.
When an ultrashort optical pulse with duration much shorter than an oscillation
period of optical phonons is irradiated on materials, the optical phonons are excited
impulsively and coherently. Oscillation of the induced optical phonons is detected via
a transient change in reflectivity or transmissivity. Such impulsively excited phonons
are called coherent phonons.1 Experimental studies of coherent optical phonons
started in the mid-1980s. Keith Nelson and his coworkers demonstrated femtosecond
time-resolved measurements of optical phonons in the organic molecular crystal
(α-perylene) [1]. They used 70 fs optical pulses with wavelength of 620 nm and
performed pump–probe experiments. The α-perylene crystal has two pairs of four
planar molecules in a unit cell. They observed coherent oscillations due to vibrational
(33 and 80 cm−1 ) and translational (104 cm−1 ) modes.
1 It should be mentioned that the coherent phonons are not necessarily a coherent state of the phonon.
Generation mechanisms of the optical coherent phonons are often classified into
three mechanisms: the impulsive stimulated Raman scattering (ISRS) [31, 32] for
transparent conditions, the displacive excitation of coherent phonons (DECP) [33]
for opaque conditions, and the screening of the surface-charge field for a polar semi-
conductor such as gallium arsenide [2, 34]. The equation of motion for an optical
phonon is expressed in a normal coordinate Q using a classical harmonic oscillator
model:
d2 Q dQ
2
+ 2γ + ω02 Q = F(t), (5.1)
dt dt
where γ and ω0 are the vibrational damping constant and frequency, respectively,
F(t) is the driving force.
The ISRS mechanism was first proposed by Yan et al. [31]. When an ultrashort
optical pulse passes through a Raman-active material, phonons, which have vibra-
tional periods longer than the pulse width, are excited by the ISRS process, because
the Stokes frequency is contained within the bandwidth of the pulse. The driving
force is given by
1 ∂α
F(t) = N E 2, (5.2)
2 ∂Q
shift is due to photoexcited electrons (n(t): electron density per unit volume) in the
excited state. They assumed the linear dependence of the shift on the electron density:
Q 0 (t) = κn(t). The driving force is expressed as
where E p and g(t) are the energy and a normalized pulse-shape function of the
optical pulse, respectively, ρ is a constant of proportionality for carrier generation,
and β is a rate constant for relaxation. The oscillation shows a cos(ω0 t) dependence
in DECP mechanism. The DECP mechanism is analogous to the mechanism using
a displaced potential for the electronic excited state, which is proposed for coherent
oscillations in large organic molecules.
For polar semiconductors, another generation mechanism due to a change in a
surface-charge field is proposed [2, 34]. The driving force of this mechanism is the
sudden depolarization of the crystal lattice due to the ultrafast change of intrinsic
surface-charge field in its depletion layer which is given by
t
F(t) = J j (t )dt , (5.4)
∞
where J j (t) is a current associated with drift of photoexcited carriers in the surface-
charge field.
Kuznetsov and Stanton [35] proposed a microscopic theory in which the dynam-
ics of electrons and phonons are described by kinetic equations using quantum-
mechanical operators. They used the Hamiltonian:
†
Ĥel = εαk ĉαk ĉαk + ωq b̂q† b̂q
k,α q
†
+ Mkq b̂q + b̂q† ĉαk ĉαk+q , (5.5)
α,k,q
where ĉ† , ĉ are the electron creation and annihilation operators in k space, respec-
tively, εαk is the energy dispersion in band α = {c, v} (conduction or valence band).
ωq , b̂† , and b̂ are the phonon angular frequency, the phonon creation, and annihilation
operators. The last term is an interaction term describing electron–phonon coupling,
and Mkq is the coupling constant related to a deformation potential. They defined
a coherent amplitude Dq ≡ b̂ + b̂† of the qth phonon mode using the statistical
average b̂ and b̂† and obtained the equation of motion using Dq :
∂2
α α
Dq + ω 2
Dq = −2ωq Mkq n k,k+q , (5.6)
∂t 2 q
α,k
70 5 Coherent Phonons: Experiment
where n αk,k+q is the electronic density matrix. The derived equation is similar to the
phenomenological equation for classical oscillators.
Numerical simulations for generating coherent phonons have been studied using
the time-dependent density functional theory (TDDFT) by Yabana and his coworkers
[36, 37]. They used a classical electronic field for the optical pulse, the first principle
theory for electronic states using DFT, and classical dynamics for the atomic motion.
The ISRS process is confirmed to be a mechanism for generating the driving force
in a diamond sample.
5.2 Experiments
The coherent phonons are usually excited by irradiation of an ultrashort optical pulse
and detected by using transient optical transmission (or reflection) measurements
using a pump–probe technique. Recently, using ultrashort X-ray pulses, the coherent
phonon oscillation is also detected via a time-resolved X-ray diffraction technique.
The coherent optical phonons are excited and measured via pump–probe transmission
(or reflection) measurements using femtosecond laser pulses. The pump pulse with
duration shorter than a vibration period of a phonon induces the phonons with finite
timing. Induced phonons oscillate in phase and cause a modulation in macroscopic
electric susceptibility, which can be detected via transmissivity (or reflectivity) of
the probe pulse. Figure 5.1 shows a typical example of the pump–probe experiments.
Femtosecond optical pulses with a central wavelength of 800 nm are generated
by a Ti:sapphire oscillator, which is excited by 532 nm light from a continuous wave
laser. The femtosecond pulse is separated into two pulses (pump and probe pulses)
by a beam splitter. The pump and probe pulses pass through different optical paths,
which have their lengths controlled. Since the light travels with the constant light
speed (2.99 × 108 m/s), the arrival time at the sample position is different for the
pump and probe pulses. The difference of 299 nm in the optical paths corresponds
to 1 fs. In this setup, path length for the probe pulse is controlled by using a shaker
(Scan delay unit), which oscillates at 20 Hz. Intensity (signal intensity: Is ) of the
probe pulse transmitted through a sample is detected by a photodiode (PD2), which
does not need a high time resolution. The probe pulse intensity (reference intensity:
Ir ) is monitored by a photodiode (PD1). The differential signal (Is − Ir ) is introduced
to a current amplifier and detected by an oscilloscope. The pump and probe pulses
are focused on the sample by using a lens or a parabolic mirror.2 The output from the
2A parabolic mirror is often used to focus an ultrashort optical pulse shorter than approximately
30 fs in order to avoid strong chirp effects by a lens.
5.2 Experiments 71
Fig. 5.1 Schematic of the experimental setup for coherent phonon measurements using a pump and
probe technique. This is a setup for the transient transmission measurement. BS: beam splitter, CM:
chirp mirror, PD: photodiode, Amp: current amplifier, PC: personal computer. Solid and dashed
lines represent the optical path and the electric connection, respectively
5
Transmittance change (ΔT/T0) 10
laser oscillator is reflected by a pair of chirp mirrors in order to compensate for the
group velocity dispersion of optics and to become the shortest pulse at the sample
position.
Figure 5.2 shows a typical example of the femtosecond time-resolved transmit-
tance measurement of diamond using the pump and probe technique and sub-10 fs
laser pulses at room temperatures.3 There is a strong peak at zero delay due to the non-
3 The used sample is a single crystal of diamond with a (100) face. The sample was fabricated by
chemical vapor deposition and obtained from EPD corporation and has intermediate type between
Ib and IIa with a size of 5 mm square and 0.7 mm thickness [38]. The optical pulse is characterized
immediately behind the output port of the laser oscillator using a spectrometer and a frequency-
resolved autocorrelation (FRAC) technique. The spectrum showed a maximum-intensity wavelength
72 5 Coherent Phonons: Experiment
linear response of diamond by overlapping the pump and probe pulses in the sample.
After the strong peak, there is a modulation caused by the coherent optical phonons
in diamond. Inset in Fig. 5.2 is the enlarged signal in delay between 600 and 800 fs.
The oscillation period is 25.1 ± 0.03 fs (frequency of 39.9 ± 0.05 THz), which is
the same value of the optical phonon of diamond at Γ -point (k = 0) obtained by
Raman spectroscopy.4 The used laser pulse is near infrared, and its energy (1.5 eV)
is well below the band gap (7.3 eV) of diamond. Then, the generation mechanism in
this experiment is ISRS. The oscillation is well reproduced by a damped harmonic
oscillation
ΔT (t)
= C exp(−t/τ ) sin(ωt + θ ), (5.7)
T0
5.2.1.1 Semimetals
Coherent phonons in semimetals were first reported by Cheng et al. [3]. They mea-
sured femtosecond time-resolved reflection from bismuth and antimony samples
using 70 fs pulses of laser light at 1.98 eV (625.8 nm) and the pump and probe tech-
nique. Bismuth and antimony crystallize in the A7 structure, which is a trigonally
distorted cubic structure with two atoms per unit cell. There are two optical phonon
modes: isotropic A1g and anisotropic E g modes. The transient reflection intensity
showed coherent oscillations with frequencies of 2.9 and 4.5 THz, which correspond
to the A1g optical phonon frequencies at Γ -point (k = 0) for bismuth and antimony,
respectively. The anisotropic E g mode phonon oscillation was not observed in their
time-resolved measurement.
By using an electro-optical (EO) sampling technique [34], the E g mode phonon
oscillation was observed in bismuth [39]. In this technique, the polarization of the
probe pulse is set perpendicular to that of the pump pulse. The reflected probe pulse
is directed into a polarizing beamsplitter, and the parallel and perpendicular compo-
nents are detected by using balanced photodetectors. Thus, the difference between
two orthogonal components Rx − R y is detected. A schematic of the EO sampling
is shown in Fig. 5.3. Using this technique, the isotropic oscillation ( A1g ) signal is
suppressed and the weak anisotropic oscillation (E g ) signal can be observed [39].
In addition to the A1g coherent phonons, the E g coherent phonons oscillating with
oscillation period of 475 fs (2.1 THz) were observed. The phonon decay time was
of 792 nm with a bandwidth of approximately 200 nm, and the estimated pulsed width was 8.2 fs at
full width at half maximum from the FRAC measurement.
4 Coherent phonons in diamond were first reported by Ishioka et al. [28] by using sub-10 fs, 395 nm
Fig. 5.3 Schematic of reflective electro-optic (EO) sampling technique. The reflected probe pulse
is directed to a polarizing beam splitter (PBS), and p- and ps-polarized components are separated
and detected by photodiodes (PD1 and PD2). PD1 and PD2 have reverse-bias voltages. An arrow
shows a polarization direction of pulses
obtained to be 3.71 and 2.21 ps for A1g and E g , respectively, by using curve fitting
with two damping oscillations.
5.2.1.2 Semiconductors
-6
Refelectivity change ( ΔR/R) 10
-4
sampling. The coherent
oscillation due to LO (b)
phonons is shown in the 4
intrinsic GaAs. In the 2
n-doped GaAs, there is 0
beating due to two coherent -2
oscillations (LO phonon and -4
LOPC). In the p-doped
GaAs, very small oscillation 4
(c)
of LOPC overlaps with that 2
of LO phonons 0
-2
-4
-1 0 1 2 3 4
Pump probe delay (ps)
Coherent optical phonons have also been detected by using ultrafast time-resolved
X-ray diffraction, where coherent phonons are excited by a femtosecond optical pulse
and detected by using an ultrashort X-ray pulse [42–46]. The first clear experiments
were demonstrated by Sokolowski-Tinten et al. [42] on a bismuth film sample. They
performed femtosecond time-resolved X-ray diffraction using a laser pump and X-ray
probe technique with 120-fs laser pulses and laser-plasma X-rays (LPX).5 The X-ray
diffraction intensity of the (222) reflection showed oscillation just after the pump-
laser pulse irradiation with an oscillation period of 467 fs (frequency of 2.12 THz).
This oscillation is assigned to the A1g optical phonon. The observed frequency was
downshifted from that of the pristine sample (2.92 THz) because the sample was
highly excited by the laser irradiation. The diffraction intensity is dependent on the
square modulus of the structure factor F(h, k, l, t), where h, k, and l are Miller
indices and t is time. The time dependence of the structure factor is expressed by
N
F(h, k, l, t) = f i exp −iG j · r j + δ j , (5.9)
j=1
where f j , r j , and δ j are the atomic scattering factor, the crystallographic atomic
position, and the atomic-deviation vector of the jth atom in a unit cell. For coher-
ently excited phonons, the atomic-deviation vector is δ j = u j sin(ωt + θ ), where
u j and θ are the maximum deviation vector and initial phase. Thus, the diffraction
intensity oscillates according to the optical phonon oscillation. Unlike the optical
reflection or transmission measurements, the value of the atomic displacement is
directly determined using the X-ray diffraction experiment. Sokolowski-Tinten et al.
[42] reported that the maximum atomic displacement is approximately 5–8% of the
nearest-neighbor distance (0.31 nm) between the bismuth atoms in the A7 structure
at pump-laser fluence of 6 mJ/cm2 .
5 When the metal target is irradiated by an intense ultrashort laser pulse, with intensity much larger
than 1016 W/cm2 , the rising edge of the pulse has enough energy to produce plasma. Electrons in the
plasma are accelerated more than keV by the main part of the pulse. The high-energy electrons collide
with target atom, excite inner shell electrons, and produce characteristic X-rays. For example, short-
pulsed K α and K β emissions are reported for a copper target for 42-fs laser-irradiation experiment
with a power density between 3 × 1016 and 2 × 1017 W/cm2 [47].
76 5 Coherent Phonons: Experiment
principle and to reduce fluctuations or quantum noise, squeezed states have attracted
much attention, especially in quantum optics. Squeezed states in condensed matters
including squeezed phonons are also studied and have promising applications such
as gravitation-wave detection and quantum communication.
The squeezed phonons were first reported using transient transmission experiment
with femtosecond laser pulses for a KTaO3 crystal by Garrett et al. [48]. The measured
transmittance coherently oscillated with a frequency close to twice the frequency of
the transverse acoustic (2TA) phonon near the zone boundary. These oscillations are
due to two phonon-squeezed states excited by the second-order Raman scattering.
They estimated that the squeeze factor is approximately 3 × 10−6 of an integrated
pulse intensity of I0 = 19 µJ/cm2 .
Generation of phonon-squeezed states by the second-order Raman scattering is
also theoretically proposed by Hu and Nori [49]. They considered both the sponta-
neous Raman and impulsive Raman processes. In the impulsive Raman process, they
used a δ function for the optical pulse and the Hamiltonian defined by
where Ĥq is the Hamiltonian of the phonon system, and λq is the product of the
second-order polarizability tensor and electric fields of the optical pulses. By solving
the Schrödinger equation with separating the free oscillator terms and the two-phonon
creation and annihilation terms, the time-dependent wave function was obtained as
where ζq = −iλq exp(−iλq /). Then, the second-order Raman process causes the
two-mode quadrature-squeezed operator on the vacuum state.
In the general squeezed states, the mean value and variance of position oscillate
with ω and 2ω, where ω is the phonon frequency, respectively. This feature has also
been reported in transient reflection measurements of Bi and GaAs using femtosec-
ond laser pulses with a pump–probe protocol [50]. They measured the amplitude
(m) of the oscillatory waveform in transient reflectivity. They repeated the measure-
ment 100 times and calculated the variance, σ 2 = (m − m)2 , at each delay. The
transient reflectivity in Bi shows 2.93-THz oscillation, corresponding to the A1g
phonon frequency (ω), in amplitude and 2ω frequency in variance. Similar features
are observed for the longitudinal optical phonons (frequency of 8.54 THz) in GaAs.
They suggested that the observed phase-dependent noise indicates the elliptical shape
for the uncertainty contour and the squeezing of phonons.
There are several reports on the squeezed phonons, and further discussions are
still continuing [51–53].
5.4 Summary 77
5.4 Summary
The coherent phonons are phonons excited by an ultrashort optical pulse and used
to study dynamics of optical phonons. A brief history of the experiments and phe-
nomenological theories of the coherent optical phonons is described. The experi-
mental setup and measurement technique for the coherent phonons in diamond are
presented as examples. Several selected examples of coherent phonons in semicon-
ductors and semimetals are presented. In addition, the squeezed phonons, which are
essential nonclassical states, are described.
References
1. De Silvestri, D., Fujimoto, J.G., Ippen, E.P., Gamble, E.B., Williams, Jr. L.R., Nelson, K.A.:
Femtosecond time-resolved measurements of optic phonon dephasing by impulsive stimulated
Raman scattering in α-perylene crystal from 20 to 300 K. Chem. Phys. Lett. 116, 146–152
(1985)
2. Cho, G.C., Kütt, W., Kurz, H.: Subpicosecond time-resolved coherent-phonon oscillation in
GaAs. Phys. Rev. Lett. 65, 764–766 (1990)
3. Cheng, T.K., Brorson, S.D., Kazeroonian, A.S., Moodera, J.S., Dresselhaus, G., Dresselhaus,
M.S., Ippen, E.P.: Impulsive excitation of coherent phonons observed in reflection in bismuth
and antimony. Appl. Phys. Lett. 57, 1004–1006 (1990)
4. Chwalek, J.M., Uher, C., Whitaker, J.F., Mourou, G.A., Agostinelli, J.A.: Subpicosecond time-
resolved studies of coherent phonon oscillations in thin-film YBa2 Cu3 O6+x (x<0.4). Appl.
Phys. Lett. 58, 980–982 (1991)
5. Misochko, O.V., Hase, M., Ishioka, K., Kitajima, M.: Observation of an amplitude collapse
and revival of chirped coherent phonons in bismuth. Phys. Rev. Lett. 92, 19740 (2004)
6. Katsuki, H., Delagnes, J.D.C., Hosaka, K., Ishioka, K., Chiba, H., Zijlestra, E.S., Garcia, M.E.,
Takahashi, H., Watanabe, K., Kitajima, M., Matsumoto, Y., Nakamura, K.G., Ohmori, K.: All-
optical control and visualization of ultrafast two-dimensional atomic motions in a single crystal
of bismuth. Nat. Commun. 4, 2801 (2013)
7. Cheng, Y.-H., Gao, F.Y., Teitelbaum, S.W., Nelson, K.A.: Coherent control of optical phonons
in bismuth. Phys. Rev. B 96, 134302 (2017)
8. Dekorsy, T., Chi, G.C., Kurz, H.: Coherent phonons in condensed media. Light. Scatt. Solids
VIII 76, 169–209 (2000). and references therein
9. Först, M., Dekorsy, T.: Coherent phonons in bulk and low-dimensional semiconductors. In: De
Silvestri, S., Cerullo, G., Lanzani, G. (eds.) Coherent Vibration Dynamics. CRC Press, New
York (2008). and references therein
10. Yee, K.J., Lee, K.G., Oh, E., Kim, D.S., Lim, Y.S.: Coherent optical phonons oscillations in
bulk GaN excited by far below band gap photons. Phys. Rev. Lett. 88, 105501 (2002)
11. Hase, M., Kitajima, M., Constantinecu, A.M., Petek, H.: The birth of a quasiparticle in silicon
observed in time-frequency space. Nature 426, 51–54 (2003)
12. Hase, M., Katsuragara, M., Constantinecu, A.M., Petek, H.: Frequency comb generation at
terahertz frequencies by coherent phonon excitation in silicon. Nat. Photonics 6, 243–247
(2012)
13. Yoshino, S., Oohata, G., Mizoguchi, K.: Dynamical Fano-like interference between Rabi oscil-
lations and phonons in a semiconductor microcavity system. Phys. Rev. Lett. 115, 157402
(2015)
14. Misochko, O.V., Georgiev, N., Dekorsy, T., Helm, M.: Two crossovers in the pseudogap regime
of YBa2 Cu3 O7−δ supreconductors observed by ultrafast spectroscopy. Phys. Rev. Lett. 89,
067002 (2002)
78 5 Coherent Phonons: Experiment
15. Mansart, B., Boschetto, D., Savoia, A., Rullier-Albenque, F., Forget, A., Colson, D., Rousse,
A., Marsi, M.: Observation of a coherent optical phonon in the iron pnictide superconductor
Ba(Fe1−x Cox )2 As2 (x = 0.06 and 0.08). Phys. Rev. B 80, 172504 (2009)
16. Takahashi, H., Kamihara, Y., Koguchi, H., Atou, T., Hosono, H., Katayama, I., Takeda, J.,
Kitajima, M., Nakamura, K.G.: Coherent optical phonons in the iron oxypnictide
SmFeAsO1−x Fx (x = 0.075). J. Phys. Soc. Jpn. 80, 013707 (2011)
17. Okano, Y., Katsuki, H., Nakagawa, Y., Takahashi, H., Nakamura, K.G., Ohmori, K.: Opti-
cal manipulation of coherent phonons in superconducting YBa2 Cu4 O7−δ thin films. Faraday
Discuss. 153, 375–382 (2011)
18. Lee, I.H., Yee, K.J., Lee, K.G., Oh, E., Kim, D.S.: Coherent optical phonon mode oscillations
in wurtzite ZnO excited by femtosecond pulses. J. Appl. Phys. 93, 4939–4941 (2003)
19. Hu, J., Misochko, O.V., Takahashi, H., Koguchi, H., Eda, T., Nakamura, K.G.: Ultrafast zone-
center coherent lattice dynamics in ferroelectric lithium tantalate. Sci. Technol. Adv. Mater.
12, 034409 (2011)
20. Wall, S., Wegkamp, D., Foglia, L., Appavoo, K., Nag, J., Haglund, R.F., Jr. Stähler, Wolf,
J.: Ultrafast changes in lattice symmetry probed by coherent phonons. Nat. Commun 3, 721
(2012)
21. Wang, Y., Xu, X., Venkatasubramanian, R.: Reduction in coherent phonon lifetime in
Bi2 Te3 /Sb2 Te3 superlattices. Appl. Phys. Lett. 93, 113114 (2008)
22. Kamarajyu, N., Kumar, S., Sood, A.K.: Temperature-dependent chirped coherent phonon
dynamics in Bi2 Te3 using high-intensity femtosecond laser pulses. EPL 92, 47007 (2010)
23. Norimatsu, K., Hu, J., Goto, A., Igarashi, K., Sasagawa, T., Nakamura, K.G.: Coherent optical
phonons in a Bi2 Se3 single crystal measured via transient anisotropic reflectivity. Solid State
Commun. 157, 58–61 (2013)
24. Flock, J., Dekorsy, T., Misochko, O.V.: Coherent lattice dynamics of the topological insulator
Bi2 Te3 probed by ultrafast spectroscopy. Appl. Phys. Lett. 105, 011902 (2014)
25. Norimatsu, K., Hada, M., Yamamoto, S., Sasagawa, T., Kitajima, M., Kayanuma, Y., Nakamura,
K.G.: Dynamics of all Raman active coherent phonons in Sb2 Te3 . J. Appl. Phys. 117, 143102
(2015)
26. Hu, J., Igarashi, K., Sasagawa, T., Nakamura, K.G., Misochko, O.V.: Femtosecond study of
A1g phonons in the strong 3D topological insulators: From pump-probe to coherent control.
Appl. Phys. Lett. 112, 031901 (2018)
27. Mishina, T., Nitta, K., Matsumoto, Y.: Coherent lattice vibration of interlayer shearing mode
of graphite. Phys. Rev. B 62, 2908–2911 (2000)
28. Ishioka, K., Hase, M., Kitajima, M., Petek, H.: Coherent optical phonons in diamond. Appl.
Phys. Lett. 89, 231916 (2006)
29. Gambetta, A., Manzoni, C., Menna, E., Meneghetti, M., Cerullo, G., Lanzani, G., Tretiak,
S., Piryatinski, A., Saxena, A., Matin, R.L., Bishop, A.R.: Real-time observation of nonlinear
coherent phonon dynamics in singe-walled carbon nanotubes. Nat. Phys. 2, 515–520 (2006)
30. Kim, J.-H., Nugraha, A.R.T., Booshehri, L.G., Hároz, E.H., Satom, K., Sanders, G.D., Yee,
K.-J., Lim, Y.S., Stanton, C.J., Saito, R.: Coherent phonons in carbon nanotubes and graphene.
Chem. Phys. 413, 55–80 (2013)
31. Yan, Y.-X., Gamble Jr. E.B., Nelson, K.A.: Impulsive stimulated scattering: General importance
in femtosecond laser pulse interactions with matter, and spectroscopic applications. J. Chem.
Phys. 83, 5391 (1985)
32. Merlin, R.: Generating coherent THz phonons with light pulses. Solid State Commun. 102,
107–220 (1997)
33. Zeiger, H.J., Vidal, J., Cheng, T.K., Ippen, E.P., Dresselhaus, G., Dresselhaus, M.S.: Theory
for displacive excitation of coherent phonons. Phys. Rev. B 45, 768–778 (1992)
34. Pfeifer, T., Dekorsy, T., Kütt, W., Kurz, H.: Generation mechanism for coherent LO phonons
in surface-space-charge fields of III–V-compounds. Appl. Phys. A 55, 482–488 (1992)
35. Kuznetsov, A.V., Stanton, C.J.: Theory of coherent phonon oscillations in semiconductors.
Phys. Rev. Lett. 43, 3243–3246 (1994)
References 79
36. Shinihara, Y., Kawashita, Y., Iwata, J.-I., Yabana, K., Otobe, T., Bertsch, G.F.: First-principles
description for coherent phonon generation in diamond. J. Phys.: Condens. Matter 22, 38412
(2010)
37. Shinihara, Y., Yabana, K., Kawashita, Y., Iwata, J.-I., Otobe, T., Bertsch, G.F.: Coherent phonon
generation in time-dependent density functional theory. Phys. Rev. B 82, 155110 (2010)
38. Sasaki, H., Tanaka, R., Okano, Y., Minami, F., Kayanuma, Y., Shikano, Y., Nakamura, K.G.:
Coherent control theory and experiment of optical phonons in diamond. Sci. Rep. 8, 96909
(2018). https://doi.org/10.1038/s41598-018-27734-1.
39. Hase, M., Mizoguchi, K., Harima, H., Nakashima, S., Tani, M., Sakai, K., Hangyo, M.: Optical
control of optical phonons in bismuth filma. Appl. Phys. Lett. 69, 2474–2476 (1996)
40. Cho, G.C., Dekorsy, T., Bakker, H.J., Hövel, R., Kurz, H.: Generation and relaxation of coherent
majority phonons. Phys. Rev. Lett. 77, 4062–4065 (1996)
41. Hu, J., Zhang, H., Sun, Y., Misochko, O.V., Nakamura, K.G.: Temperature effect on the coupling
between coherent longitudinal phonons and plasmons in n-type and p-type GaAs. Phys. Rev.
B 97, 165307 (2018)
42. Sokolowski-Tinten, K., Blome, C., Blums, J., Cavalleri, A., Dietrich, C., Tarasevitch, A.,
Uschmann, I., Förster, E., Kammler, M., Horn-von-Hoegen, M., von der Linde, D.: Femo-
tosecond X-ray measurement of coherent lattice vibrations near the lindemann stability limit.
Nature 422, 287 (2003)
43. Nakamura, K.G., Ishii, S., Ishitsu, S., Shiokawa, M., Takahashi, H., Dharmalingam, K., Irisawa,
J., Hironaka, Y., Ishioka, K., Kitajima, M.: Femtosecond time-resolved X-ray diffraction from
optical coherent phonons in CdTe(111) crystal. Appl. Phys. Lett. 93, 061905 (2008)
44. Johnson, S.L., Beaud, P., Vorobeva, E., Milne, C.J., Murray, É.D., Fahy, S., Ingland, G.: Directly
observing squeezed phonon states with femtosecond x-ray diffraction. Phy. Rev. Lett. 102,
175503 (2009)
45. Harmand, M., Coffee, R., Bionta, M.R., Chollet, M., French, D., Zhu, D., Fritz, D.M., Lemke,
H.T., Medvedev, N., Ziaja, B., Toleikis, S., Cammarata, M.: Achieving few-femtosecond time-
sorting at har X-ray free-electron lasers. Nat. Photonics 7, 215–218 (2013)
46. Johnson, S.L., Beaud, P., Möhr-Vorobeva, E., Caviezel, A., Ingold, G., Milne, C.J.: Direct
observation of non-fully-symmetric coherent optical phonons by femtosecond x-ray diffraction.
Phys. Rev. B 87, 054301 (2013)
47. Yoshida, M., Fujimoto, Y., Hironaka, Y., Nakamura, K.G., Kondo, K., Ohtani, M., Tsunemi,
H.: Generation of picosecond hard x rays by tera watt laser focusing on a copper target. Appl.
Phys. Lett. 73, 2393–2395 (1998)
48. Garrett, G.A., Rojo, A.R., Sood, A.K., Whitaker, J.F., Merlin, R.: Vacuum squeezing of solids:
macroscopic quantum states driven by light pulses. Science 275, 1638–1640 (1997)
49. Hu, X., Nori, F.: Phonon squeezed states generated by second-order Raman scattering. Phys.
Rev. Lett. 79, 4605–4608 (1997)
50. Misochko, O.V., Sakai, K., Nakashima, S.: Phase-dependent noise in femtosecond pump-probe
experiments on Bi and GaAs. Phys. Rev. B 61, 11225–11228 (2000)
51. Misochko, O.V., Hu, J., Nakamura, K.G.: Controlling phonon squeezing correlation via one-
and two- phonon interference. Phys. Lett. A 375, 4141–4146 (2011)
52. Esposito, M., Titimbo, K., Zimmermann, K., Giusti, F., Randi, F., Boschetoo, D., Floreanini,
R., Benatti, F., Fausti, D.: Photon number statistics uncover the fluctuations in non-equilibrium
lattice dynamics. Nat. Commu. 6, 10249 (2015)
53. Benatti, F., Esposito, M., Fausi, D., Floreanini, R., Titimbo, K., Zimmermann, K.: Generation
and detection of squeezed phonons in lattice dynamics by ultrafast optical excitations. New J.
Phys. 19, 02302 (2017)
Chapter 6
Coherent Phonons: Quantum Theory
In this chapter, we describe a unified quantum mechanical description for the genera-
tion and detection of coherent optical phonons using a simple model. Two electronic
states coupled with displaced harmonic oscillators, which represent phonons, are
used. The dipole interaction between an optical pulse and the system is assumed
and treated as a perturbation interaction. Time evolution of the density operator is
calculated using the second-order perturbation approximation.
There are several quantum theories for generating coherent phonons. The different
theoretical treatments are often used to describe DECP and ISRS processes [1–5].
Here, we attempt to describe the generation of coherent phonons using a unified
quantum mechanical theory [6] in both opaque and transparent conditions.
6.1.1 Hamiltonian
Here, we consider a two-level system for the electronic state and a one-dimensional
harmonic oscillator for the optical phonon at the Γ point with a wave vector k = 0.1
Figure 6.1 shows a schematic of the model potential. |g and |e indicate the electronic
1 Quantum mechanical models with the two-level electronic states are often used to study vibrational
|g>
ground and excited states, respectively. The Hamiltonian Ĥ0 of the materials system
is
Ĥ0 = Ĥg |gg| + ε + Ĥe |ee| (6.1)
where ω is the angular frequency and b̂† and b̂ are the creation and annihilation oper-
ators of the phonon, respectively [6]. Ĥg and Ĥe are Hamiltonians for the harmonic
oscillator in the electronic ground and excited states. The zero-point energy of the
harmonic oscillator is set to be zero. α represents shift between the harmonic oscilla-
tor of the electronic ground and that in the excited states.2 ε is a energy gap between
the electronic ground and excited states. In general, the creation and annihilation
operators are dependent on the wave vector k and expressed as b̂k† and b̂k . Here, we
use only the b̂† and b̂ which represent the operators at k ∼ 0. The reason is that,
according to the phase matching (or momentum conservation), the wave vectors of
phonons which can be excited by the optical processes are only those lying close
to the Γ -point (k ∼ 0), because the optical wavelength is much larger than the lat-
tice constant. The eigenmodes of optical phonons form a continuum around k = 0.
The electromagnetic field of the optical pulse interacts only with atoms within the
penetration depth (δL). In opaque conditions, the penetration depth is usually much
smaller than the sample thickness. This effect relaxes the condition of phase match-
ing and allows the modes with δk ∼ 1/(δL) around k = 0 to be excited. In addition,
the spot size of the pump pulse is smaller than the surface area of the crystal, and the
phase-matching condition along the lateral direction is also relaxed.
The dipole interaction between the materials system and the incident light is
assumed and the interaction Hamiltonian is represented by
d
i |Ψ (t) = Ĥ0 + ĤI (t) |Ψ (t). (6.5)
dt
The solution is given formally by using a time-ordered exponential:
t
1 1
|Ψ (t) = exp Ĥ0 t exp+ Ĥ I (t )dt
i i −∞
×|Ψ (−∞), (6.6)
with
i −i
Ĥ I (t ) = exp Ĥ0 t ĤI (t ) exp Ĥ0 t , (6.7)
where |Ψ (−∞) is the Ket vector of the initial state at t = −∞. Ĥ I (t ) is given by
using (6.1) and (6.3):
i
Ĥ I (t) = exp Ĥg |gg| + (ε + Ĥe |ee|)t
× μE 0 f (t)(e−iΩt |eg| + eiΩt |ge|)
−i
× exp ( Ĥg |gg| + (ε + Ĥe )|ee|)t
i −i
= μE 0 f (t) exp (ε − Ω + Ĥe )t |eg| exp Ĥg t + H.c. , (6.8)
3 When we set B̂ = |ee|, the exponential exp( Ĥe B̂t) is expressed by exp( Ĥe B̂t) =
n ( Ât) Since B̂ 2 = (|ee|)(|ee|) = |ee| = B̂, then B̂ n = B̂ for n ≥ 1. Finally, we get
n B̂ n .
includes not only the ISRS mechanism, which is previously proposed (Chap. 5)
for a transparent condition, but also the phonon-generation mechanism in an opaque
condition via the photo-induced electronic polarization. The screening of the surface-
space-charge field is also included in either the IA or ISRS processes depending on
the excitation condition, because both the photoexcited electron–hole pair and the
polarization can change the surface-space-charge field, which causes deformation of
the potential.4 Even at the resonance condition, the ISRS process is different from
the absorption and emission process, because the population in the electronic state
is not induced by the ISRS process.
where these operators act on the phonon states. Using these operators, Ĥ I (t) is
expressed by
Ĥ I (t) = μE 0 f (t) † (t)|eg| B̂(t) + B̂ † (t)|ge| Â(t) . (6.12)
4 The harmonic potential can be deformed under a long-range external field such as the surface-
charge field. Supposing a uniform electric field (F(x) = −dx) acting along the phonon coordinate
(x), the effective charge field can be treated as uniform in such a short scale. When the external
field potential is applied to a harmonic potential U (x) = kx 2 /2, the potential changes to U (x) =
U (x) + F(x) = k(x − d/k)2 /2 − d2 /(2k). Then, the potential minimum position and energy shift
are d/k and −d2 /(2k), respectively. The slope d of the external fields in the electronic excited state
is lower than that in the ground state, because the surface screening is suppressed by a electron–hole
pair or electronic polarization. Then, the effective harmonic potential in the excited state is displaced
from that in the ground state.
86 6 Coherent Phonons: Quantum Theory
t
1
exp+ Ĥ I (t )dt
i −∞
2 t t
1 t 1
=1+
Ĥ I (t )dt + Ĥ I (t ) Ĥ I (t )dt dt
i −∞ i −∞ −∞
μE 0 t
=1+ f (t ) † (t )|eg| B̂(t ) + B̂ † (t )|ge| Â(t ) dt
i −∞
t
μE 0 2 t
+ f (t ) † (t )|eg| B̂(t ) + B̂ † (t )|ge| Â(t )
i −∞ −∞
× f (t ) Â (t )|eg| B̂(t ) + B̂ † (t )|ge| Â(t ) dt dt .
†
(6.13)
When the operators ( F̂(t) and Ĝ(t)) act on the |g, 0 state, we can neglect the terms
with |ge| and |ee| and the effective operators are
t
F̂(t) = f (t ) † (t )|eg| B̂(t ) dt
−∞
t t
Ĝ(t) = f (t ) f (t ) B̂ † (t ) † (t )|gg| Â(t ) B̂(t ) dt dt . (6.17)
−∞ −∞
6.1 Generation Mechanism with Displaced Harmonic Oscillator 87
ρ̂(t) = |ψ(t)ψ(t)|
−i μE 0 μE 0 2
= exp Ĥ0 t 1+ F̂(t) + Ĝ(t) |g, 0
i i
μE 0 μE 0 2 † i
× g, 0| 1 + F̂ † (t) + Ĝ (t) exp Ĥ0 t . (6.18)
−i −i
Here, we focus our attention on the terms with μ2 , which are the second-order
interaction with light and denote this as ρ̂ (2) (t):
−i μE 0 μE 0 i
ρ̂ (2) (t) = exp Ĥ0 t F̂ † (t) exp
F̂(t)|g, 0g, 0| Ĥ0 t
i −i
2
−i μE 0 i
+ exp Ĥ0 t Ĝ(t)|g, 0g, 0| exp Ĥ0 t
i
−i μE 0 2 † i
+ exp Ĥ0 t |g, 0g, 0| Ĝ (t) exp Ĥ0 t , (6.19)
−i
where we denote the terms in the right-hand side as ρ̂1(2) (t), ρ̂2(2) (t), and ρ̂3(2) (t). ρ̂1(2) (t)
corresponds to the impulsive absorption process ρ̂2(2) (t), and ρ̂3(2) (t) corresponds to
the impulsive stimulated Raman scattering process.
where we used
1 2
−i
exp Ĥg t |0 = 1 + ωb̂† b̂ + ωb̂† b̂ + · · · |0 = |0, (6.22)
2
because b̂† b̂|0 = 0. A similar calculation gives the latter half in the right-hand side
in (6.20) as
t
i i
e, 0| dt f (t ) exp ε − Ω + Ĥe (t − t ) exp Ωt . (6.23)
−∞
which shows the system is in the electronic excited state at time t. The expectation
value for the electronic excitation state is
μE 0 2 t −i
e|ρ̂1(2) (t)|e = dt f (t ) exp ε − Ω + Ĥe (t − t ) |0
−∞
t
i
0| dt f (t ) exp ε − Ω + Ĥe (t − t )
−∞
2
μE 0
= |ψ p (t)ψ p (t)|, (6.25)
The final electronic state is the excited state and this process corresponds to the
absorption process.
6.1 Generation Mechanism with Displaced Harmonic Oscillator 89
Here
i
i
B̂ (t ) Â (t )|g = exp
† †
Ĥg t exp ε − Ω + Ĥe t |g, (6.28)
and
−i −i
g| Â(t ) B̂(t ) = g| exp ε − Ω + Ĥe t exp Ĥg t , (6.29)
then, we get
i
B̂ † (t ) † (t )|gg| Â(t ) B̂(t ) = exp Ĥg t
−i −i
× exp ε − Ω + Ĥe (t − t ) exp Ĥg t |gg|. (6.30)
In addition,
i i
g, 0| exp Ĥ0 t = |g, 0 exp Ĥg |gg|t = g, 0|, (6.32)
then, we get
t t
μE 0 2 −i i
ρ̂2(2) (t) = exp Ĥ0 t
f (t ) f (t ) exp Ĥg t
i −∞ −∞
−i
exp ε − Ω + Ĥe (t − t ) dt dt |g, 0g, 0|
t
μE 0 2 t −i
= f (t ) f (t ) exp Ĥg (t − t )
i −∞ −∞
−i
exp ε − Ω + Ĥe (t − t ) dt dt |g, 0g, 0|. (6.33)
Therefore, the expectation value for the electronic states, which corresponds to the
trace for the electronic states, is
t t
μE 0 2 −i i
g|ρ̂2(2) (t)|g = g| exp Ĥ0 t
f (t ) f (t ) exp Ĥg t
i −∞ −∞
−i
exp ε − Ω + Ĥe (t − t ) dt dt |g, 0g, 0||g
t
μE 0 2 t −i
= f (t ) f (t ) exp Ĥg (t − t )
i −∞ −∞
−i
exp ε − Ω + Ĥe (t − t ) dt dt |00|
= |ψ p (t)0|, (6.34)
which is the Hermitian conjugate of ρ̂2(2) (t). In the paths represented by ρ̂2(2) (t)
and ρ̂3(2) (t), the initial and final states are in the electronic ground state and the light
interaction occurs for the bra or ket state. These paths correspond to the ISRS process.
The phonon ket in the electronic excited state which is expressed by (6.26) as
t
−i −i
|ψ p (t) = f (t ) exp (ε − Ω)(t − t ) exp Ĥe (t − t ) dt |0. (6.38)
−∞
where
D̂(α) is the displacement operator and expressed by D̂(α) ≡ exp
α b̂† − α ∗ b̂ . This relation is shown by
D̂ † (α) Ĥg D̂(α) = ω D̂ † (α)b̂† b̂ D̂(α)
= ω D̂ † (α)b̂† D̂ D̂ † b̂ D̂(α)
= ω b̂† + α b̂ + α
Using the displacement operator, the coherent state |α is expressed by |α =
D̂(α)|0. Using these relations, we get
92 6 Coherent Phonons: Quantum Theory
−i −i
exp Ĥe t = D̂ † (α) exp D̂(α) exp iωα 2 t
Ĥg t
= exp iωα 2 t D̂ † (α) exp −iωb̂† b̂ D̂(α), (6.41)
t
i
= f (t ) exp − (ε − Ω − α 2 ω)(t − t )
−∞
× D̂ † (α)|αe−iω(t−t ) dt , (6.42)
where |αe−iω(t−t ) is the time-dependent coherent state. By setting a new parameter
B ≡ −(ε − Ω − α 2 ω)/(ω), the expectation value of the phonon amplitude in
the electronic excited state is obtained as
†
Q̄ = ψ p (t)| b̂ + b̂ |ψ p (t)
2ω
t t
= f (t ) f (t )e−iωB(t−t ) eiωB(t−t )
2ω −∞ −∞
× αe−iω(t−t ) | D̂(α) b̂† + b̂ D̂ † (α)|αe−iω(t−t ) dt dt
t t
= f (t ) f (t )e−iωB(t −t )
2ω −∞ −∞
× αe−iω(t−t ) | b̂† − α + b̂ − α |αe−iω(t−t ) dt dt
t t
= f (t ) f (t )e−iωB(t −t )
2ω −∞ −∞
× α eiω(t−t ) + e−iω(t−t ) − 2 αe−iω(t−t ) |αe−iω(t−t ) dt dt . (6.43)
∞
αn
|αe−iω(t−t ) = e−|α| /2
√ e−iω(t−t )n |n,
2
(6.44)
n=0 n!
6.1 Generation Mechanism with Displaced Harmonic Oscillator 93
and
∞
αm
αe−iω(t−t ) | = e−|α| /2
m| √ eiω(t−t )m .
2
(6.45)
m=0 m!
n=0
n!
= exp −|α|2 exp |α|2 e−iω(t −t )
= exp |α|2 eiω(t −t ) − 1 . (6.46)
Then, Q̄ is obtained as
t t
Q̄ = f (t ) f (t )e−iωB(t −t ) × α eiω(t−t ) + e−iω(t−t ) − 2
2ω −∞ −∞
× exp |α|2 eiω(t −t ) − 1 dt dt . (6.47)
Finally, we obtain the expectation value of the phonon amplitude by inserting the
form of B as
μE 0 2
Q A (t) = T r [ Q̂ ρ̂2(2) ] = Q̄
t t
μE 0 2
=α f (t ) f (t ) eiω(t−t ) + e−iω(t−t ) − 2
2ω −∞ −∞
× exp −α 1 + iω(t − t ) − eiω(t −t ) ei(ε−Ω)(t −t )/ dt dt .
2
(6.48)
In the ISRS process, the final state is the electronic ground state, which is given by
94 6 Coherent Phonons: Quantum Theory
2
μE 0
g|ρ̂2(2) (t) + ρ̂3(2) (t)|g =− |ψ p (t)0| + |0ψ p (t)| . (6.50)
we get
−i 2
exp Ĥe (t − t ) |0 = eiωα (t −t ) D̂ † (α)|αe−iω(t −t ) .
(6.55)
and then
b̂e−iωb̂ = e−iωt e−iωb̂ b̂t b̂.
† †
b̂t
(6.57)
and
−i −i
0|b̂ exp Ĥg (t − t ) exp Ĥe (t − t ) |0
(t −t ))
= e−iω(t−t ) e−iωα 0|b̂ D̂ † (α)|αe−iω(t −t ) .
2
(6.59)
(6.61)
(6.62)
At the resonance condition of the excitation optical pulse (ε − Ω = 0), the ampli-
tude of the phonons excited via the absorption process is expressed
t t
Q A (t) = α dt dt f (t ) f (t )
2ω −∞ −∞
× e−iω(t−t ) + eiω(t−t ) − 2 . (6.66)
Here, f (t) is a real function and e−iω(t−t ) = cos ω(t − t ) − i sin ω(t − t ) and
eiω(t−t ) = cos ω(t − t ) − i sin ω(t − t ). Because t and t are independent and
have the same integration range, the sine-function terms become zero. Then, we get
t t
Q A (t) = α dt dt f (t ) f (t )
2ω −∞ −∞
× cos ω(t − t ) + cos ω(t − t ) − 2 . (6.67)
The similar calculation gives the expected amplitude of the phonons excited by the
ISRS process which is given by
t t
Q R (t) = α dt dt f (t ) f (t )
2ω −∞ −∞
× e−iω(t−t ) − e−iω(t−t ) + c.c.
t t
= 2α dt dt f (t ) f (t )
2ω −∞ −∞
× cos ω(t − t ) − cos ω(t − t ) . (6.68)
6.1 Generation Mechanism with Displaced Harmonic Oscillator 97
On the other hand, the phonon amplitude decreases along with increasing pulse
width.
ISRS process:
We also consider the t σ condition and set the upper limit of the integration ∞:
∞ ∞
−iωt
Q R (t) = Q 0 e dt dt f (t ) f (t ) eiωt − eiωt + c.c. (6.73)
−∞ −∞
which have the integral range [−∞, ∞] for s and [0, ∞] for u, respectively. t
and t are expressed by t = s + u/2 and t = s − u/2, respectively. Using these
parameters, we get
iωt iωt 1 −(t 2 + t 2 ) iωt iωt
f (t ) f (t ) e −e = exp e − e , (6.76)
πσ2 σ2
and
Here, we use
∞ √
2s 2 π −σ 2 ω2
ds exp − 2 + iωs = √ exp (6.80)
−∞ σ 2σ 8
and
∞ ωu √ σ ω2√2
u2 −σ 2 ω2
du exp − 2 sin = 2σ exp t 2 dt. (6.81)
0 2σ 2 8 0
This integral is defined by the pulse width and the phonon frequency and constant
for a given condition. When we set this constant value as
σ ω2√2
4 −σ 2 ω2
A ≡ Q 0 √ exp t 2 dt (6.82)
π 4 0
The amplitude of phonons induced by ISRS oscillates sine-like, though that of the
impulsive absorption oscillates cosine-like. It is worth noting that the optical phonon
amplitude goes to zero for the short pulse limit (σ ω → 0), because the integral A
goes to zero.
6.1 Generation Mechanism with Displaced Harmonic Oscillator 99
0.4 (a)
0.2
0.0
Mean value of phonon coordinate (Q0)
-0.2
-0.4
0.5 (b)
0.0
-0.5
-1.0
-1.5
-2.0
0.5 (c)
0.0
-0.5
-1.0
-1.5
-2.0
-0.5 0.0 0.5 1.0 1.5 2.0 2.5
Delay (T )
Fig. 6.3 Time evolution of the mean value of the phonon amplitude with pulse width (FWHM) of
0.1T (solid curves), 0.55T (dotted curves), and 0.9T (dashed curves) for the vibrational period T
of the optical phonon. The bottom axis represents the pump–probe delay in unit of T . a represents
Q R excited by ISRS process, b represents Q A excited by IA process. The total amplitude
Q = Q R + Q A is shown in c. This figure is obtained by modifying Fig. 3 of the paper, K.G.
Nakamura et al., Physical Review B 92, 144304 (2015)
Figure 6.3 shows the numerical results of the time evolution of the mean value of
the phonon amplitude and its pulse width dependence. We examined three pulse
width conditions (0.1T , 0.55T , and 0.9T ), where T is the vibrational period of the
optical phonon. For Q A (t), the approximate formula (6.71) agrees well with the
numerical results in the region of time after the passage of the pulse (delay larger than
T ). Violent oscillations of the coherent phonons change to a gradual adaption of a
new equilibrium as the pulse width becomes large. The phonon amplitude decreases
as the pulse width increases. On the other hand, for Q R (t), the phonon amplitude
at the pulse width of 0.55T is larger than that at 0.1T and 0.9T .
In order to discriminate experimentally between Q A (t) and Q R (t), it is nec-
essary to use the electronic-state-selective measurement of the coherent phonons.
100 6 Coherent Phonons: Quantum Theory
0.0
0.0 0.4 0.8 1.2 1.6
Pulse width (T )
This shows that the phonon amplitude decreases as the detuning increases.
For the ISRS process, the calculation is a little bit complicated. The phonon
amplitude is expressed by
6.1 Generation Mechanism with Displaced Harmonic Oscillator 101
∞
1
Q R (t) = √ e−σ ω /8 e−iωt due−u /2σ
2 2 2 2
2π 0
× e−iωu/2 − e−iωu/2 e−iξ u + c.c. (6.85)
The integral term gives a complex number, then an initial phase of the phonon-
amplitude oscillation is dependent on the detuning. Here, we evaluate the large
detuning case (ξ σ 1) compared to the pulse width. In this case, the real part
of the integral term is negligibly small. For the positive value of η, the integral is
calculated by
∞
u2 π σ 2 η2
du exp − 2 − iηu = σ exp −
0 2σ 2 2
σ η/√2
√ σ η2
2
2
− i 2σ exp − et dt. (6.86)
2 0
The real part is smaller than the imaginary part value for the ησ 1 case. Then, we
can get
where
√ −σ 2 ω2 ω ω
B ≡ Q 0 2 exp D ξ+ −D ξ− . (6.88)
8 2 2
d σ
D(x) = −σ 2 x D(x) + √ . (6.90)
dx 2
1
D(x) √ . (6.91)
2σ x
102 6 Coherent Phonons: Quantum Theory
Figure 6.5 shows the numerical results of the time evolution of phonon amplitude
Q A (t), Q R (t), and Q(t) with a pulse width of 0.1T with detuning of ΔE −
0, 3ω, 5ω, and 10ω. The phonon amplitude Q A (t) approaches to zero as the
detuning increases. Q R (t) starts to move in the same direction as Q A (t) at large
detuning.
0.2 (a)
0.1
0.0
Mean value of phonon coordinate (Q0 )
-0.1
-0.2
(b)
0.0
-0.5
-1.0
-1.5
-2.0
(c)
0.0
-0.5
-1.0
-1.5
-2.0
0.0 0.5 1.0 1.5 2.0 2.5
Delay (T )
Fig. 6.5 Time evolution of the phonon amplitude with detuning ΔE of 0ω (solid curves: no detun-
ing), 3ω (dotted curves), and 5ω (dashed curves). a represents Q R (t), b represents Q A (t),
and c represents Q(t). T is the vibrational period of the phonon. The pulse width was set to 0.1T .
This figure is obtained by modifying Fig. 5 of the paper, K.G. Nakamura et al., Physical Review B
92, 144304 (2015)
6.1 Generation Mechanism with Displaced Harmonic Oscillator 103
0.001
0 2 4 6 8 10
Detuning ( )
Figure 6.6 shows the phonon amplitude as a function of the detuning. Both the
phonon amplitudes Q R (t) and Q A (t) decrease as the detuning increases. For
large detuning, the phonon amplitude Q R (t) becomes larger than Q A (t) since no
light absorption occurs, and the dominant generation process of the coherent phonons
is subject to the ISRS process.
In a weak coupling (α < 1) case, the generation of coherent phonons can be cal-
culated using a four-level system consisting of two electronic states (|g and |e)
and zero-phonon and one-phonon states (|0 and |1). There are four levels: |g|0,
|g|1,|e|0, and |e|1. The four-level model is easily used to calculate dynamics
compared to that of the displaced harmonic oscillators (Sect. 6.1), while the both
show the same results in a weak coupling case.
the electron–phonon coupling (α b̂† ). The total interaction causes the multiplicative
coefficient
μE(t ) μE 0
α =α f (t )e−iΩt . (6.93)
i i
At time between t and t , the state |e, 1g, 0| has a time evolution factor of
e−i(ε/+ω)(t −t ) . (6.94)
At time t , g, 0| transfers to e, 0| via the dipole interaction with the incident pulse.
This interaction causes the multiplicative coefficient
μE ∗ (t ) μE 0
= f (t )eiΩt . (6.95)
−i −i
Through these processes, we get the time evolution of the density matrix by
μE 0 −iΩt
ρ1 (t : t , t ) = α e f (t )e−i(ε/+ω)(t −t )
i
μE 0
× f (t )eiΩt e−iω(t−t ) |e, 1e, 0|
−i
μE 0 2
=α f (t ) f (t )e−iΔ(t −t ) e−iω(t−t ) |e, 1e, 0|, (6.97)
6.2 Four-Level Model and Double-Sided Feynman Diagrams 105
Another path is g, 0| → e, 0| at time t and |g, 0 → |e, 1 at time t . Its density
matrix is given by
t
μE 0 2 t
ρ2 (t) = α dt dt f (t ) f (t )eiΔ(t −t ) e−iω(t−t )
−∞ −∞
× |e, 1e, 0|. (6.99)
In addition to these two paths, their Hermitian conjugate paths are possible. The total
density matrix ρ(t) is obtained by ρ(t) = ρ1 (t) + ρ2 (t) + H.c.
At the resonance condition (Δ = 0), the density matrix is expressed as
t
μE 0 2 t
ρ(t) = α dt dt f (t ) f (t )
−∞ −∞
× e−iω(t−t ) + e−iω(t−t ) |e, 1e, 0|
+ eiω(t−t ) + eiω(t−t ) |e, 0e, 1| . (6.100)
In the absorption processes, phonons are excited in the electronic excited state, in
which the potential energy is expressed by the displaced harmonic oscillator, and the
annihilation and creation operators are expressed by b̂ − α and b̂† − α, respectively.
The phonon coordinate operator Q̂ A is expressed by
Q̂ A = b̂ + b̂† − 2α . (6.101)
2ω
Therefore, the |e, 0e, 0| state also contributes to the shifted position, and the density
operator ρ (t) for the |g, 0g, 0| → |e, 0e, 0| transition is
2
μE 0 t
t
ρ (t) = dt dt f (t ) f (t ) e−iΔ(t −t ) + e−iΔ(t −t ) |e, 0e, 0|
−∞ −∞
2
μE 0 t t
= dt dt f (t ) f (t )|e, 0e, 0|. (6.102)
−∞ −∞
106 6 Coherent Phonons: Quantum Theory
Next, we consider the ISRS process for exciting the coherent phonons in the electronic
ground state.
Figure 6.8a and b show the double-sided Feynman diagrams of the impul-
sive stimulated Raman scattering process with the |g, 0 → |e, 1 → |g, 1 and
|g, 0 → |e, 0 → |g, 1 transitions, respectively. The final state (|g, 1g, 0|) is in
the electronic ground state with vibrational polarization. We set the initial state
|g, 0g, 0| at t = −∞. In the path shown by Fig. 6.8a, at time t , |g, 0 changes
to |e, 1 via the dipole interaction with the incident pulse and the electron–phonon
coupling (α b̂† ). The total interaction causes the multiplicative coefficient
μE(t ) μE 0
α =α f (t )e−iΩt , (6.104)
i i
which is the same as that in the IA process (Fig. 6.7). At time between t and t , the
state |e, 1g, 0| has a time evolution factor of
e−i(ε/+ω)(t −t ) . (6.105)
At time t , |e, 1 changes to |g, 1 via the dipole interaction with the incident pulse.
This interaction causes the multiplicative coefficient
μE ∗ (t ) μE 0
= f (t )eiΩt , (6.106)
i i
because the light goes out from the material system. After time t , the state |g, 1g, 0|
evolves with a factor of
e−iω(t−t ) . (6.107)
Through these processes, we get the time evolution of the density matrix as
μE 0 −iΩt
ρ1 (t : t , t ) = α e f (t )e−i(ε/+ω)(t −t )
i
μE 0
× f (t )eiΩt e−iω(t−t ) |g, 1g, 0|
i
μE 0 2
= −α f (t ) f (t )e−iΔ(t −t ) e−iω(t−t ) |g, 1g, 0|.
(6.108)
Another path (Fig. 6.8b) is |g, 0 → |e, 0 at time t and |e, 0 → |g, 1 at time t .
At time t , |g, 0 changes to |e, 0 via the dipole interaction with the incident pulse.
The interaction causes the multiply coefficient
μE(t ) μE 0
= f (t )e−iΩt . (6.110)
i i
At time between t and t , the state |e, 0g, 0| has a time evolution factor of
e−i(ε/)(t −t ) . (6.111)
108 6 Coherent Phonons: Quantum Theory
At time t , |e, 0 changes to |g, 1 via the dipole interaction with the incident pulse and
the electron–phonon coupling. This interaction causes the multiplicative coefficient
μE ∗ (t ) μE 0
−α = −α f (t )eiΩt . (6.112)
i i
It is worth noting that the electron–phonon coupling interaction is −α b̂ for the tran-
sition from the electronic excited to the ground state, because the ground-state har-
monic potential shifts in the −α direction. After time t , the state |g, 1g, 0| evolves
with a factor of
e−iω(t−t ) . (6.113)
Through these processes, we get the time evolution of the density matrix as
μE 0 −iΩt
ρ2 (t : t , t ) = e f (t )e−i(ε/)(t −t ) (−α)
i
μE 0
× f (t )eiΩt e−iω(t−t ) |g, 1g, 0|
i
μE 0 2
=α f (t ) f (t )e−iΔ(t −t ) e−iω(t−t ) |g, 1g, 0|. (6.114)
In addition to these two paths, their Hermitian conjugate paths are possible. The total
density matrix ρ(t) is obtained by ρ(t) = ρ1 (t) + ρ2 (t) + H.c. At the resonance
condition (Δ = 0), the density matrix ρ(t) is obtained to be
2 t
μE 0 t
ρ(t) = α dt dt f (t ) f (t )
−∞ −∞
−iω(t−t ) −iω(t−t )
× e −e |g, 1g, 0| + H.c. (6.116)
In the stimulated Raman scattering process, phonons are excited in the electronic
ground state, and the phonon coordinate operator Q̂ R is expressed by
Q̂ R = b̂ + b̂† . (6.117)
2ω
6.2 Four-Level Model and Double-Sided Feynman Diagrams 109
Q̂ R (t) =T r (ρ(t) Q̂ R )
t
μE 0 2 t
=2α dt dt f (t ) f (t ) cos(ω(t − t )) − cos(ω(t − t )) .
2ω −∞ −∞
(6.118)
The two-level system for the electronic state is extended to a two band model, in
which electronic excited states with a different wave vector k have different energy
levels [11]. The one-dimensional harmonic potentials are also assumed for the optical
phonons. The Hamiltonian is expressed as
Ĥ0 = Ĥg |gg| + Ĥk |kk|
k
Ĥg = ωb̂† b̂
Ĥk = εk + ωb̂† b̂ + αω(b̂ + b̂† ), (6.119)
where |k is the state for which an electron with wave vector k is excited from |g to
the conduction band with the excitation energy of εk . The creation and annihilation
operators of the LO phonon at the Γ -point with energy ω are denoted b̂† and b̂,
respectively. The deformation potential interaction having dimensionless coupling
constant α, which is independent of the k vector, is assumed. Using the rotating wave
approximation, the interaction Hamiltonian between pump pulse and the electronic
state is given by
ĤI (t) = μk E 0 f (t) e−iΩt |kg| + eiΩt |gk| , (6.120)
k
in which μk is the transition dipole moment from |g to |k and f (t) is the envelope
of the pulse. A similar calculation used for the two-level model is used to solve the
Schrödinger equation using Hamiltonians for the two-band model (Fig. 6.9).
band
and |1) under transparent conditions. The pump pulse generates phonon polarization
in the electronic ground state by ISRS and the second-order density operator is given
by
2 t
(2) μE 0 t
ρ (t) = α dt dt f (t ) f (t )
−∞ −∞
−iω(t−t ) −iω(t−t )
× e −e e−iΔ(t −t ) |g, 1g, 0| + H.c., (6.121)
|μ|2 |E 0 |2 −iωt ω
ρ (2) (t) = iα e−σ ω /8 |g, 1g, 0|,
2 2
e √ (6.123)
2
2σ C Δ
2 2
and its Hermite conjugate. The expectation value of the excited coherent optical
phonons Q can be calculated as
(2)
Q(t) = T r [Qρ (t)] = T r (b̂ + b̂† )ρ (2) (t)
2ω
= A|E 0 |2 sin(ωt), (6.124)
2ω
6.4 Optical Detection Mechanism 111
Fig. 6.10 Schematic of the pump–probe heterodyne detection of transient transmission. E pr (t)
is the electronic field of the probe pulse and Ps (t) is the induced polarization. The dotted arrow
indicates dipole radiation from the induced polarization. PD is a photo detector
where, we set
μ2 ω
e−σ ω /8 .
2 2
A=α √ (6.125)
2
2σ C Δ
2 2
The phonon amplitude is proportional to the intensity of the pump pulse. The dis-
placement of atoms defined by the phonon amplitude can not be directly detected
by the optical response, while it can be detected by using the X-ray diffraction. In
the pump and probe protocol, the probe pulse induces electronic polarization Ps (t),
which causes dipole radiation. When a heterodyne detection of the transmitted probe
pulse is investigated, the detection intensity Ih (t) is
where l is the thickness of the sample [10], E 3∗ (t) is the strength of the electronic
field of the probe pulse, and Im(A) is to get a imaginary part of A. Figure 6.10 shows
a schematic drawing of the experimental configuration.
The probe pulse irradiates the sample at time t p , which is a delay between the
pump and probe pulses. There are eight pathways (Fig. 6.11) as given below:
and their Hermite conjugates. The third-order density operator for the path 1 is
obtained by
iμ
ρ1(3) (t ) = i A|E 0 |2 E 3 e−iωt p
t
× dt f 3 (t )e−iωt e−iΩt e−i(ε+ω)(t −t )/ |e, 1g, 0|, (6.128)
−∞
112 6 Coherent Phonons: Quantum Theory
where f 3 (t ) is the Gaussian pulse envelope of the probe pulse and t p is the pump–
probe delay. The polarization (P(t)) induced by the probe pulse is obtained using
the polarization operator P̂ op ≡ μ|ge| + μ∗ |eg| by P(t) = T r (ρ (3) (t) P̂ op ). The
polarization for the path 1 is obtained as
μ2
P1 (t ) = α A|E 0 |2 E 3 e−iωt p
t
× dt f 3 (t )e−iωt e−iΩt e−i(ε+ω)(t −t )/ |e, 1g, 0|, (6.129)
−∞
and the time-integrated intensity (I1 (t p )) of the product between the probe light and
polarization is as
∞
I1 (t p ) = E 3 f 3∗ (t )P1 (t )dt
−∞
μ2
= α A|E 0 |2 |E 3 |2 e−iωt p
∞ t
× dt dt f 3 (t )eiΩt f 3 (t )e−iωt
−∞ −∞
−iΩt −i(ε+ω)(t −t )/
×e e . (6.130)
6.4 Optical Detection Mechanism 113
μ2
I1 (t p ) = α A|E 0 |2 |E 3 |2 e−iωt p
∞ π σ 2C 2
∞
e−2s /σ e−iωs due−u /(2σ ) ei(Δ−ω/2)u
2 2 2 2
×
−∞ 0
√
μ |E 0 | |E 3 | −iωt p π −σ 2 ω2 /8 iα A
2 2 2
≈ e √ e , (6.131)
π σ 2 C 2 2σ Δ − ω/2
1
I1 (t p ) = i B|E 0 |2 |E 3 |2 eiωt p , (6.132)
Δ − ω/2
where
√
π ωα 2 μ2 −ω2 σ 2 /4
B= e . (6.133)
2σ 2 C 4 Δ2 3
1
I2 (t p ) = I3 (t p ) = i B|E 0 |2 |E 3 |2 e−iωt p . (6.134)
Δ + ω/2
6.5 Summary
The generation processes of the coherent optical phonons were described quan-
tum mechanically by using the two electronic levels and the displaced harmonic
oscillators. Two generation processes (IA and ISRS) were investigated. Using the
second-order perturbation, the mean value of the phonon amplitude was calculated
for both the IA and ISRS processes. Detailed calculation was done with the Gaussian
114 6 Coherent Phonons: Quantum Theory
References
1. Scholz, R., Pfeifer, T., Kurz, H.: Density-matrix theory of coherent phonon oscillations in
germanium. Phys. Rev. B 24, 16229–16236 (1993)
2. Kuznetsov, A.V., Stanton, C.J.: Theory of coherent phonon oscillations in semiconductors.
Phys. Rev. Lett. 43, 3243–3246 (1994)
3. Hu, X., Nori, F.: Quantum phonon optics: coherent and squeezed atomic displacement. Phys.
Rev. B 53, 2419–2424 (1996)
4. Merlin, R.: Generating coherent THz phonons with light pulses. Solid State Commun. 102,
107–220 (1997)
5. Stevens, T.E., Kuhl, J., Merlin, R.: Coherent phonon generation and two stimulated Raman
tensors. Phys. Rev. B 65, 144304 (2002)
6. Nakamura, K.G., Shikano, Y., Kayanuma, Y.: Influence of pulse width and detuning on coherent
phonon generation. Phys. Rev. B92, 144304 (2015)
7. Pollard, W.T., Fragnito, H.L., Bigot, J.-Y., Shank, C.V., Mathies, R.A.: Quantum-mechanical
theory for 6 fs dynamic absorption spectroscopy. Chem. Phys. Lett. 168, 239–245 (1990)
8. Banin, U., Bartana, A., Ruthman, S., Kosloff, R.: Impulsive excitation of coherent vibrational
motion ground surface dynamics induced by intense short pulse. J. Chem. Phys. 101, 8461–
8481 (1994)
9. Cerullo, G., Manzoni, C.: Time domain vibrational spectroscopy: principle and experimental
tools. In: De Silvestri, S., Cerullo, G., Lanzani, G. (eds.) Coherent Vibration Dynamics. CRC
Press, Boca Raton (2008)
10. Mukamel, S.: Principles of Nonlinear Optical Spectroscopy. Oxford University Press, New
York (1995)
11. Nakamura, K.G., Ohya, K., Takahashi, H., Tsuruta, T., Sasaki, H., Uozumi, S., Norimatsu, K.,
Kitajima, M., Shikano, Y., Kayanuma, Y.: Spectrally resolved detection in transient-reflectivity
measurements of coherent optical phonons. Phys. Rev. B94, 024303 (2016)
Chapter 7
Coherent Control of Optical Phonons
Coherent control is a technique to control quantum states using laser light and was
originally developed for controlling chemical reactions using coherent two-photon
process to assist chemical reactions via electronic excited states [1–3]. It has been
performed for other physical properties, for example, electronic, vibrational, and
rotational states of atoms and molecules, and electrons, excitons, spins, and phonons
in condensed matter [4–8]. Here, we describe the coherent control of optical phonons
using multiple ultrashort pulses.
The coherent control of optical phonons was first demonstrated on α-perylene molec-
ular crystals at 5 K using multiple femtosecond pulses by Weiner et al. [9]. They
irradiated a femtosecond pulse train consisting of more than 10 pulses, which was
generated through a pulse-shaping technique, on the sample for exciting the opti-
cal phonons via the ISRS process in a transparent condition. They controlled the
oscillation amplitude of the specific phonon mode to be approximately 5 THz owing
to the frequency of the pulse train. A few years later, the coherent control of lon-
gitudinal optical (LO) phonon dynamics was demonstrated in a single crystal of a
© Springer Nature Switzerland AG 2019 115
K. Nakamura, Quantum Phononics, Springer Tracts in Modern Physics 282,
https://doi.org/10.1007/978-3-030-11924-9_7
116 7 Coherent Control of Optical Phonons
-6
fs (a), 242.5 fs (b), 250.9 fs
-20
-30
-400 -200 0 200 400 600 800
Delay between pump and probe pulses (fs)
Amplitude (normalized)
and phase (b) of the
controlled oscillation after
the pump 2 against the 1.5
pump–pump delay τ . The
amplitude is normalized 1.0
using that obtained after
excitation by the pump 1. 0.5
The solid circles are the
experimental data, and the
0.0
solid curves are obtained by
the theoretical calculation.
This figure is obtained by
0.2 (b)
modifying Fig. 4 of the paper,
0.0
Initial phase (π)
-0.4
-0.6
-0.8
230 240 250 260
Delay between pump pulses (fs)
in the pump 2. The third process is important only when the two pump pulses overlap
for the nonresonant condition (transparent condition). All processes are within the
second-order perturbation. The density operator for the second diagram (ρ2(2) ) is the
same as the first one (ρ1(2) ), which is shown in Chap. 6, except for the time delay:
where E 1 and E 2 are the electric field amplitude of the pump 1 and pump 2, τ is the
delay between pump 1 and pump 2, and
μ2 ω
e−σ ω /8 .
2 2
A=α √ (7.2)
2 2C 2 Δ2
The result shows that the phonon amplitude after the second pump pulse irradiation
is a sum of two sinusoidal functions induced by each pulse. The phonon amplitude
and initial phase at the timing of irradiation of pump 2 are calculated for diamond
and shown in Fig. 7.3 as solid curves.
7.3 Coherent Control Theory 119
Fig. 7.4 Diagram of the impulsive stimulated Raman scattering paths. a passes through |e, 1 state
and b passes through the |e, 0 state. Time flows from the felt to the right. The upper part of the
figure shows an envelope of the optical pulse. Interaction between the optical pulse and the system
occurs at time t and t
It is worth noting that the above coherent control is due to constructive or destruc-
tive interference between phonon states excited by the pump 1 and pump 2. This is
because we cannot distinguish which pulse excites the phonons while keeping the
phonon coherence. The phonons are excited by the second-order perturbation, which
interacts with the light twice in one pulse. This is different from the classical control
of pendulum oscillations with two impacts by two pulses.
When the delay between the pump 1 and pump 2 is very short, the third path repre-
sented as (c) in Fig. 7.4 is important. The density operator is calculated as
μ 2 t t
ρ3(2) (t, τ ) = −α e−iΩτ e−iωt dt dt f 1 (t ) f 2 ( f − τ )
−∞ −∞
ε − Ω
× exp −i (t − t ) eiωt − eiωt |g, 1g, 0| + H.c. (7.4)
When the two pulses overlap, there is another transition path, in which the excitation
and relaxation are induced by the pump 2 and pump 1, respectively. The density
operator of this path is
120 7 Coherent Control of Optical Phonons
μ 2 t t
ρ4(2) (t, τ ) = −α e iΩτ −iωt
e dt dt f 1 (t ) f 2 ( f − τ )
−∞ −∞
ε − Ω
× exp −i (t − t ) eiωt − eiωt |g, 1g, 0| + H.c. (7.5)
With ρ3(2) (t, τ ) and ρ4(2) (t, τ ) included, the expectation value of the phonon amplitude
should be modified with the optical frequency Ω. This oscillation can be observed
when the delay between two pump pulses is controlled with a step shorter than the
optical cycle.
In the opaque conditions, the impulsive absorption paths are possible in addition to
the ISRS paths described above. The final state should be |e, 1e, 0| or |e, 0e, 1|.
The paths (a) and (b) are the same as the single pulse excitation described in
Chap. 6. The density operators for the path (c) and (d) in Fig. 7.5 are
Fig. 7.5 Diagrams of the coherent control of optical phonons with the impulsive absorption paths.
Transition occurs by only the pump 1 (a) or the pump 2 (b). (c) and (d) represent the transition
occurring by both the pump 1 and pump 2. Time flows from the felt to the right. The upper part
of the figure shows an envelope of the optical pulse. Interaction between the optical pulse and the
system occurs at time t and t
7.3 Coherent Control Theory 121
μ 2 t t
ρ3(2) (t, τ ) = α e−iΩτ dt dt f 1 (t ) f 2 ( f − τ )
−∞ −∞
ε − Ω
× exp −i (t − t )
× e−iω(t−t ) |e, 1e, 0| + eiω(t−t ) |e, 0e, 1| , (7.6)
and
μ 2 t t
ρ4(2) (t, τ ) = α eiΩτ dt dt f 1 (t ) f 2 ( f − τ )
−∞ −∞
ε − Ω
× exp −i (t − t )
× e−iω(t−t ) |e, 1e, 0| + eiω(t−t ) |e, 0e, 1| . (7.7)
This shows that the phonon amplitude is also modulated by the optical frequency.
There are many applications of the coherent control of optical phonons such as the
mode-selective excitation and phase-transition control. We describe several examples
of the coherent control experiments of the coherent phonons.
with different frequencies of 3 and 2 THz, respectively. Both phonon modes are
excited by an ultrashort optical pulse and detected via the EO sampling technique.
In the A1g and E g modes, the Bi atoms oscillate in the longitudinal and lateral direc-
tions, when the (0001) surface is used. The relative intensity ratio between the A1g
and E g modes, which corresponds to the displacement direction of atoms in the
two-dimensional space, was controlled by controlling the delay between two pump
pulses. The reflectivity change shows beat structures, which are produced by overlap-
ping two oscillations, whose temporal evolution changes drastically as they change
the delay between two pump pulses on the attosecond timescale. They mapped these
beat structures into a two-dimensional space atomic displacement with the density
functional theory calculation of the reflectivity change.
Not only the LO phonons (ω L /2π = 8.7 THz) but also the optical phonon–plasmon
coupled (LOPC) mode oscillations (ω− /2π = 7.7 THz) are coherently controlled
by using a pair of pump pulses at delays between 315 and 345 fs [15]. In n-GaAs,
the oscillation amplitudes of the LO phonons and LOPC mode were harmonically
modulated according to their frequencies by the pump–probe delay. On the other
hand, in p-GaAs, the oscillation amplitude of the LOPC mode was modulated in
phase with that of the LO phonons at the period of LO phonons. They analyzed
the results using a phenomenological model by assuming that the LOPC formation
is delayed from the LO-phonon excitation. The lifetime of the LOPC plays a key
role in understanding different results in two samples. The lifetime of the LOPC
is ∼200 fs in p-GaAs, and ∼800 fs in n-GaAs. Therefore, when the second pump
pulse arrives at the sample, the LOPC excited by the first pump pulse has already
decayed significantly in p-GaAs. Thus, the LOPC excited by the second pump pulse
has nearly no interference with that excited by the first one, and the dynamics are
determined mainly by the interference of LO phonons. The excited LO phonons
cause the phonon–plasmon coupling. The delay time of the LOPC formation was
estimated to be 100 fs for n-GaAs and 120 fs for p-GaAs. In addition to the control
of phonon amplitudes, they reported that the lifetime of the LOPC was coherently
controlled, while that of LO phonons was independent of the pump–pulse delay.
Ge2 Sb2 Te5 (GST) is a chalcogenide phase-change material, in which phase transfor-
mation between the amorphous and crystal phase is optically induced. Makino et al.
[16] demonstrated that the phase change from amorphous into crystalline states is
manipulated by controlling atomic motions through selectively exciting vibrational
modes that involve Ge atom using a pair of femtosecond pulses. The GST superlattice
7.4 Selected Examples 123
sample was irradiated by the two pump pulses with delay of 276 fs, which is close
to the time period of the local A1 mode of the octahedral GeTe6 (in crystalline
phase). The pump pulses coherently enhanced the vibrational amplitude toward the
crystalline phase.
7.5 Summary
References
1. Tannor, D.J., Rice, S.A.: Control of selectivity of chemical reaction via control of wave packet
evolution. J. Chem. Phys. 83, 5013 (1985)
2. Brumer, P., Shapiro, M.: Control of unimolecular reactions using coherent light. Chem. Phys.
Lett. 126, 541 (1986)
3. Tannor, D.J., Kosloff, R., Rice, S.A.: Coherent pulse sequence induce control of selectivity of
reactions: exact quantum mechanical calculations. J. Chem. Phys. 85, 5805 (1986)
4. Ramsay, A.J.: A review of the coherent optical control of the exciton and spin states of semi-
conductor quantum dots. Semicond. Sci. Techol. 25, 103001 (2010). and references therein
5. Katsuki, H., Takei, N., Sommer, C., Ohmori, K.: Ultrafast coherent control of condensed matter
with attosecond precison. Acc. Chem. Res. 51, 1174–1184 (2018)
6. Okano, Y., Katsuki, H., Nakagawa, Y., Takahashi, H., Nakamura, K.G., Ohmori, K.: Opti-
cal manipulation of coherent phonons in superconducting YBa2 Cu4 O7−δ thin films. Faraday
Discuss. 153, 375–382 (2011)
7. Cheng, Y.-H., Gao, F.Y., Teitelbaum, S.W., Nelson, K.A.: Coherent control of optical phonons
in bismuth. Phys. Rev. B 96, 134302 (2017)
8. Hu, J., Igarashi, K., Sasagawa, T., Nakamura, K.G., Misochko, O.V.: Femtosecond study of
A1g phonons in the strong 3D topological insulators: From pump-probe to coherent control.
Appl. Phys. Lett. 112, 031901 (2018)
9. Weiner, A.M., Leaird, D.E., Wiederrecht, G.P., Nelson, K.A.: Femtosecond multiple-pulse
impulsive stimulated Raman scattering spectroscopy. J. Opt. Soc. Am. B 8, 1264 (1991)
10. Dekorsky, T., Kütt, W., Pfeifer, T., Kurz, H.: Coherent Control of LO-Phonon Dynamics in
Opaque Semiconductors by Femtosecond Laser Pulses. Europhys. Lett. 23, 223 (1993)
11. Hase, M., Mizoguchi, K., Harima, H., Nakashima, S., Tani, M., Sakai, K., Hangyo, M.: Optical
control of coherent optical phonons in bismuth films. Appl. Phys. Lett. 69, 2474 (1996)
12. Takahashi, H., Kato, K., Nakano, H., Kitajima, M., Ohmori, K., Nakamura, K.G.: Optical con-
trol and mode selective excitation of coherent phonons in YBa2 Cu3 O7−δ . Solid State Commun.
149, 1955 (2009)
13. Katsuki, H., Delagnes, J.C., Hosaka, K., Ishioka, K., Chiba, H., Zijlstra, E.S., Garcia, M.E.,
Takahashi, H., Watanabe, K., Kitajima, M., Matsumoto, Y., Nakamura, K.G., Ohmori, K.: All-
optical control and visualization of ultrafast two-dimensional atomic motions in a single crystal
of bismuth. Nat. Commun. 4, 2801 (2013)
124 7 Coherent Control of Optical Phonons
14. Hase, M., Fons, P., Mitrofanov, K., Kolobov, A.V., Tominaga, J.: Femtosecond structural trans-
formation of phase-change materials far from equilibrium monitored by coherent phonons.
Nat. Commun. 6, 8367 (2015)
15. Hu, J., Misochko, O.V., Goto, A., Nakamura, K.G.: Delayed formation of coherent LO phonon-
plasmon coupled modes in n- and p-type GaAs measured using a femtoseocnd coherent control
technique. Phys. Rev. B 86, 235145 (2012)
16. Makino, K., Tominaga, J., Hase, M.: Ultrafast optical manipulation of atomic arrangements in
chalcogenide alloy memory materials. Opt. Express 19, 138475 (2011)
Appendix A
Linear Diatomic Chain: Normal Coordinate
and Hamiltonian
Here, we describe the normal coordinates of the linear diatomic chain and diagonalize
its Hamiltonian.1
1
ikxκl −iω j (k)t
l
qκ, j (k) = √ A 0, j (k)a κ, j (k)e + c.c. . (A.1)
2 N mκ
When we set q > 0, the displacement of the atomκ at the lth cell is expressed by
1
A0, j (k)aκ (k)eikxκ −iω j (k)t + c.c.
l
qκl = √
2 N m κ j,k≥0
1
A0, j (−k)aκ (−k)e−ikxκ −iω j (−k)t + c.c. .
l
+ √ (A.2)
2 N m κ j,k>0
where
1
Q kj = A0, j (k)e−iω j (k)t + A0, j (k)eiω j (−k)t
2
1
= A0, j (k)e−iω j (k)t + A∗0, j (−k)eiω j (k)t . (A.4)
2
∗
j (−k) = aκ, j (k), we calculate qκ as
l∗
Using aκ,
⎛ ⎞
1
qκl∗ = √ ⎝ a ∗ (k)e−ikxκ Q k∗
l
j +
∗
aκ, j (−k)e
ikxκ −k∗ ⎠
l
Qj
N m κ j,k≥0 κ, j j,k>0
⎛ ⎞
1
= √ ⎝ a ∗ (k)e−ikxκ Q k∗
l
j +
∗
aκ, j (−k)e
ikxκ −k∗ ⎠
l
Qj . (A.5)
N m κ j,k≥0 κ, j j,k>0
−k
Since the displacement qκl is real, qκl∗ = qκl . Then, we get Q k∗
j = Q j . One of the
most important merits of using the normal coordinates is that the Hamiltonian is
diagonalized by using the normal coordinate (as shown in the next subsection).
pκl 2
T = , (A.6)
l,κ
2m κ
d l
pκl = m κ q = m κ q̇κl
dt κ
mκ
= aκ, j (k) exp(ikxκl ) Q̇ j (k). (A.7)
N k, j
1
= mκ aκ, j (k)aκ, j (k ) exp(i(k + k )la)
2N l,κ k,k , j, j
1
T = aκ, j (k)aκ, j (−k) Q̇ j (k) Q̇ j (−k)
2 κ k, j, j
1 ∗ ∗
= aκ, j (k)aκ, j (k) Q̇ j (k) Q̇ j (k)
2 κ k, j, j
1
= Q̇ j (k) Q̇ ∗j (k), (A.9)
2 k, j
∗
where we use κ aκ, j (k)aκ, j (k) = δ j, j .
Next we calculate the potential energy V using (4.38):
1
V = f S + gS , (A.10)
2
where
2
S= q2l − q1l
l
2
S = q1l+1 − q2l . (A.11)
l
we get
⎛ ⎞
1
q2l
2
= ⎝ √ a2, j (k) exp(ik(la + x2 ))Q j (k)⎠
l l k, j
N m2
⎛ ⎞
1
×⎝ √ a2, j (k ) exp(ik (la + x2 ))Q j (k )⎠
k, j
N m 2
1
= a2, j (k)a2, j (k ) exp(i(k + k )la)
N m 2 l k,k , j, j
× exp(i(k + k )x2 )Q j (k)Q j (k )
1
= a2, j (k)a2, j (−k)Q j (k)Q j (−k)
m 2 k, j, j
1 ∗ ∗
= a2, j (k)a2, j (k)Q j (k)Q j (k). (A.13)
m 2 k, j, j
⎛ ⎞
1
q1l q2l = ⎝ √ a1, j (k) exp(ik(la + x1 ))Q j (k)⎠
l l k, j
N m 1
⎛ ⎞
1
×⎝ √ a2, j (k ) exp(ik (la + x2 ))Q j (k )⎠
k ,j N m 2
1
∗ ∗
= √ a1, j (k)a2, j (k)Q j (k)Q j (k) exp(ik(x 1 − x 2 ))
m 1 m 2 k, j, j
1
∗ ∗
= √ a1, j (k)a2, j (k)Q j (k)Q j (k)ψ12 , (A.15)
m 1 m 2 k, j, j
1
= √ a ∗ (k)a2, j (k)Q j (k)Q ∗j (k)ψ12
∗
. (A.16)
m 1 m 2 k, j, j 1, j
1 f +g f +g
V = Q j (k)Q ∗j (k) ∗
a1, j (k)a1, j (k) +
∗
a2, j (k)a2, j (k)
2 k, j, j m1 m2
1 ∗ ∗ ∗ ∗
− √ a1, j (k)a2, j (k) ( f ψ12 + gΨ12 ) + a1, j (k)a2, j (k) f ψ12 + gΨ12 .
m1m2
(A.19)
1 ∗
−√ ( f ψ12 + gΨ12 ) = D12 (k) = D21 (k)
m1m2
1 ∗ ∗ ∗
−√ f ψ12 + gΨ12 = D12 (k) = D21 (k), (A.20)
m1m2
and we get
130 Appendix A: Linear Diatomic Chain: Normal Coordinate and Hamiltonian
1
V = Q j (k)Q ∗j (k) D11 (k)a1, j (k)a1, ∗ ∗
j (k) + D22 (k)a2, j (k)a2, j (k)
2 k, j, j
∗ ∗
+ D21 (k)a1, j (k)a2, j (k) + D12 (k)a1, j (k)a2, j (k)
1
= Q j (k)Q ∗j (k) D11 (k)a1, j (k) + D12 (k)a2, j (k) a1, ∗
j (k)
2 k, j, j
∗ ∗
+ D22 (k)a2, j (k) + D21 (k)a1, j (k) a 2, j (k) . (A.21)
1
V = Q j (k)Q ∗j (k) ω2j a1, j (k)a1,
∗ ∗
j (k) + ω j a2, j (k)a2, j (k)
2
2 k, j, j
1
= Q j (k)Q ∗j (k)ω2j . (A.22)
2 k, j
1
L =T −V = Q̇ j (k) Q̇ ∗j (k) − Q j (k)Q ∗j (k)ω2j , (A.23)
2 k, j
and the generalized momentum P j (k) conjugate to the normal coordinate Q j (k) is
defined from the Lagrangian:
∂L
P j (k) = = Q̇ j (k). (A.24)
∂ Q̇ j (k)
The Hamiltonian expressed using the normal coordinates and the conjugated mo-
mentum is
1
H =T +V = P j (k)P j∗ (k) + Q j (k)Q ∗j (k)ω2j . (A.25)
2 k, j
Appendix B
Ultrashort Laser Technology
where E 0 is the amplitude of the electric field, σ is the pulse width, and ω0 is the
angular frequency of the light. The spectrum of the pulse is calculated by the Fourier
transformation of the pulse as
∞
1 σ
E(t)e−iωt dt = √ E 0 e−(ω−ω0 ) σ /4 .
2 2
E(ω) = √ (B.2)
2π −∞ 2
The spectrum has a Gaussian form at the center angular frequency of ω0 . The inten-
sities of the pulse and the spectrum are defined by I (t) = E(t)E ∗ (t) and I (ω) =
E(ω)E ∗ (ω), respectively. The widths (full√ width of half maximum) of the pulse
and the spectrum are obtained as Δt = σ 2 ln 2 and Δω = 2π × Δν = 4 ln 2/σ ,
respectively [2]. The relation between the pulse width and the frequency is
2 ln 2
ΔtΔν = ≈ 0.441. (B.3)
π
In more general case, the pulse includes a time-dependent phase term (eiθ(t) and
e−iθ(ω) ) and ΔtΔν is bigger than 0.441. The pulse satisfying (B.3) is called the
Fourier-transform-limited pulse.
When the Fourier-transform-limited pulse passes through optical materials such
as a lens and a beamsplitter, the pulse gets a frequency-dependent phase change [3].
The pulse after passing the optics is expressed as
∞
1
E (t) = √ E(ω)eiθ(ω) eiωt dω. (B.4)
2π ∞
1 d3 θ
+ (ω − ω0 )3 + · · · , (B.5)
6 dω3 ω0
where θ0 is a constant, the second term is a group velocity delay, and the third term
is a group velocity dispersion. If we consider only the third term using
1 d2 θ
δ≡ , (B.6)
2 dω2 ω0
The result indicates that the frequency is dependent on time and the pulse width is
larger than that of the original pulse. This effect is called “chirp”.
where τ is delay between the pump 1 and pump 2. The intensity I (τ ) detected by
the detector is
∞ ∞
I (τ ) = |E 1 (t) + E 2 (t)| dt =
2
(E 1 (t) + E 2 (t))(E 1∗ (t) + E 2∗ (t))dt
−∞ −∞
∞
2t 2 2(t − τ )2
= |E 0 |2 exp − 2 + exp −
−∞ σ σ2
2
t + (t − τ )2
+ 2 exp − cos(ωτ )dt
σ2
√ τ2
= |E 0 |2 2π σ 1 + exp − 2 cos(ωτ ) . (B.10)
2σ
Appendix C
Mathematical Formula
Let us consider the Gaussian integral with two complex parameters (α and β)
© Springer Nature Switzerland AG 2019 135
K. Nakamura, Quantum Phononics, Springer Tracts in Modern Physics 282,
https://doi.org/10.1007/978-3-030-11924-9
136 Appendix C: Mathematical Formula
∞
e−α(x+β) dx,
2
I (α, β) = (C.4)
−∞
where Re(α) > 0.2 The integrand function goes to zero at x → ±∞ for the Re(α)>0
condition, and the integral I (α, β) converges. The partial derivative of I (α, β) by
β is
∞ ∞
∂ I (α, β) ∂ −α(x+β)2
−2α(x + β)e−α(x+β) dx
2
= e dx =
∂β −∞ ∂β −∞
2 ∞
= e−α(x+β) = 0. (C.5)
−∞
Fourier transformation of the Gaussian function is also important and often used in
the text.
∞ ∞
−ax 2 ikx
e−a(x −ikx/a) dx
2
e e dx =
−∞ −∞
∞
e−a(x−ik/(2a)) e−k /(4a) dx
2 2
=
−∞
π −k 2 /(4a)
= I (a, −ik/(2a))e−k /(4a)
2
= e . (C.11)
a
References
1. Bruesch, P.: Phonons: Theory and Experiments I. Springer Series in Solid-State Sciences, vol.
34. Springer, Berlin (1982)
2. Rulliere, C. (ed.): Femtosecond Laser Pulses, Principles and Experiments, 2nd edn. Springer
Science+Business Media Inc, Berlin (2005)
3. Watanabe, S.: Cyotanparusu hassei gijyutu (in Japanese: Generation of ultrashort pulse). Kogaku
24, 378–383 (1995)
4. Trebino, R.: Frequency-Resolved Optical Grating: The measurement of ultrashort laser pulses.
Kluwer Academic Publishers, Norwell (2000)
5. Nomoto: Gauss sekibun no ippannka (in Japanese: General Solution of Gaussian Integral). http://
www.eng.niigata-u.ac.jp/~nomoto/15.html
Index
A Cosine-like oscillation, 73
Acoustic mode, 62 Coupling constant, 69
Acoustic phonon, 58 Creation operator, 29, 64, 105
A1g -mode phonon, 68 Current, 69
Annihilation operator, 29, 34, 64, 105
Atomic deviation, 75
D
Damped harmonic oscillation, 72
B Damping constant, 68
Baker–Hausdrof lemma, 44 Dawson function, 101
Band model, 109 Decay time, 72
Bandwidth, 68 Deformation potential, 109
Bogoliubov transformation, 39, 44 Density functional theory, 122
Boson, 58 Density matrix, 20
Bra vector, 1 Density operator, 11, 85
Dephasing, 74
Depolarization, 69
C Destructive interference, 116, 119
Canonical quantization, 55 Detuning, 100
Chemical reaction, 115 Diagonal component, 18
Chirp mirror, 71 Diamond, 116
Classical number, 3 Dimensionless operator, 57
Closed system, 4 Dimensionless parameter, 55
Coherence, 12 Dipole interaction, 82, 107
Coherent control, 115 Dispersion curve, 52
Coherent phonons, 67 Dispersion relation, 52, 55
Coherent state, 34, 39 Displaced harmonic oscillator, 37, 105
Collective oscillations, 58 Displaced potential, 69
Commulator, 44 Displaced vacuum state, 38
Commutation relation, 25, 26, 30, 55, 57, 64 Displacement, 51
Commutator, 7 Displacement operator, 37, 91
Complete set, 4 Displacive excitation of coherent phonons,
Complex vector space, 1 68
Conduction bad, 109 Double factorial function, 42
Conduction band, 73 Double femtosecond pulses, 116
Constructive interference, 116, 119 Double-sided Feynman diagram, 17, 84,
Cosine-like, 97 103, 117
© Springer Nature Switzerland AG 2019 139
K. Nakamura, Quantum Phononics, Springer Tracts in Modern Physics 282,
https://doi.org/10.1007/978-3-030-11924-9
140 Index
H
E Hamiltonian, 4
Eigenstate, 27 Harmonic oscillator, 25, 35
Eigenvalue, 2, 27 Haung–Rhys factor, 82
Eigenvalue equation, 27 Heisenberg picture, 7
Eigenvector, 2 Hermitial operator, 27
Electric susceptibility, 70 Hermitian conjugate, 2, 105
Electromagnetic field, 7, 82 Hermitian operator, 2
Electron annihilation operator, 69 Heterodyne detection, 111
Electron creation operator, 69 High-temperature transition superconduc-
Electron density, 69 tor, 121
Electron density matrix, 70
Electronic excited state, 69
Electronic ground state, 84, 108 I
Electronic state, 81, 88 Identity operator, 3
Electronic-state-selective measurement, 99 Imaginary part, 111
Electron-hole pair, 73 Impulsive absorption process, 87
Electron-phonon coupling, 104, 106 Impulsive Raman, 76
Electro-optical sampling, 72 Impulsive stimulated Raman scattering, 68
Energy gap, 82 Impulsive stimulated Raman scattering pro-
Envelope, 96 cess, 87
Equation of motion, 4 Initial state, 85
Equilibrium position, 51, 59 Inner product, 2
Error ellipse, 46 Interaction-free, 18
Even number state, 43 Interaction Hamiltonian, 9, 82
Excitation density, 73 Interaction picture, 7
Excitation energy, 109
Excited state, 12
Expectation value, 4, 7 K
Ket vector, 1
Kinetic energy, 54
F Klemens channel, 74
Femtosecond laser pulse, 70
Fermi level, 73
Fluctuation, 76 L
Fock state, 29 Laser-plasma X-rays, 75
Force, 59 Lattice vibration, 53
Fourier transformation, 56 Lifetime, 122
Fourier-transform limited, 83 Linear atomic chain, 51
Four-level system, 103 Linear diatomic chain, 58
Liouville-von Neumann equation, 13
Longitudinal optical phonon, 68
G Long-wavelength approximation, 62
Γ -point, 72 Lowest state, 29
Gaussian function, 31, 96
Gaussian integral, 97
Gaussian pulse, 110 M
General squeezed state, 48 Matrix, 12
Glauber state, 34 Matrix form, 18
Gravitational-wave detection, 76 Measurement, 3
Ground state, 12, 29 Michelson-type interferometer, 116
Index 141
V
T
Vacuum state, 29, 75
Taylor expansion, 42
Valence band, 73
Time-dependent coherent state, 39
Variance, 37, 76
Time-dependent density functional theory,
70 Vibrational period, 99
Time-dependent squeezed state, 43 Von Neumann equation, 13
Time-evolution operator, 5, 8, 16
Time-ordered exponential, 6, 83
Ti:sapphire oscillator, 70 W
Trace, 12 Wave function, 30
Transient optical transmission, 70 Wave packet, 39
Transient reflectivity, 68 Wave vector, 52
Transient transmittance, 117
Transition dipole, 83
Transition process, 84, 103 Z
Transparent condition, 68, 110, 115 Zero-phonon state, 84
Transverse acoustic phonon, 76 Zero-point energy, 29, 82