Krylov complexity for 1-matric quantum mechanics
Abstract
This paper investigates the notion of Krylov complexity, a measure of operator growth, within the framework of 1-matrix quantum mechanics (1-MQM). Krylov complexity quantifies how an operator evolves over time by expanding it in a series of nested commutators with the Hamiltonian. We analyze the Lanczos coefficients derived from the correlation function, revealing their linear growth even in this integrable system. This growth suggests a link to chaotic behavior, typically unexpected in integrable systems. Our findings in both ground and thermal states of 1-MQM provide new insights into the nature of complexity in quantum mechanical models and lay the groundwork for further studies in more complex holographic theories.
I Introduction
The notion of “complexity” is playing an increasingly important role in several physical contexts [1], from computational condensed matter to holographic spacetime [2]. This quantity should reflect how "complicated" a physical system is. In quantum mechanics, this complexity could relate to states compared to a basic reference state, to operators, or a mix of both [3]. In other words, this idea of growth is connected to how we understand complexity in quantum systems. Essentially, complexity measures how difficult it is to create a state from a basic starting point. There are two main ways to look at complexity:
-
•
How complex a state is.
-
•
How an operator spreads out over time.
Recently, there have been many discoveries in these areas [4, 5, 6, 7, 8, 9, 10, 11]. For more details, you can look at studies on quantum systems [12, 13, 14, 15], quantum field theories [16, 2, 17, 18], black hole physics [19, 20], as recent reviews of these topics. The main idea is to describe how entangled a state is using simple parts, like gates in a quantum circuit or tensor network. It is used for the state complexity, and then complexity is measured by the size of the smallest circuit that can represent the state using these parts. For more details, see [21].
Recent work has shifted focus from states to operator growth in many-body systems [22, 23]. The authors of reference [24] introduced a new way to understand complexity by looking at how operators evolve over time. Instead of using a fixed set of operators, they use a basis that changes with time. Starting with an initial operator they consider how it evolves in the Heisenberg picture, given by . This evolution can be expanded into a series using nested commutators with the Hamiltonian. These nested commutators act as building blocks for describing how the operator evolves. The concept of K-complexity from [24] measures the "effective dimension" or the size of the space that the evolving operator explores. Krylov space is defined as the linear span of nested commutators, where is the system’s Hamiltonian and is an operator in question. More precisely, the operator may be expanded in terms of nested operators as follows
(1) |
where . Given a proper inner product in the space of operators, these nested operators are not orthogonal or normalized. However, we can construct an orthogonal and normalized basis called the Krylov basis. This is done using the Gram-Schmidt process.
Starting with an autocorrelation function of a simple local operator , you can use the recursion method to find Lanczos coefficients . These coefficients describe how the operator grows in the Krylov subspace. The original work [24] suggested the universal operator growth hypothesis, which links the long-term behavior of to the dynamics of the system. There is non-trivial evidence supporting the connection between the behavior of and integrability/chaos, yet it does not seem to be universal. In [24], authors explore the possible connection of the Krylov complexity with the circuit complexity.
We investigated aspects of Krylov complexity in a system of 1-matrix quantum mechanics (1-MQM). The notion of Krylov complexity has the advantage that it is well-defined for the class of quantum mechanical theories which appear in the context of the AdS/CFT correspondence, in particular large N gauge theories in various dimensions. We explored aspects of Krylov complexity in the model of quantum mechanics of a single Hermitian matrix. Using the mapping at large N to a gas of non-interacting fermions moving in an external potential We are able to compute the 2-point function of single-trace operators in this model and from that extract the asymptotic form of the so-called Lanczos coefficients. While the system is integrable these coefficients display linear growth which in the literature has been conjectured to be related to chaotic features. The work is an important first step towards the goal of computing Krylov complexity in more realistic holographic theories, involving more than one matrix, for example, the BFSS matrix model.
We study the notion of Krylov complexity in both the ground state and thermal state. In the ground state, we see the linear growth of the Lancsoz coefficients albeit the theory is integrable. In the thermal state, we find that the Lancsoz coefficients contain the even and odd linear branches. Moreover, we will find the radius of convergence when we are using the correlator to find the Krylov complexity. Till that point, we see only growth of the Krylov complexity. It is almost the same point as the first peak of the correlation function.
The structure of the paper is as follows: In Chapter 2, we review the concept of the Kyrilov complexity. After that in Chapter 3, we study the behavior of the Krylov complexity and Lancsoz coefficients for the number of decoupled harmonic oscillators. In Chapter 4, we review the basics of the 1-MQM and find the correlators of the theory both in the ground state and thermal one. In Chapter 5, we discuss the notion of complexity in 1-MQM. And in the end, in Chapter 6, we discuss the radius of convergence of the Krylov complexity.
II Krylov Complexity
We start with the definition of the notion of the Krylov complexity. It is defined as the recursion method in [25] and recently has been used in [24].
II.1 Krylov state complexity
Consider a quantum system with a time-independent Hamiltonian . A state is time evolved under the Schrodinger equation . Its solution has a formal power series expansion
(2) |
while . The time-evolved state is a linear combination of
(3) |
The subspace which is spanned by (3) is called Krylov subspace. Notice that in general, this basis is not orthogonal. The Gram-Schmidt procedure applied to generate an orthogonal basis when we define for one subspace of the full Hilbert space explored by the evolution of . In general, this code subspace can be infinite dimension.
Using the ordinary inner product, one can orthogonalize the basis (3) through the Lanczos algorithm:
-
1.
-
2.
-
3.
For ,
-
4.
Set
-
5.
If stop, otherwise set , and go to step 3 [26].
In the case that is finite, the Lanczos algorithm will end at some point that . The result of the Lanczos algorithm is two sets of Lanczos coefficients and .
We can expand the time-evolved state in terms of the Krylov basis as
(4) |
by substituting it into the Schrodinger equation, one gets
(5) |
and the initial condition is .
The Krylov state complexity of the state is defined as
(6) |
II.2 Krylov operator complexity
Similar to the Krylov state complexity, we can define Krylov complexity for quantum operators. Motivated by the time evolution of the operators , one can create the Krylov basis for a given operator in terms of the nested commutators with the Hamiltonian as they determine the time Taylor expansion of the Heisenberg operator.
Consider a time-independent Hamiltonian of a quantum system and a given Hermitian operator . The operator undergoing a Heisenberg evolution
(7) |
Just as states evolved under the Hamiltonian operator, operators evolved under the Liouvillian operator
(8) |
This is a linear combination of the sequence of operators
(9) |
where stands for [27]. The linear span of operators forms an invariant subspace . A convenient way to study the growth of a simple operator is to realize them as states, , and to introduce a notion of an inner product. It can be any non-degenerate inner product in the operator algebra such as the trace inner product for finite-dimensional Hilbert space (also known as infinite temperature inner product or Frobenius norm)
(10) |
and we write for the norm [28]. Thereby any operator within this subspace can be thought of as a vector in the linear vector space. Such a vector space endowed with a valid inner product is called the Krylov subspace.
The set of operators (9) are not orthogonal. The idea is to apply the Gram-Schmidt to orthogonalize it. As in the case of the state complexity, it is called Lanczos algorithm. It is as follows
-
1.
-
2.
-
3.
For :
-
4.
Set
-
5.
If stop; otherwise set and go to step3 [26].
The output of the algorithm is a sequence of positive numbers, , called the Lanczos coefficients and an orthogonal set of operators called the Krylov basis.
The time-evolved operator can now be expanded on the Krylov basis
(11) |
where can be thought of as the wavefunction over the Krylov basis. From the orthogonality, we obtain
(12) |
The time evolution of the operator follows
(13) |
thus via the Heisenberg equation satisfies the equation
(14) |
with boundary condition and . From unitarity, since the initial operator is normalized at the first step of the Lanczos algorithm, the wavefunction is normalized at all times .
Krylov complexity or K-complexity is defined as the time-dependent average position over the Krylov chain
(15) |
which can be viewed as the expectation value of the Krylov operator
(16) |
Intuitively, describes the mean width of a wavepacket in the Krylov space and hence quantitatively measures how the size of the operator increases as time goes by [29].
II.3 Krylov operator complexity over pure and mixed states
Given an normalized operator , by acting the operator on a pure state , one can construct a state
(17) |
The choice of pure state depends on the two-point function of the operator that we have in hand. For example, when we have the zero-temperature two-point function of , one can take the state to be the ground state of the theory.
A time-dependent state for can be expanded by . Although they do not create on an orthonormal basis. We need to apply the Gram-Schmidt procedure to make them orthogonal. By using it, one can obtain the Krylov basis such that as follows
(18) |
This construction of the basis is called Arnoldi iteration for general matrices. If is a Hermitian matrix, then (18) is simplified as
(19) |
(20) |
while . As before, this construction is called the Lanczos algorithm.
If is an eigenstate of , let us say , we have
(21) |
which is Hermitian. Assuming that and are Hermitian and we have an appropriate inner product by trace and Hermitian conjugation, we find that
(22) |
is the Hamiltonian eigenvalue in the absence of , which would be not directly related to the spreads of operators. On the other hand, , especially at large , represents how much the operator spreads into an orthogonal direction in the Hilbert space at a later time.
By introducing an inner product between operators at finite temperature one can generalize the above procedure
(23) |
where is the inverse temperature. We define
(24) |
Once the inner product is defined, one can construct the Krylov basis as . On top of it
(25) |
which is Hermitian. Hence, one can use the Lanczos algorithm instead of the Arnoldi iteration. For mixed states, it is more convenient that define the Lanczos coefficients in terms of operators as
(26) |
One can also obtain
(27) |
In the zero temperature limit, this reduces to the Lanczos algorithm for the pure state case when is the ground state of the theory and in the infinite temperature limit, it reaches the discussion of the Krylov operator complexity [30].
II.4 Recursion Method and Moment Expansion
This part is mostly based on [25].
The dynamical behavior of a quantum system is determined by its Hamiltonian and the operator representing the observable we’re interested in tracking over time. Our objective is to compute the dynamical correlation function , which provides insights into how evolves with time. Here, we assume that the correlators are even in the time. This evolution is governed by the Heisenberg equation of motion:
(28) |
Here, the commutator , known as the quantum Liouvillian operator , plays a crucial role. It’s a Hermitian superoperator. The formal solution to the equation of motion is expressed as:
(29) |
To implement the recursion method effectively, besides and , we need to define an inner product for operators within the Hilbert space associated with and . This choice influences the nature of the resulting dynamic correlation function.
The heart of the Liouvillian representation in the recursion method lies in the orthogonal expansion of the observable under examination:
(30) |
For classical systems, comprises an orthonormal set of functions in phase space. In contrast, for quantum systems, it constitutes an orthonormal set of operators. Regardless, these sets span a Hilbert space, typically of infinite dimensionality. The Liouvillian operator acts on the vectors within this space. The orthogonal expansion is executed in two successive steps.
-
•
Determine a particular orthogonal basis in the Hilbert space of the dynamical variables by applying the Gram-Schmidt procedure with the Liouvillian as the generator of the new direction.
-
•
Insert the expansion (30) into the equation of motion to obtain a set of differential equations for the time-dependent coefficients .
As the first step, we note that the general inner product between the vectors and for arbitrary vanishes
(31) |
This simplifies the Gram-Schmidt orthogonalization process and results in the subsequent set of recurrence relations for the vectors
(32) |
(33) |
with and . The sequence of numbers contains all the information for the reconstruction of the fluctuation function .
In the second step, we plug in the orthogonal expansion (30) into the equation of motion. The differential operator acts on the and the Liouvillian acts on the , which yields the following set of coupled linear differential equations for the function :
(34) |
with , . Unlike the vectors , the functions can not be determined recursively.
If our goal is to determine the fluctuation function of the dynamical variable , then it is sufficient to know just one of the functions . Follows directly from the orthogonal expansion
(35) |
There is a way to calculate the sequence for specific correlation functions of a given model system. It is called the moment expansion. The normalized fluctuation function can be expanded in a Taylor series
(36) |
with . The coefficients are the frequency moments of the normalized spectral density
(37) |
while
(38) |
for a given set of moments with the first coefficients are determined by
(39) |
for and , and with set values ,, .
The set of coefficients is equivalent to the square of the set of as discussed earlier.
Recursion method of the quantum Hamiltonian system in its ground state:
This application of the recursion method is tailored for investigating dynamic correlation functions within the quantum Hamiltonian system’s ground state. An essential preliminary step in more practical scenarios involves identifying the ground state wave function of the system.
For a given quantum Hamiltonian and its ground state wave function , our goal is to determine the normalized correlation function of the dynamical variable represented by the Hermitian operator
(40) |
In such a case the result of the Lanczos algorithm is two sets of coefficients and . The relation between the moments and these sets of coefficients are most conveniently expressed in terms of two arrays of auxiliary quantities and :
Given a set of moments , the coefficients and are obtained by initializing
(41) |
for and then applying the recursion relations [30]
(42) |
for and . The resulting coefficients are
(43) |
III Simple example: Krylov Complexity of free field theory
For a single harmonic oscillator, the Euclidean two-point function must obey the equation
(44) |
The solution to this equation is
(45) |
which can also be found by using the path integral method. To find the finite-temperature two-point function, one can use the method of images
(46) |
For simplicity, consider the case that , then we have
(47) |
Therefore, the thermal correlator is given as
(48) |
In order to find the complexity we can use the inner product which can be motivated or inspired by a two-sided correlator on the TFD state or KMS inner product as
(49) |
To find the inner product between the single harmonic oscillator and its time-shifted we can use the thermal two-point function and shift the time as
(50) |
considering a free quantum field on a circle of length . We can expand it in modes and get a collection of harmonic oscillators with frequency . In the following, we consider a J number of modes over the ground state and thermal states respectively.
III.1 Krylov complexity of the operator over the ground state
The correlator for different modes of harmonic oscillator in the ground state is
(51) |
while in the normalization factor such that . The moments are
(52) |
Here both sets of odd and even moments are nonzero, thus we get the nonzero values for both sets of and . In Fig. 1, one can find the non-zero value of and for different value of .
III.2 Krylov complexity of the operator over the thermal state
To find the correlator we use the inner product defined in (49). The correlator for different mode of the harmonic oscillator in the thermal state with inverse temperature is given by
(53) |
while
(54) |
and the normalization factor such that . The moments are
(55) |
One can calculate the Lanczos coefficients using (43). As it is clear and thus
(56) |
In Fig. 2 and 3, one can see the behavior of the non-zero for different values of and . In general, in this case, has two branches. For small , it increases linearly, and at some point, it starts to decrease and goes to zero. The number of non-zero valued increases as we increase the and it is almost twice the value of for this range of . Considering both positive and negative modes in the thermal case, the number of non-zero is equal to the number of different modes (in this case 2J). As increases, the linear behavior of the plots is dominant, and for in Fig. 3 one can see that we just have two linear branches. Moreover, by increasing the the branches get more separated, and in the high value of , it means the small value of the second branch is getting to vanish and we will reach the one linear branch as in the ground state. However, for a fixed , the slopes of two branches remain constant. As one can see in Fig. 2 the linear growth part of the plots for different are on top of each other.
In Fig. 4, one can see the behavior of when . It contains two linear branches and the slopes of two branches for different values of are different. Finally, in Fig. 5, one can see the behavior of the Krylov complexity for different values of . As correlators are periodic in time, the Krylov complexity is also periodic with period of .
In [31], the authors discussed the simple example of harmonic oscillator analytically. In particular, they find that for a very generic choice for the frequency of the modes, only the first Lancsoz coefficients are nonzero. For just one harmonic oscillator the theory describes with the Hamiltonian
(57) |
while . The position operator can be written in terms of the creation and annihilation operator . The calculation for the momentum operator is similar to the position operator. From the partition function, one can include the normalization factor
(58) |
we find
(59) |
One can apply the Lanczos algorithm starting from a normalized operator
(60) |
The first recursion gives
(61) |
while
(62) |
Then the second operator in the recursion actually vanishes . So the harmonic oscillator is a rather trivial model with the Lanczos algorithm terminating at the second step.
To generalize this case, consider a quantum system of decoupled harmonic oscillators of different frequencies
(63) |
with a properly normalized initial operator
(64) |
It is easy to compute the moments in the case
(65) |
The determinant in
(66) |
vanishes for
(67) |
so the Lanczos algorithm terminates at the steps with .
For a more general discussion on free theory, one can look at [32]. They consider free massive scalar and Dirac fermion in spacetime dimension. In the first case, Lancsoz coefficients split into even and odd branches, growing linearly with albeit with different intercepts. grows linearly with the universal slope, but even and odd branches have different finite terms. In the second case of free massless fermions is not an even function, hence besides , Lancsoz coefficients also include . In [32], one can see the numerical results as a function of .
They also consider a CFT on a sphere and calculate Lancsoz coefficients and Krylov complexity associated with the thermal two-point function of the model. They consider free massless scalar compacted on a . The corresponding two-point function has some singularity on the imaginary time axis. The correlator is in terms of a parameter which is the radius of measured in the units of which is the radius of . Their numerical results are good for . The Lanczos coefficients split into even and odd branches which grow linearly with but with different slopes. The same as our results in the thermal case. The behavior of Krylov complexity is the same as the of our results (see Fig. 5).
IV Matrix Quantum Mechanics
This chapter is based on [33, 34]. The one dimension takes to be timelike and the Lagrangian defines the theory
(68) |
where is a Hermitian matrix variable. The Lagrangian is invariant under a global symmetry, with the conserved angular momentum
(69) |
The quantum theory then is defined by the Hamiltonian
(70) |
and we restrict ourselves to the singlet sector .
A set of basic singlet vertex operators is given by
(71) |
To work with the collective field theory approach, a natural set of singlet operators is given by the vertex operators
(72) |
and one considers the collective field as its Fourier transform
(73) |
In terms of the eigenvalues
(74) |
one has and thus
(75) |
is simply the density of eigenvalues . The Collective field is constrained by
(76) |
and other constrained which disappear as .
To reformulate the theory with as the coordinate, one not only needs to change variables in the Hamiltonian but also to rescale the wavefunctions by the Jacobian of the transformation from to . While the Jacobian is singular for finite , the may be found from the hermiticity of the Hamiltonian. One can compute
(77) |
One can easily verify the following useful identity
(78) |
The Fourier transform of is the singular form
(79) |
In the end, one can write down the following field theory Hamiltonian
(80) |
where is the momentum conjugation to , and represent a multiplier for the density constraint and we also have some additional singular terms associated with the derivative terms. The kinetic energy piece is local. The effective potential is given by
(81) |
One can evaluate the integral and find
(82) |
We also have two other terms which are of lower order
(83) |
They do not contribute to the planar limit but begin to contribute in the first torus correction.
We should find the classical equation of motion. Since the constraints (76) should satisfy, the ground state has and in the leading order minimize
(84) |
This gives
(85) |
where is the point at which the square root vanishes. the planar ground state energy is then given by
(86) |
We now proceed to the computation of the propagator. This corresponds to the study of fluctuations in the collective field method. By shifting the field
(87) |
the propagator is determined by the quadratic action
(88) |
It is convenient to introduce a new variable as
(89) |
For a classical particle moving in the potential , is the time taken for the particle to go from the origin to the point . The range of is given by where is the time period of the classical motion and it is determined by
(90) |
where are the turning points of the classical motion. by redefining the field variable
(91) |
we will give
(92) |
Notice that the background field has disappeared. The only remnant is the new integration region for the variable . With the further transformation
(93) |
the action is brought into the form
(94) |
the propagator of the scalar field are obtained by implementing the constraint
(95) |
which leads to the Dirichlet boundary condition on . The small fluctuation eigenfunctions are found to be
(96) |
with the frequencies
(97) |
The propagator is then
(98) |
To find the two-point function in the matrix model, we have
(99) |
In terms of the collective field, one can find that
(100) |
therefore
(101) |
By substituting
(102) |
we reach to
(103) |
and thus
(104) |
and the connected two-point function in terms of the propagator in (98) can be written as
(105) |
To evaluate the integration over in (98) for we can take the integration over upper half plane and in the case of over the lower half plane and we will find that
(106) |
From now on we assume that and thus
(107) |
therefore
(108) |
IV.1 Quadratic Potential
We start with the free theory. Taking , we have
(109) |
thus the variable in terms of can be find as
(110) |
We have and so
(111) |
In the end, for free theory, we find that
(112) |
For some value of , the result is as below
(113) |
IV.2 Quatric Potential
Now let us consider the interacting theory, the simplest potential is
(114) |
and we set here. Hence, we have
(115) |
by change of variable we will reach to
(116) |
where
(117) |
is the elliptic integral of the first kind. The turning point of the classical particle is at . Therefore
(118) |
where
(119) |
is the complete elliptic integral of the first kind. Solving in terms of , one can find that
(120) |
where is the Jacobi elliptic function.
In order to find the connected two-point function, we need to calculate . To proceed, we can use the series definition of The Jacobi elliptic function
(121) |
In our case
(122) |
and . Let us first calculate the two-point function for the singlet . Thus, we have
(123) |
We have
(124) |
and in the end
(125) |
while
(126) |
Finally, we reach to
(127) |
V Krylov complexity for 1-MQM via the Lanczos algorithm
Now it is time to attempt to find the notion of Krylov complexity for the 1-MQM.
V.1 Over the Ground State
The correlator in the ground state is
(128) |
while is the normalization factor such that . The moments are given by
(129) |
In Fig. 6, one can see the moments and Lanczos coefficients of the 1-MQM in the ground state for different values of .
The Lanczos coefficients have a linear behavior in that the absolute value of the slope increases for higher values of the parameter. The slopes of the coefficients are negative while in the case of , they are positive.
Finally, in Fig. 7, one can find the Krylov complexity for different values of . The peak of the complexity grows while g increases. The complexity is periodic as the correlation function is periodic. However, we should consider the behavior of the complexity as a function of time for the time less than the radius of convergence in the time direction. The period of complexity is related to the and it decreases while increases and it is expected that it saturates for infinite .
V.2 Over the Thermal State
The correlator for the inner product (49) at inverse temperature is
(130) |
Thus, the moments are given by
(131) |
As it is obvious from the formula odd moments are zero and thus the set of the Lanczos coefficients
(132) |
In Fig. 8, one can see the plots for the set of and the even moments in this case. The coefficients have two linear branches with two different positive slopes. One of the slopes is almost the same for different values of while another slope increases while the parameter grows. In other words, one slope is a function of while another one is constant and -independent. (Look at the example in chapter 4 in [25])
In Fig. 9 and 10, one can see the Krylov complexity of the 1-MQM over the thermal states for different values of and . For a fixed , the periodicity of the Krylov complexity decreases as increases. In this case, unlike over the ground state, the peak of the complexity remains the same for different values of and . Moreover, for the fixed , the periodicity remains the same for different values of .
VI Toda chain flow in Krylov space and radius of convergence of Krylov complexity
VI.1 Toda chain flow in Krylov space
This section is mostly based on [35]. They begin by reviewing the basics of the recursion method. First, we start with the time-correlation function of some operator ,
(133) |
it is defined based on the Hermitian form in the space of operators
(134) |
here are some hermitian positive semi-definite operators which commute with the Hamiltonian H.
It is convenient to introduce
(135) |
such that
(136) |
In [35], authors focus on the Euclidean time evolution. For a given where is Euclidean time, an operator evolved in Minkowski time is .
The adjoin action of in the Krylov basis can be represented by Jacobi matrix L,
(137) |
(138) |
(139) |
As a generalization of (136) we define
(140) |
and evolution in terms of the Lanczos coefficient
(141) |
the original correlation function is then
(142) |
Lanczos coefficients can be promoted to be t-dependent.
Therefore, we can apply the recursion method to define the Krylov basis starting from the same initial for any given value of t. This defines the orthogonal basis ,
(143) |
where and are now t-dependent and thus and are time-dependent as well.
An important observation is that and written in terms of two different bases . They are related by a change of coordinates
(144) |
The basis has been transformed into basis by the matrix . One can express in terms of
(145) |
Explicit time dependence of G(t) provides that
(146) |
It follows that satisfies the Toda equation. The relation between , and is given by
(147) |
(148) |
One can introduce . In particular
(149) |
which says that the time-correlation function analytically continued to Euclidean time is a tau-function of the Toda hierarchy.
Furthermore, since and are orthonormal bases, they must be related by an orthogonal transformation ( for more detail look at [35])
(150) |
Evolving this equation in time by . We find
(151) |
This QR decomposition of [36].
In [35], the authors apply the relation of Lanczos coefficients to the Toda chain to clarify chaos in quantum many-body systems. An accurate counting of nested commutators appearing in the Taylor series expansion of will be singular at some finite .
In general, chaotic behavior is reflected by the linear growth of both and . While the slope of can not exceed twice the slope of .
To study the singular behavior of the time-correlation function, we assume that together with its derivatives are smooth functions for , and diverges at . From here follows that are regular for . Using QR decomposition, we have
(152) |
and conclude that is regular for and diverge at . We can decompose into orthogonal Krylov basis
(153) |
while
(154) |
This is a manifestation of delocalization in Krylov space. At , the operator spreads across the whole Krylov space.
Just note that this singularity is along the imaginary axis as we consider the Euclidean time.
VI.2 Radius of convergence of the Krylov complexity
The Krylov complexity is defined in (15) can be written in terms of . Thus, in case that in our calculation, the coefficients are regular in the time-band . Thus the Krylov complexity is also regular in this time band and the radius of convergence for Krylov complexity is , while is the radius of convergence of the correlation function.
Now in the 1-MQM model we consider in this project, first we should find the radius f convergence of the correlator.
For a given series , the series converge if
(155) |
In the case of 1-MQM over the ground state we have
(156) |
therefore
(157) |
After analytically continuation of t, for a complex , one can get
(158) |
Therefore
(159) |
Thus the radius of convergence of correlation function is at most
(160) |
In Fig. 7, one can see that the first peak of the Krylov complexity for all values of is approximately at . The radius of convergence of the Krylov complexity is discussed in the previous part.
VII Discussion
In this paper, we investigate the notion of Krylov complexity for 1-MQM which is a toy model of the gauge theory dual of an AdS black hole in both ground states and thermal states. In both cases, we observe the linear growth. However, over the thermal states, it divides into odd and even branches. This work is a warm-up for the more realistic examples of holography as BFSS. Even studying the n-matrix quantum mechanics can be much more challenging as the system is not solvable, thus, we treat the chaotic systems. Since the theory is not solvable, it is a very hard task to find the correlation functions in the theory. Here albeit the 1-MQM is not chaotic, the Lancsoz coefficients have the linear growth behavior. However, in the literature, it has been conjectured that the linear growth of the Lancsoz coefficients is a sign of chaos [24]. It seems that the conjecture is not universal and it should be corrected in a way that only if we have a chaotic system, the corresponding Lancsoz coefficients have linear growth. Furthermore, in the case of the thermal state both in the harmonic oscillator and 1-MQM we see that the coefficients are divided in the odd and even branches with linear growth. The reason why this happens is not clear but it might be because of using the new inner product (49) as it is the case also in [32]. Moreover, it remains an open question: if it is possible in some scenarios that we have even more than 2 branches. The result that we find for the complexity is periodic since the correlators are periodic. However, we calculate the radius of convergence of the Krylov complexity where it is half of the radius of convergence of the correlator. From the numerics, it is obvious that this time is the same point as the first peak of the Krylov complexity. Thus, in the radius of convergence we see only the growth of Krylov complexity. Nevertheless, finding the Krylov complexity after this point is an open question.
Acknowledgements.– I would like to thank M. Alishahiha for useful discussions and communication. In particular, I would like to thank my supervisor, K. Papadodimas for the useful discussion and comment on the draft. I would like also to thank my co-supervisor M. Brtolini for his support during this work. I would like to thank CERN-TH for its hospitality during the preparation of this work. The research is partially supported by the INFN Iniziativa Specifica- String Theory and Fundamental Interactions project.
Appendix A Collective field theory formalism
Broadly, the collective approach involves a variable transformation. Consider an operator Hamiltonian
(161) |
in a manner that allows its representation using an infinite set of new variables
(162) |
This set would be generally over-complete for finite . One can make a standard canonical transformation and express the theory using . Thus, the wave function of the theory should be written in terms of . this can come about as a restriction on invariant singlet subspace of the full Hilbert space. On the wave functional, the kinetic term takes the form
(163) |
where
(164) |
The kinetic term in the new collective representation is not Hermitian. It is because of the fact that the new scalar product involves a Jacobian.
Using a similar transformation, one finds the following Hermitian Hamiltonian
(165) |
whith and being a conjugate set of fields variables when
References
- Nielsen and Chuang [2010] M. A. Nielsen and I. L. Chuang, Quantum computation and quantum information (Cambridge university press, 2010).
- Susskind [2020] L. Susskind, Three lectures on complexity and black holes, Tech. Rep. (Springer, 2020).
- Rabinovici et al. [2022] E. Rabinovici, A. Sánchez-Garrido, R. Shir, and J. Sonner, Journal of High Energy Physics 2022, 1 (2022).
- Balasubramanian et al. [2022] V. Balasubramanian, P. Caputa, J. M. Magan, and Q. Wu, Physical Review D 106, 046007 (2022).
- Caputa and Liu [2022] P. Caputa and S. Liu, Physical Review B 106, 195125 (2022).
- Bhattacharjee et al. [2023] B. Bhattacharjee, X. Cao, P. Nandy, and T. Pathak, Journal of High Energy Physics 2023, 1 (2023).
- Caputa et al. [2023] P. Caputa, N. Gupta, S. S. Haque, S. Liu, J. Murugan, and H. J. Van Zyl, Journal of High Energy Physics 2023, 1 (2023).
- Afrasiar et al. [2023] M. Afrasiar, J. K. Basak, B. Dey, K. Pal, and K. Pal, Journal of Statistical Mechanics: Theory and Experiment 2023, 103101 (2023).
- Alishahiha [2023] M. Alishahiha, Physics Letters B 842, 137979 (2023).
- Vasli et al. [2024] M. J. Vasli, K. Babaei Velni, M. R. Mohammadi Mozaffar, A. Mollabashi, and M. Alishahiha, Eur. Phys. J. C 84, 235 (2024), arXiv:2307.08307 [hep-th] .
- Alishahiha and Banerjee [2023] M. Alishahiha and S. Banerjee, SciPost Phys. 15, 080 (2023), arXiv:2212.10583 [hep-th] .
- Jefferson and Myers [2017] R. A. Jefferson and R. C. Myers, Journal of High Energy Physics 2017, 1 (2017).
- Chapman et al. [2019] S. Chapman, J. Eisert, L. Hackl, M. P. Heller, R. Jefferson, H. Marrochio, and R. Myers, SciPost physics 6, 034 (2019).
- Caceres et al. [2020] E. Caceres, S. Chapman, J. D. Couch, J. P. Hernandez, R. C. Myers, and S.-M. Ruan, Journal of High Energy Physics 2020, 1 (2020).
- Chagnet et al. [2022] N. Chagnet, S. Chapman, J. de Boer, and C. Zukowski, Physical review letters 128, 051601 (2022).
- Stanford and Susskind [2014] D. Stanford and L. Susskind, Physical Review D 90, 126007 (2014).
- Susskind [2018] L. Susskind, arXiv preprint arXiv:1802.01198 (2018).
- Chattopadhyay et al. [2023] A. Chattopadhyay, A. Mitra, and H. J. van Zyl, Physical Review D 108, 025013 (2023).
- Chapman and Policastro [2022] S. Chapman and G. Policastro, The European Physical Journal C 82, 128 (2022).
- Mück and Yang [2022] W. Mück and Y. Yang, Nuclear Physics B 984, 115948 (2022).
- Nielsen [2005] M. A. Nielsen, arXiv preprint quant-ph/0502070 (2005).
- von Keyserlingk et al. [2018] C. W. von Keyserlingk, T. Rakovszky, F. Pollmann, and S. L. Sondhi, Physical Review X 8, 021013 (2018).
- Chan et al. [2018] A. Chan, A. De Luca, and J. T. Chalker, Physical Review X 8, 041019 (2018).
- Parker et al. [2019] D. E. Parker, X. Cao, A. Avdoshkin, T. Scaffidi, and E. Altman, Physical Review X 9, 041017 (2019).
- Viswanath and Müller [1994] V. Viswanath and G. Müller, The recursion method: application to many body dynamics, Vol. 23 (Springer Science & Business Media, 1994).
- Hashimoto et al. [2023] K. Hashimoto, K. Murata, N. Tanahashi, and R. Watanabe, Journal of High Energy Physics 2023, 1 (2023).
- Kundu et al. [2023] A. Kundu, V. Malvimat, and R. Sinha, Journal of High Energy Physics 2023, 1 (2023).
- Bhattacharya et al. [2023] A. Bhattacharya, P. Nandy, P. P. Nath, and H. Sahu, Journal of High Energy Physics 2023, 1 (2023).
- Lv et al. [2023] C. Lv, R. Zhang, and Q. Zhou, arXiv preprint arXiv:2303.07343 (2023).
- Iizuka and Nishida [2023] N. Iizuka and M. Nishida, Journal of High Energy Physics 2023, 1 (2023).
- Du and Huang [2022] B.-n. Du and M.-X. Huang, arXiv preprint arXiv:2212.02926 (2022).
- Avdoshkin et al. [2022] A. Avdoshkin, A. Dymarsky, and M. Smolkin, arXiv preprint arXiv:2212.14429 (2022).
- Das [1992] S. R. Das, Spring school on superstrings, Trieste, Italy , 172 (1992).
- Das and Jevicki [1990] S. R. Das and A. Jevicki, Modern Physics Letters A 5, 1639 (1990).
- Dymarsky and Gorsky [2019] A. Dymarsky and A. Gorsky, arXiv preprint arXiv:1912.12227 (2019).
- Symes [1980] W. Symes, Physica D: Nonlinear Phenomena 1, 339 (1980).