Open access peer-reviewed chapter

Quantum Algorithms for Nonlinear Equations in Fluid Mechanics

Written By

Rene Steijl

Reviewed: 13 November 2020 Published: 10 December 2020

DOI: 10.5772/intechopen.95023

From the Edited Volume

Quantum Computing and Communications

Edited by Yongli Zhao

Chapter metrics overview

639 Chapter Downloads

View Full Metrics

Abstract

In recent years, significant progress has been made in the development of quantum algorithms for linear ordinary differential equations as well as linear partial differential equations. There has not been similar progress in the development of quantum algorithms for nonlinear differential equations. In the present work, the focus is on nonlinear partial differential equations arising as governing equations in fluid mechanics. First, the key challenges related to nonlinear equations in the context of quantum computing are discussed. Then, as the main contribution of this work, quantum circuits are presented that represent the nonlinear convection terms in the Navier–Stokes equations. The quantum algorithms introduced use encoding in the computational basis, and employ arithmetic based on the Quantum Fourier Transform. Furthermore, a floating-point type data representation is used instead of the fixed-point representation typically employed in quantum algorithms. A complexity analysis shows that even with the limited number of qubits available on current and near-term quantum computers (<100), nonlinear product terms can be computed with good accuracy. The importance of including sub-normal numbers in the floating-point quantum arithmetic is demonstrated for a representative example problem. Further development steps required to embed the introduced algorithms into larger-scale algorithms are discussed.

Keywords

  • partial differential equations
  • fluid mechanics
  • nonlinear equations
  • quantum Fourier transform
  • floating-point arithmetic

1. Introduction

Quantum computing [1] and quantum communication are research areas that have seen significant developments and progress in recent years, as is apparent from the work covered in this book. In this chapter, the focus is on the development of quantum algorithms for solving nonlinear differential equations, highlighting key challenges that arise from the non-linearity of the equations to be solved. For this application of quantum computing, progress has so far been relatively limited and in this work, a promising approach to deriving efficient quantum algorithms is proposed. Although the focus is on non-linear equations related to fluid mechanics, the approach put forward here is applicable to a much wider range of problems. Furthermore, in developing the proposed method, efficient quantum circuits involving floating-point arithmetic were created, in contrast to the more commonly used fixed-point arithmetic employed in a range of quantum algorithms. This aspect of the work described here should also be useful for a wider audience. In this work, the development of quantum algorithms for the nonlinear governing equations for fluid mechanics is described with a particular focus on representing the non-linear product terms in the equations. A key aspect of the derived quantum circuits in the present work is the (temporary) representation of the solution in the computational basis, along with the the use of a floating-point data representation in the arithmetic operations. The quantum circuits for obtaining the non-linear product terms are new developments and form the main contribution of this work. In recent years, a small number of works have considered quantum computing applications to fluid mechanics [2, 3, 4, 5, 6, 7, 8]. A brief review of this previous work will be presented in Section 2 and will provide context to the proposed approach. Related work on algorithms with representation in the computational basis is reviewed in this chapter. This chapter is structured as follows. Section 2 describes the background to the current work. Section 3 reviews the key challenges related to treating nonlinear differential equations in a quantum computing context, followed by a discussion of the nonlinear governing equations in fluids dynamics in Section 4. Section 5 then describes how nonlinear terms in governing equations can be evaluated in quantum algorithms using the computational basis. Section 6 and Section 7 discuss the quantum circuits used for computing the square of a floating-point number and the multiplication of two floating-point numbers, respectively. The simulation and verification of the derived quantum circuits is presented in Section 8. The complexity of the circuits is analyzed in Section 9. Finally, conclusions from this work and suggestions for further work are presented in Section 10.

Advertisement

2. Background of present work

For a small number of applications, quantum algorithms have been developed that display a significant speed-up relative to classical methods. Computational quantum chemistry is proving to be one of the key areas of application. Important developments for a wider range of applications include quantum algorithms for linear systems [9, 10] and the Poisson equation [11]. Applications to computational science and engineering problems beyond quantum chemistry have only recently begun to appear [4, 5, 6, 12, 13, 14]. Despite this research effort, progress in defining suitable engineering applications for quantum computers has been limited.

Significant progress has been made in recent years in the development of quantum algorithms for linear ordinary differential equations (ODEs) as well as linear partial differential equations (PDEs) [15, 16, 17, 18, 19]. However, in contract to this progress for linear equations, there has not been similar progress in the development of quantum algorithms for nonlinear ODEs and nonlinear PDEs. An early work by Leyton and Osborne [20] presented an innovative and highly ambitious algorithm. However, the computational complexity of this work involves exponential dependency on the time interval used in the time integration. A small number of more recent works have addressed nonlinear differential equations and typically algorithms for very specific problems were obtained [8]. Therefore, much research work is needed into quantum algorithms for a wider range of nonlinear problems.

Early work in quantum computing relevant to the field of Computational Fluid Dynamics (CFD) mainly involved the work on quantum lattice-gas models by Yepez and co-workers [2, 3]. This work typically used type-II quantum computers, consisting of a large lattice of small quantum computers interconnected in nearest neighbor fashion by classical communication channels. In contrast to these quantum lattice-gas based approaches, the present study focuses on quantum algorithms designed for near-future ‘universal’ quantum computers. The potential of quantum computing in the context of direct numerical simulation of flows was reviewed recently by Griffin et al. [7], showing that a number of further developments are needed to make this approach viable.

Typically, there are two methods of encoding the result of a quantum algorithm: encoding within the computational basis of the quantum state and encoding within the amplitudes of the quantum state. The widely-used Quantum Fourier Transform (QFT) uses the second approach. The QFT with complexity Olog2N for problem size N has exponential speed-up compared to the classical fast Fourier transform (complexity ONlogN) and plays an important role in quantum computation as an essential part of many quantum algorithms. The exponential speed-up realized is due to superposition and quantum parallelism. However, in same quantum algorithms, the Fourier coefficients may be needed in the computational basis [21].

Here, the two different encoding methods are illustrated using the discrete Fourier Transform (DFT). The QFT performs the DFT in terms of amplitudes as,

j=0N1xjjk=0N1ykkE1

The QFT performs a DFT on a list of complex numbers, and the result is stored as amplitudes of a quantum state vector. In order to extract the individual Fourier components, measurements need to be performed on the quantum state vector. Therefore, the QFT is not directly useful for determining the Fourier-transformed coefficients of the input state. However, the QFT is widely used as a subroutine in larger algorithms. In contrast to the amplitude encoding in Eq. (1), Zhou et al. [21] presented a quantum algorithm computing the Fourier transform in the computational basis (termed QFTC). This quantum algorithm encodes Fourier coefficients with fidelity 1δ and digit accuracy ε for each Fourier coefficient. Its time complexity depends polynomially on logN, and linearly on 1/δ and 1/ε. The QFTC, enables the Fourier-transformed coefficient to be encoded in the computational basis as follows,

k0kykE2

where yk corresponds to the fixed-point binary representation of yk11 using two’s complement format. In the algorithm proposed by Zhou et al. [21], the input vector x is provided by an oracle Ox such that,

Ox0=j=0N1xjjE3

which can be efficiently implemented if x is efficiently computable or by using the qRAM that takes complexity logN under certain conditions [21]. Comparing Eq. (1) and Eq. (2), it is clear that encoding in the computational basis requires a number of additional qubits depending on the required fixed-point representation.

Advertisement

3. Nonlinear problems on quantum computers

An early work by Leyton and Osborne [20] introduced a quantum algorithm to solve nonlinear differential equations with an unfavorable complexity. Since then, very few works have considered quantum algorithms for nonlinear equations. In contrast, algorithms for linear differential equations have continued to receive significant attention. As an example, advanced quantum spectral methods for differential equations were published recently by Childs and Liu [19].

A key contributing factor to the limited progress in algorithms for non-linear problems is the inherent linearity of quantum mechanics. For quantum algorithms encoding information as amplitudes of a quantum state vector, nonlinear (product) terms cannot be obtained by multiplying these amplitudes by themselves, as a result of the no-cloning theorem that prohibits the copying of an arbitrary quantum state. Furthermore, all quantum-gate operations (with the exception of measurements) in the quantum-circuit model used here need to be unitary and reversible. These requirements add further challenges to representing nonlinear terms when using the amplitude-based encoding approach. Specifically, in a normalized quantum state vector all amplitudes in the vector are 1 (unless only a single amplitude is non-zero), therefore an operator performing products of the amplitudes cannot be unitary since the resulting quantum state vector will no longer have a unit norm.

One possible way around these problems associated with nonlinear terms would be a hybrid quantum-classical approach where the nonlinear products are computed on a classical computer. However, due to the complexity introduced by measuring the quantum state (needed before each transfer of information to the classical computer) and the cost of (re-)initialization of the quantum computer with the result of these products, this is not a promising line of development. It is highly unlikely to lead to a quantum speed-up. Recently, Variational Quantum Computing (VQC) was introduced as an effective hybrid classical-quantum approach [22, 23], firstly for applications in quantum chemistry and more recently for a wider range of linear and nonlinear problems [24]. The VQC approach constructs the required solution from a layered network, as illustrated in Figure 1. As shown in Figure 1(a), multiple layers are used (4 in the illustration), each taking as input multiple qubits (6 in example shown). Using depth 5 in the example, the quantum circuits defined by Uλ involve 13 two-qubit gates as shown in Figure 1(b). Each of these gates has a parameter λi113 associated with it. A classical computer is used to create optimized parameters λ employing an iterative approach that takes the measured state of the ancilla qubit as input. A further key part of the approach is the problem-specific Quantum Nonlinear Processing Unit (QNPU). Recently, Lubasch et al. [24] published an example for the QNPU for the nonlinear Burgers equation. In applications of the VQC approach, the efficiency strongly depends on the choice of the number of parameters λ used in Uλ. The work by Lubasch et al. [24] showed that exponential speed-up is only possible if the depth of Uλ scales with the number of qubits and not with the overall problem size. It is clear that the proposed VQC approach is an important development toward QC applications to nonlinear problems. It therefore constitutes a leading candidate for applications to fluid dynamics. However, it is also clear that further investigation is needed to further assess its suitability for a range of applications.

Figure 1.

Illustration of the Variational Quantum computing (VQC) approach (adapted from Lubasch et al. [24]).

Advertisement

4. Nonlinear governing equations in fluid mechanics

The Navier–Stokes equations for an incompressible, Newtonian fluid can be written as,

Ut+UxU=1ρxp+νΔU;xU=0E4

where U, p, ρ and ν are the velocity, pressure, density and kinematic viscosity, respectively. x denotes the coordinate in space. The second term on the right-hand side of Eq. (4) is the nonlinear convection term that poses a key challenge to developing efficient quantum algorithms for the Navier–Stokes equations. Efficient quantum algorithms for linear convection equations discretized on regular Cartesian meshes with periodic boundary conditions have been devised in recent years [6]. When studying numerical methods for the Navier–Stokes equations, it is often useful to switch to Burgers’ model equation, to obtain a single nonlinear partial differential equation that retains a nonlinear convection term similar to the Navier–Stokes equations. Using the VQC approach, Lubasch and co-workers recently published example quantum circuits to model the Burgers equation [24]. Griffin et al. [7] discuss two approaches for treating the nonlinear term in the Navier–Stokes equations: the VCQ approach of Lubasch et al. [24] and a linearized approach. These authors conclude that, at present, VQC represents the most promising approach for Navier–Stokes equations. Their study also highlights that much further research work is needed to create efficient algorithms for fluid dynamics applications. It is relatively easy to show that the linearization approach to solving non-linear governing equations on a Quantum Computer is generally unfeasible. In applying linearization to nonlinear governing equations, the idea is to use a linearization about the present state of the solution, and then advance this linearized problem in time. This creates a linearization error, which is small if the time step is small. However, even if this linearization error can be tolerated, the linearization approach is problematic in a quantum computing context. This is due to the need for repeated measuring of the quantum state (so that the gates that implement the linear operator may be updated with the current solution) and repeated re-initialization of the quantum state. The complexity associated with repeated measuring and re-initialization is so large that any benefit of a quantum algorithm over a classical algorithm is very likely to vanish.

The development of quantum algorithms for fluid dynamics is clearly at a very early stage and therefore it is essential that different approaches are considered.

Advertisement

5. Representing nonlinear terms in computational basis

In the present work, an alternative approach to introducing the nonlinear terms of nonlinear differential equations into a quantum algorithm is investigated. Specifically, the assumption is made that in a large-scale quantum algorithm for the solution of the nonlinear (partial) differential equations, the solution is encoded in terms of amplitude in the quantum state vector, i.e. the approach used in a wide range of algorithms including the QFT. Then, for the nonlinear terms of the equations, the following steps are suggested. First, within the larger quantum algorithm, a quantum algorithm is embedded that converts the solution from the quantum-amplitude representation to a representation in the computational basis. Recently, quantum algorithms for this ‘analog-to-digital conversion’ were published by Mitarai et al. [25]. Using the representation of the solution in the computational basis, the required nonlinear terms are then efficiently evaluated using quantum circuits presented later in this chapter. Once computed, a conversion back to quantum-amplitude representation is to be used, enabling the rest of the quantum algorithm to proceed. For this ‘digital-to-analog’ conversion, quantum algorithms were recently studied and published by SaiToh [26]. For the representation in the computational basis, a fixed-point approach is typically employed to represent real or complex numbers in quantum algorithms. The number of additional qubits required when using computational-basis encoding depends directly on the number of qubits required to represent the real and complex numbers needed in the algorithm. In the present work, a different approach is put forward: instead of using fixed-point arithmetic, a floating-point representation is used.

In the literature, quantum arithmetic using floating-point numbers has received very little attention so far. Haener et al. [27] described an investigation into quantum circuits for floating-point addition and multiplications and compared automatically generated circuits from Verilog implementations with hand-crafted optimized circuits. Their study provides evidence that floating-point arithmetic is a viable candidate for use in quantum computing, at least for typical scientific applications, where addition operations usually do not dominate the computation. Following on from these conclusions, the present work investigates the use of floating-point arithmetic as part of evaluating nonlinear terms in the computational basis.

5.1 Previous works on algorithms in computational basis

Quantum arithmetic in the computational basis constitutes an important component of many quantum algorithms, and as a result reversible implementations of algebraic functions (addition, multiplication, inverse, square root, etc.) have been widely studied. In contrast, there is relatively little work on quantum algorithm implementation of higher-level transcendental functions, such as logarithmic, exponential, trigonometric and inverse trigonometric functions. Examples of applications of trigonometric and inverse trigonometric functions in the computational basis can be found in the famous HHL algorithm [9] and in the state preparation algorithm introduced by Grover and Rudolph [28]. More recently, a quantum algorithm for approximating the QR decomposition of a N×N matrix in the computational basis was published by Ma et al. [29], with polynomial speed-up over the best classical algorithm.

5.2 Fixed-point and floating-point arithmetic

A fixed-point number held in an nq qubit register can be defined as the following quantum state,

w=wnint1wnint2w0integerw1wnintnqfractionalE5

where wj0,1, j=nintnq,nintnq+1,,0,,nint1 [30]. This state represents the number w=jwj2j. The nint leftmost qubits are used to represent the integer part of the number and the remaining nfrac=nqnint qubits represent its fractional part. In this example, no sign qubit is used so that only positive numbers can be represented (for most applications an additional sign qubit would be required). Since fewer than nq bits may suffice for the representation of the input, a number of the leftmost qubits in the register may be set to 0. Clearly, the fixed point system is very limited in terms of the size of the numbers it can store. Therefore, soon after computers were introduced for numerical computing the switch to floating-point arithmetic was made. In a computer implementation of a floating point number with base 2, a non-zero signed number x, defined through a normalized representation, is expressed as,

x=±S×2E,where1S<2E6

where the numbers S and E are the mantissa and the exponent, respectively. The binary expansion of the mantissa is

S=b0.b1b2b32withb0=1E7

Here, it is important to note that always b0=1 for non-zero numbers in a normalized representation. This will be used in the present work to achieve savings in the number of required qubits, as detailed later. In the binary representation, the bits following the binary point are the fractional part of the mantissa. Once floating-point numerical computation on classical computers became commonplace, the industry standard IEEE 754 was introduced [31]. A similar standard for floating-point representations on a quantum computer does not yet exist, but is desirable [30]. A key feature of the IEEE standard is that it requires correctly rounded operations: correctly rounded arithmetic operations, correctly rounded remainder and square root operations and correctly rounded format conversions. Typically, rounding to the nearest floating pointing number available in the destination (output register) is used. In the quantum circuits in the present work, rounding down to nearest is used, for reasons of simplicity. Detailed analysis of quantum-circuits developed here for squaring and multiplication operations shows that ‘correctly’ rounding to nearest involves a significant increase in circuit complexity (i.e. using quantum equivalents of guard and sticky bits, that are well established in arithmetic on classical computers [31]). A key aspect of the IEEE 754 that has been incorporated in the present work is the definition of sub-normal numbers. To illustrate the concept of subnormal numbers, the IEEE 754 standard representation of single format numbers using a 32-bit word is considered. The first bit is the sign bit, followed by 8 bits representing the exponent. Then, 23 bits are used to store a 24-bit representation of the mantissa, i.e. b0 is not stored. Numbers with exponent bits 000000002=010 and 111111112=25510 are defined special cases. The smallest normalized number is 1.002×2126=2126. Sub-normal numbers are used to represent smaller numbers, i.e. in this case the exponent field has a zero bit string but the fraction field has a nonzero bit string. Zero is represented with a zero bit string for the fractional field. For all subnormal numbers, the 00000000 used for the exponent represents 2126 and by using the 23 fractional field bits, equally-spaced numbers in the range 0.00012×2126 (with 22 zero bits after the binary point) to 0.11112×2126 (with 23 one bits after the binary point) are encoded.

5.3 Quantum floating-point format used in present work

Based on the floating point representation defined in the IEEE standard, the present work introduces a floating-point system with fewer bits (i.e. qubits in this case) than the 32 used for single format numbers. This is the direct result of the limited number of qubits available on current and near-term quantum computers. To optimize the range of floating-point numbers that can be represented with the approach used here, the following key aspects of the IEEE standard were adopted:

  • For the mantissa only the fractional part is stored,

  • Exponent bit strings 00002 and 11112 are used for special cases, i.e. dealing with 0, subnormal numbers as well as cases with overflow,

  • The remaining range of exponent bit strings is used for a range of exponential centred around 20=01112,

  • Sub-normal numbers are used to extend the range of small numbers,

  • Rounding down to nearest is used as rounding mode,

  • Only unsigned numbers are considered for simplicity. Signed numbers can easily be obtained by adding a further ‘sign’ qubit.

In this work, a floating-point number is represented as an nq=NM+NE quantum register. In the quantum-circuit implementation, the most significant (leftmost) mantissa qubit is not stored, using the hidden-bit approach used in the IEEE 754 standard. Therefore, NM1 qubits define the fractional part of the mantissa in the developed quantum circuits. NE defines the number of qubits used to define the exponent. In the following, examples with NE=3 and NE=4 and NM35 are considered. For NE=3, the number 1.00 is defined by 00011 when NM=3. Similarly, 0000111 defines the number 1.000 for NM=4 and NE=4. For NE=3, the smallest normalized number that can be represented is 1/4 independent of the number mantissa qubits. Then, exponent state 000 defines zero and sub-normal numbers, as shown in Table 1 for NM=3, NM=4 and NM=5.

NM=3NM=4NM=5
011000=3/160111000=7/3201111000=15/64
010000=1/80110000=3/1601110000=7/32
001000=1/160101000=5/3201101000=13/64
0100000=1/801100000=3/16
0011000=3/3201011000=11/64
0010000=1/1601010000=5/32
0001000=1/3201001000=9/64
00010000=1/32
00001000=1/64

Table 1.

Sub-normal numbers for floating-point numbers with 3 qubits as exponential.

Similarly, using 4 qubits for the exponent (NE=4) means that the smallest normalized number is 1/64. For NM=4 and NM=5, Table 2 shows the corresponding sub-normal numbers.

NM=4NM=5
01110000=7/512011110000=15/1024
01100000=3/256011100000=7/512
01010000=5/512011010000=13/1024
01000000=1/128011000000=3/256
00110000=3/512010110000=11/1024
00100000=1/256010100000=5/512
00010000=1/512010010000=9/1024
000100000=1/512
000010000=1/1024

Table 2.

Sub-normal numbers for floating-point numbers with 4 qubits as exponential.

In line with the IEEE 754 standard, exponent state 11 denotes numbers for which an overflow has occurred. For NE=3, the largest normalized number available is 111110 which equates to 14 and 15 for NM=3 and NM=4, respectively. Similarly, for NE=4, the largest normalized number available is 1111110 which equates to 240 and 248 for NM=4 and NM=5, respectively.

Advertisement

6. Quantum circuits for squaring floating-point numbers

For a floating-point number defined by NM mantissa and NE exponent bits, a total of NM1+NE qubits is needed to define the state in the quantum circuits introduced here. An example with NM=3 and NE=3 will now considered, using registers imb1imb0 and ieb2ieb1ieb1 to define the fractional part of the mantissa and the exponent of the input number, respectively. For the multiplication operation described later a second input floating-point number is defined using ima1ima0 and iea2iea1iea1. The output of the squaring and multiplication operations is a floating-point number r defined by imr1imr0 and ier2ier1ier1 (initialized at 0). In addition to the input and output registers, the quantum circuits will need additional qubits to hold results of intermediate results, e.g. for NM=3 a 6-qubit sub-register imp5imp0 is used. To facilitate the quantum-multiplication operations, a further ancilla qubit a0 is used. For quantum circuits without measures to deal with sub-normal numbers and overflow, the quantum state for NM=3 and NE=3 is defined in a 2×NM1+NE+2×NM+1=17-qubit register

ieb2ieb1ieb0imb1imb0a0imp5imp0ier2ier1ier0imr1imr0E8

For NM=4 and NE=4, the required number of qubits increases to 2×NM1+NE+2×NM+1=23. The quantum circuit performing the squaring operation for NM=3 and NE=3 is detailed here as example (in realistic applications NM>3 will typically be needed). Figure 2 shows the quantum circuit used in the first step of computing the square of a quantum floating point with NM=3 and NE=3. This step involves computing the square of the mantissa, with this result temporarily stored in imp5imp0. In this circuit, QFT6 prepares this temporary register for the three subsequent product steps denoted by P1, P2 and P3, involving doubly-controlled phase operations. Specifically, three-qubit gates are used applying a phase rotation conditional on state of a0 and either imb1 or imb0. The Pi steps are controlled-summation operations in the shift-and-add approach to computing the products, i.e. the circuits in Pi are derived from quantum adders controlled by an additional qubit. Once the controlled phase changes in the circuits P1, P2 and P3 have been applied, inverse QFT6 on imp5imp0 creates the desired output state. In case the square of the mantissa 2, i.e. imp5=1, the result exponent needs to be incremented by 1. This is achieved by apply a controlled-NOT to ier0 (which was initialized at 0) with imp5 as control. In the next step, result mantissa qubits imr1imr0 are set using temporary results in imp5imp0, where the required gate operations are conditional on the state of imp5. Then the steps shown in Figure 2 are ‘uncomputed’ so that the sub-register imp5imp0 is set to 0 again. The next step is illustrated in Figure 3, where the output exponent is obtained. This step involves the initialization of the temporary register imp3imp0 with 2×Eb (i.e. twice the input exponent). Then, the bias of 0112=3 is removed (denoted by 011). This bias removal uses two’s complement to create a modified modulo-5 adder that removes a value 0112=3 from imp4imp0. Then, the result exponent sub-register ier2ier1ier0 is prepared for the subsequent modulo-3 addition (denoted by MADD3) by applying QFT3. Next, the modulo-3 adder is used to add the qubits imp2imp1imp0 into ier2ier1ier0. By applying the inverse QFT3 on ier2ier1ier0 the required state is obtained. The remaining steps shown in the quantum circuit in Figure 3 are used to ‘uncompute’ and clean-up the temporary register, e.g. using inverse QFT3 and a modified modulo-5 adder to re-apply the bias 0112=3. The circuits described so far do not take into account the special situation arising from creating sub-normal numbers as output as well as cases with ‘overflow’ results. This is discussed next.

Figure 2.

Quantum circuit used to compute square of mantissa (for NM=3 and NE=3).

Figure 3.

Quantum circuit used to obtain exponent for squaring operation (NM=3 and NE=3).

For certain normalized input numbers the squaring operation leads to outputs truncated to 0 or to the non-zero sub-normal numbers discussed in Section 5.3. The quantum circuits discussed so far need to be modified in a number of ways to deal with this possible sub-normal output. Figure 4 illustrates the required changes for NM=3 and NE=3. Two additional qubits are needed. Qubit isub=0 is used as indication that result is a sub-normal number. Qubit icut=0 is similarly used to define cases with output truncated to 0. Both qubits are initialized to 1. Then, before the mantissa multiplication step takes place, a first modification is introduced, shown on the left-hand side of Figure 4. For NE=3, only inputs with exponent 000 will need truncating to 0, as shown in the first 4-qubit controlled-NOT gate flipping icut to 0. For NE=3, inputs with exponent 001 are guaranteed to lead to sub-normal output (or 0), and for these cases isub is set to 0, using the second 4-qubit controlled-NOT gate with isub as target. The mantissa-multiplication step shown in Figure 2 remains unchanged (i.e. qubits isub and icut are not used). The next required modification relates to the ‘copying’ of the result of the mantissa multiplication to output register imr1imr0 and the application of increments to the output exponent. The additional logic needed is shown on the right-hand side of Figure 3. First, for imp5=1, setting ier0=1 becomes conditional of both isub=1 and icut=1. The next two 4-qubit gates are used to guarantee that correct output with exponent 111 is created for inputs with exponents 101 and 110. The remaining gate operations perform the ‘copying’ of the mantissa squared into imr1imr0 taking into account the possible sub-normal output (cases with isub=0). The steps for isub=1 are the same as in the corresponding circuit for squaring without the sub-normal number modifications. A further set of circuit modifications to deal with sub-normal numbers is required in the quantum circuit used to obtain the output exponent. Figure 5 shows the additional operations required relative to the original quantum circuit shown in Figure 3. Three additional CNOT operations are introduced just before performing the QFT3. For isub=0 and icut=0 the initialization of imp1imp0 is modified so that the subsequent steps will produce the correct result for the exponent. The three CNOT operations also appear in the ‘uncompute’ stage at the right-hand side of the circuit. Further changes comprise two 4-qubit controlled-NOT operations on ier2 and ier1 required to create 111 exponents for inputs with exponent 110.

Figure 4.

Quantum circuits used in obtaining output mantissa for squaring operation, including sub-normal numbers and underflow/overflow protection (NM=3 and NE=3).

Figure 5.

Quantum circuit used to obtain exponent for squaring operation, including sub-normal numbers and under/overflow protection (NM=3 and NE=3).

For a fixed value of NE it is important to note that the additional complexity introduced by increasing NM is limited. In fact, the quantum circuit shown on the left-hand side of Figure 4 does not depend on NM. Similarly, the quantum circuits used to obtain the result exponent are independent of NM. The circuit shown on the right-hand side of Figure 4, representing the definition of imr1imr0 for cases with normalized or sub-normal output requires modification. Figure 6 shows how imr2imr1imr0 are set for NM=4 using a set of gate operations that has grown linearly with NM. The circuit shown accounts for sub-normal numbers and includes underflow/overflow protection.

Figure 6.

Quantum circuit used to set output mantissa for squaring operation, including sub-normal numbers and underflow/overflow protection (NM=4 and NE=3).

Advertisement

7. Quantum circuits for multiplication of floating-point numbers

In the interest of brevity, only the main features of the quantum circuits used for multiplication of two quantum floating-point numbers are summarized here. Figure 7 illustrates the quantum circuit used to compute the product of the mantissas of two inputs. Compared to the circuit shown in Figure 2 the main difference is that ancilla qubit a0 is now set using the mantissa of a second input. A further difference relative to the squaring operation occurs in the circuit used to obtain the result exponent. Here, instead of setting 2× the exponent using a bit shift, the sum of the two input exponents needs to be computed employing a quantum full adder.

Figure 7.

Quantum circuit used in multiplying the mantissa of two input numbers (NM=3 and NE=3).

Advertisement

8. Results of simulation and verification of quantum circuits

The proposed quantum circuits for squaring and multiplying floating-point numbers as part of the computational-basis representation, were systematically verified by gate-level simulation of the circuits for a wide range of cases with and without sub-normal numbers as well as cases with overflow results. The C++ quantum computer simulator detailed in previous work [4] was used for this purpose. To illustrate the process, the quantum algorithm used to square numbers with NM=3 and NE=3 is considered, with the following 19-qubit register (algorithm demonstrated accounts for sub-normal numbers as well as underflow/overflow protection, see Eq. (8) for reference):

ieb2ieb1ieb0imb1imb0a0imp5imp0ier2ier1ier0imr1imr0icutisubE9

where ieb2ieb1ieb0 and imb1imb0 define the exponent and the fractional part of the mantissa of the input, respectively. Qubits icut and isub are initialized as 1, while all other qubits are initialized as 0. The quantum state in the simulation is then initialized with a single non-zero (unit) amplitude, with the index in the quantum state vector defined by the binary representation of input exponent and fractional part of mantissa. With the rounding mode fixed at rounding down to nearest, the intended output can be easily computed before the quantum circuit is simulated. In effect, this defines the index of the single non-zero (unit) amplitude of the output quantum state that should be returned in case the circuit is correct. Upon finalizing the quantum computer simulation the actual quantum state vector obtained is compared against the previously-computed required output. For this verification to be meaningful, the following range of possible inputs and outputs were considered: (i) input and output are both normalized numbers, (ii) input is normalized number and output is a sub-normal number, (iii) input is a sub-normal number and result truncated to 0, (iv) input is a normalized number, with output overflow. For NM=3 and NE=3, Table 3 summarizes the input and output states for examples of each of the 4 categories considered. For inital and output the single non-zero amplitudes are shown. Since the simulator employed here stores the full 2nq state vector for nq qubits, only circuits with 28 qubits were considered as a result of limited computational resources and the large number of cases considered (>100). For the squaring operation, NM36 and NE34 were considered, while for multiplication the range of NM needed to be reduced, i.e. NM34.

InputInitial stateOutput state
7/2 (i)ψinit10011000000000000112=1ψout10011000000011010112=1
7/16 (ii)ψinit00111000000000000112=1ψout00111000000000011102=1
3/16 (iii)ψinit00011000000000000112=1ψout00011000000000000012=1
6 (iv)ψinit10110000000000000112=1ψout10110000000011100112=1

Table 3.

Results from quantum circuit simulation for representative range of inputs (squaring NM=3, NE=3).

Advertisement

9. Complexity analysis

Before analyzing the quantum circuits introduced here in terms of complexity, first the choice of NM and NE for representing realistic flow fields is considered.

9.1 Representing Taylor-green vortex flow

In a two-dimensional flow field, the non-linear terms appearing in the Navier–Stokes equations, shown in Eq. (4), involve the square of the velocity components in x and y directions, i.e. u2 and v2, as well as, the product uv. Here, the example flow field defined by the two-dimensional Taylor-Green vortex is considered, where velocity and pressure are defined in a square domain 02π2 with periodic boundary conditions as,

u=cosxsiny;v=sinxcosy;p=14cos2x+cos2yE10

Considering a 100×100 uniform mesh, the effect of representing the flow field variables with a reduced-precision floating-point format is analyzed first. Table 4 summarizes the results, highlighting the importance of including sub-normal numbers in the floating-point representation. Since a sign bit is not used here, the absolute values of u,v,p were actually used. Flow variables defined in Eq. (10) are in the range 11, so that by increasing NE from 3 to 4, far fewer sub-normal numbers are used to represent the flow field. As a result, removing the sub-normal number capability (as shown in bottom half of table), results in smaller errors for NE=4. For realistic applications of the proposed quantum floating point format, the relatively small overhead incurred by introducing sub-normal numbers in the quantum circuits clearly suggests that sub-normal numbers should be included.

NMNEL2uLuL2pLp
Rounding down - using sub-normal numbers
3326.8050.12474113.29080.0623342
437.699640.06249833.793960.0310842
531.930690.03124830.8830950.0154592
630.4778620.01562330.2335420.00780768
730.1103580.007810780.06117840.00390143
830.02476150.003904530.01355010.00194831
446.360020.06249831.575080.0310387
541.626790.03124830.3872610.0154137
640.4096630.01562330.108470.00762945
740.09589820.007810780.02960860.0037232
840.02098940.003904530.006478540.00192175
Rounding down - without sub-normal numbers
3386.86250.248583111.8960.249507
4370.44130.248583108.3520.249507
5365.82350.248583107.3590.249507
6364.63490.248583107.1350.249507
7364.32620.248583107.0690.249507
8364.25290.248583107.0500.249507
446.38810.06249831.61140.0310387
541.655030.03124830.424050.0154137
640.4379760.01562330.145340.0151248
740.1242230.01472180.06650740.0151248
840.04931630.01472180.04338270.0151248

Table 4.

Approximation errors in Taylor-green vortex flow field due to reduced-precision floating-point representation. L and L2 norms of errors relative to IEEE double-precision representation for velocity (u) and pressure (p) for different NM and NE. 100×100 uniform mesh.

For NE=4, the representation of u2 and uv is considered. Specifically, the error shown is that introduced by the multiplication: the difference between the ‘exact’ product of the reduced-precision representation of u and v and the corresponding reduced precision representation of the products is shown in Table 5. The results highlight that although sub-normal numbers played a relatively smaller role in representing velocity components, in the computation of the nonlinear terms, the inclusion of sub-normal numbers is more important for the minimization of approximation errors.

NMNEL2u2Lu2L2uvLuv
Rounding down - using sub-normal numbers
440.8015960.03515620.1619250.0146484
540.38480.02441410.05207720.00732422
640.1010350.0139160.0180160.00378418
740.03821580.007385250.00534490.00186157
840.01086210.003799440.001235370.000919342
Rounding down - without sub-normal numbers
440.872220.03515620.305110.0147705
540.4616890.02441410.2137560.0153809
640.180350.01514050.1883710.0154495
740.1192220.01514050.1796710.0154495
840.09271760.01526090.1775510.015553

Table 5.

Approximation errors of velocity products in Taylor-green vortex flow field due to reduced-precision floating-point representation. L and L2 norms of errors relative to IEEE double-precision representation for velocity (u2) and pressure (uv) for different NM and NE. 100×100 uniform mesh.

9.2 Mantissa multiplication step

QFT and inverse QFT are used involving 2NM qubits, so that the complexity in terms of two-qubit (controlled-phase) gates scales as NM2, where the well-known complexity of the standard QFT implementation is used. The complexity of the phase-addition steps involved in the multiplication are detailed in Table 6. For the two-qubit gates the number can be seen to scale as NM2, while the number of three-qubit gates shows a NM3 scaling.

CPHASEC2PHASEθmin
3×39272π/26
4×414662π/28
5×5201302π/210

Table 6.

Number of controlled-phase gates (CPHASE) and doubly-controlled-phase (C2PHASE) for phase-addition operator in quantum-multiplier. Also, smallest rotation angle is shown.

9.3 Computation of exponent

QFT and inverse QFT are used involving NE, NE+1 and NE+2 qubits, representing a smaller complexity than the QFT used in mantissa multiplications. The main contributions to complexity of exponent computation stems from the modulo and full-adders involving a number of qubits scaling linearly with NE. The polynomial complexity in terms of qubits for the adders implemented here is shown in Table 7.

CPHASEθminCPHASEθmin
MADD362π/23FADD392π/24
MADD4102π/24FADD4142π/25
MADD5152π/25FADD5202π/26
MADD6212π/26

Table 7.

Number of controlled-phase gates (CPHASE) in phase-addition step for modulo adder (MADD) and full adder (FADD). Also, smallest rotation angle is shown.

9.4 Discussion

The quantum circuits presented here for squaring two floating-point numbers in the format proposed show that by accounting for sub-normal numbers and under/overflow an additional number of multi-qubit controlled-NOT gates is needed. However, for the examples analyzed a polynomial dependence on NM and NE was observed. This means that in terms of quantum-algorithm complexity this implementation has the desired efficiency. The relatively small complexity as compared to circuits used for mantissa multiplication highlights that for most applications it is desirable to include the capability of using sub-normal numbers and provide under/overflow protection in the quantum circuits. The analysis in this section also shows that for a realistic application, a well-considered scaling of the governing equations to O1 variables is even more important here than in classical implementations using IEEE single- or double-precision arithmetic. Using the limited number of qubits available on current and near-term quantum computers (<100), the proposed approach to introducing non-linearity is a good candidate in cases where NM and NE can be chosen significantly smaller than in equivalent classical floating-point representations.

Advertisement

10. Conclusions

The challenges associated with representing non-linear differential equations in terms of quantum circuits were discussed in this chapter. In this work, a new approach for representing product-terms in nonlinear equations suitable for near-term (e.g. NISQ generation) quantum computers was proposed. A key aspect discussed is the (temporary) representation of the variables in the computational basis. Furthermore, the use of a suitably-chosen floating-point format was detailed. The importance of including sub-normal numbers, such as defined in the IEEE 758 standard for floating-point arithmetic on classical computers, was demonstrated. Based on the current findings, a number of suggestions for further work can be put forward. The presented circuits performed arithmetic for a single set of input data, i.e. equivalent to data for a single point in a computational domain. Extending the approach to a multi-dimensional computational mesh is a first step to consider. A complexity analysis will be needed to assess the potential speed-up relative to classical discretization approaches for the considered equations. A further step involves investigating how the proposed approach can be made part of a larger quantum algorithm, where a mix of amplitude-based encoding and computational-basis encoding occurs. A key aspect is therefore the development of efficient quantum circuits to perform the required conversions between the two different encoding approaches. Finally, further work is needed to establish how the approach presented here can be used in a wider range of quantum computing applications.

References

  1. 1. Nielsen MA, Chuang IL. Quantum Computation and Quantum Information: 10th Anniversary Edition. 2nd ed. Cambridge: Cambridge University Press; 2010
  2. 2. Yepez J. Quantum lattice-gas model for computational fluid dynamics. Phys. Rev. E. 2001;63(4):046702. DOI:10.1103/PhysRevE.63.046702
  3. 3. Berman GP, Ezhov AA, Kamenov DI, Yepez J. Simulation of the diffusion equation on a type-II quantum computer. Phys. Rev. A. 2002;66(1):012310. DOI:10.1103/PhysRevA.66.012310
  4. 4. Steijl R, Barakos GN. Parallel evaluation of quantum algorithms for computational fluid dynamics. Computers&Fluids 2018;173:22-28. DOI: 10.1016/j.compfluid.2018.03.080
  5. 5. Steijl R. Quantum Algorithms for Fluid Simulations. In: Bulnes F, Stavrou VN, Morozov O, Bourdine AV, editors. Advances in Quantum Communication and Information. IntechOpen; 2020. DOI: 10.5772/intechopen.86685
  6. 6. Todorova BN, Steijl R. Quantum Algorithm for the collisionless Boltzmann equation. J. Comp. Phys. 2020;409:109347. DOI:10.1016/j.jcp.2020.109347
  7. 7. Griffin KP, Jain SS, Flint TJ, Chan WHR. Investigations of quantum algorithms for direct numerical simulation of the Navier-Stokes equations. Center for Turbulence Research Annual Research Briefs. 2019;347-363
  8. 8. Gaitan F. Finding flows of a Navier–Stokes fluid through quantum computing. npj Quantum Information. 2020; 6:61. DOI:10.1038/s41534-020-00291-0
  9. 9. Harrow AW, Hassidim A, Lloyd S. Quantum algorithm for linear systems of equations. Phys. Rev. Lett. 2009;103(15):150502. DOI:10.1103/PhysRevLett.103.150502
  10. 10. Clader BD, Jacobs BC, Sprouse CR, Preconditioned quantum linear system algorithm. Phys. Rev. Lett. 2013;110(25):25054. DOI:10.1103/PhysRevLett.110.250504
  11. 11. Cao Y, Papageorgiou A, Petras I, Traub J, Kais S. Quantum algorithm and circuit design solving the Poisson equation. New J. Phys. 2013;15:013021. DOI:10.1088/1367-2630/15/1/013021
  12. 12. Scherer A, Valiron B, Mau S-C, Alexander S, van den Berg E, Chapuran TE. Concrete resource analysis of the quantum linear-system algorithm used to compute the electromagnetic scattering cross section of a 2D target. Quantum Inf. Proc. 2017;16(3):60. DOI:10.1007/s11128-016-1495-5
  13. 13. Montanaro A, Pallister S. Quantum Algorithms and the finite element method. Phys. Rev. A. 2016;93(3):032324. DOI:10.1103/PhysRevA.93.032324
  14. 14. Xu G, Daley AJ, Givi P, Somma RD. Turbulent mixing simulation via a quantum algorithm. AIAA J. 2018;56(2):687-699. DOI:10.2514/1.J055896
  15. 15. Berry DW. High-order quantum algorithm for solving linear differential equations. J. Phys. A. 2014;47(10):105301. DOI:10.1088/1751-8113/47/10/105301
  16. 16. Berry DW, Childs AM, Ostrander A, Wang G. Quantum Algorithm for Linear Differential Equations with Exponentially Improved Dependence on Precision. Comm. Math. Phys. 2017;356(3):1057-1081. DOI:10.1007/s00220-017-3002-y
  17. 17. Fillion-Gourdeau F, Lorin E. Simple digital quantum algorithm for symmetric first-order linear hyperbolic systems. Numerical Algorithms. 2019;82:1009-1045. DOI:10.1007/s11075-018-0639-3
  18. 18. Costa PCS, Jordan S, Ostrander A. quantum algorithm for simulating the wave equation. Phys. Rev. A. 2019;99(1):012323. DOI:10.1103/PhysRevA.99.012323
  19. 19. Childs AM, Liu J-P. Quantum spectral methods for differential equations. Comm. Math. Phys. 2020; 375(2):1427-1457. DOI:10.1007/s00220-020-03699-z
  20. 20. Leyton SK, Osborne TJ. A quantum algorithm to solve nonlinear differential equations. arXiv.org 2008;0812.4423
  21. 21. Zhou SS, Loke T, Izaac JA, Wang JB. Quantum Fourier transform in computational basis. Quantum Inf. Proc. 2017; 16(3):82. DOI:10.1007/s11128-017-1515-0
  22. 22. Peruzzo A, McClean J, Shadbolt P, Yung M-H, Zhou X-Q, Love PJ, Aspuru-Guzil A, O’Brien JL. A variational eigenvalue solver on a photonic quantum processor. Nature Comms 2014; 5:4213. DOI:10.1038/ncomms5213
  23. 23. McClean JR, Romero J, Babbush R, Aspuru-Guzik A. The theory of variational hybrid quantum-classical algorithms. New J. Phys. 2016; 18:023023. DOI:10.1088/1367-2630/18/2/023023
  24. 24. Lubasch M, Joo J, Moinier P, Kiffner M, Jaksch D. Variational quantum algorithms for nonlinear problems. Phys. Rev. A. 2020;101(1):010301. DOI:10.1103/PhysRevA.101.010301
  25. 25. Mitarai K, Kitagawa M, Fijii K. Quantum analog-digital conversion. Phys. Rev. A. 2019;99(1):012301. DOI:10.1103/PhysRevA.99.012301
  26. 26. SaiToh A. quantum digital-to-analog conversion algorithm using decoherence. Quantum Inf. Proc. 2015; 14(8):2729-2748. DOI:10.1007/s11128-015-1033-x
  27. 27. Haener T, Soeken M, Roetteler M, Svore KM. Quantum circuits for floating-point arithmetic. In: Kari J, Ulidowski I, editors. Reversible Computation. RC 2018. Lecture Notes in Computer Science, vol 11106. Springer; 2018. DOI:doi.org/10.1007/978-3-319-99498-7-11
  28. 28. Grover L, Rudolph T. Creating superpositions that correspond to efficiently integrable probability distribution. arXiv 2002;0208112
  29. 29. Ma G, Li H, Zhao J. Quantum QR decomposition in the computational basis. Quantum Inf. Proc. 2019; 19:271. DOI:10.1007/s11128-020-2777-4
  30. 30. Bhaskar MK, Hadfield S, Papageorgiou A, Petras I. Quantum algorithms and circuits for scientific computing. Quantum Info. Comput. 2016;16(3-4):197-236. DOI:10.5555/3179448.3179450
  31. 31. Overton M.L. Numerical Computing with IEEE Floating Point Arithmetic. 1st ed. Philadelphia: SIAM; 2001. 97p

Written By

Rene Steijl

Reviewed: 13 November 2020 Published: 10 December 2020