Geometric Interpretation of Sensitivity to Structured Uncertainties in Spintronic Networks
Abstract
We present a geometric model of the differential sensitivity of the fidelity error for state transfer in a spintronic network based on the relationship between a set of matrix operators. We show an explicit dependence of the differential sensitivity on the fidelity (error), and we further demonstrate that this dependence does not require a trade-off between the fidelity and sensitivity. Rather, we prove that for closed systems, ideal performance in the sense of perfect state transfer is both necessary and sufficient for optimal robustness in terms of vanishing sensitivity. We demonstrate the utility of this geometric interpretation of the sensitivity by applying the model to explain the sensitivity versus fidelity error data in two examples.
I Introduction
Exploiting the unique properties of quantum systems has the potential to revolutionize applications in communications, sensing, and computing [1, 2]. However, harnessing any technological “quantum advantage” requires external controls that not only meet exacting performance criteria, but retain a high level of performance in the face of model uncertainty. While classical synthesis methods with guaranteed performance margins are applicable to a subset of open quantum systems undergoing weak continuous measurement [3, 4], such methods are ill-suited to the marginally stable dynamics of closed quantum systems such as idealized spintronic networks. As such, post-synthesis selection of acceptable controllers based on an assessment of the robustness or the sensitivity of the performance measure is common. Various methods for delivering such assessments can be found in the literature, ranging from stochastic [5, 6] to analytic [7, 8, 9].
Previous work has applied robustness assessments across a range of controllers to determine whether a trade-off between performance and robustness is a fundamental limitation in quantum control problems. Such a trade-off might be expected from the fundamental limitation of classical feedback control, encapsulated by the frequency domain identity , where and represent the error and log-sensitivity, resp. In [9], we showed that there is a trade-off between performance and robustness in the time domain in the sense that the normalized logarithmic sensitivity of the fidelity error diverges as the error approaches zero. However, both [8] and [10] suggest that such a trade-off does not necessarily hold for suboptimal controllers for closed systems, in that the best-performing controllers (lowest fidelity error) are not necessarily those with the worst robustness (magnitude of the log-sensitivity). Furthermore, [11] provided examples of state transfer in spintronic networks where controllers with the best fidelity exhibited the smallest unnormalized differential sensitivity to parameter variations. Similarly, the results of [12] indicated a high concordance between the fidelity error and differential sensitivity for gate control problems, indicating that the best performing controllers are also the most robust.
The purpose of this paper is to investigate why such differing trends between error and sensitivity exist for state transfer problems. We develop a geometric interpretation of the differential sensitivity of the fidelity error to parametric uncertainty and use it to derive an explicit relationship between the magnitude of the sensitivity and the fidelity of transfer. The established relationship shows that there need not be a trade-off between the sensitivity and fidelity error and provides insight into the most important factors that determine the magnitude of the sensitivity. As part of the exposition, we provide necessary and sufficient conditions for vanishing sensitivity. Finally, we analyze two case studies to demonstrate the utility of the derived sensitivity model.
II Preliminaries
II-A Physical Model and Control Problem
As in [10], we consider a coupled network of spin- particles with real Hamiltonian
(1) |
is the coupling strength between spin and and is the respective Pauli operator acting on spin , formally the -fold tensor product of copies of the identity matrix with a Pauli matrix in the th position. We consider the case of uniform coupling with for all pairs . We restrict the dynamics to the single excitation subspace (SES), the subspace of the Hilbert space isomorphic to corresponding to a single excited spin in the network. This is justified when the goal is to transfer the state of the system from that of a single excited ”input” spin to a single excited ”output” spin with all other spins in the ground state, tantamount to transfer of a single qubit of information. While control of quantum systems in an open-loop manner via optimally shaped time-varying fields is a mainstay in quantum control [1], we consider the alternative paradigm of shaping the energy landscape of a closed system to facilitate the desired evolution [13, 14]. In the interest of model simplicity, we introduce time-invariant external fields to maximize the probability of state transfer from a given pure input state to a pure output state . As detailed in [15], these control fields enter the Hamiltonian as scalars on the diagonal of the SES Hamiltonian. We also consider only XX coupling since in the SES any coupling terms generated by are diagonal and can be absorbed into the controls. In terms of the Schrodinger dynamics, we have the initial value problem
(2) |
where is the controlled Hamiltonian in the SES and units are chosen such that . The are generated by maximizing the probability of transfer, or fidelity, at a read-out time given by . See [10, 15] for details on the controller synthesis problem.
II-B Uncertainty Model
To analyze the robustness of the controlled system, we use an uncertainty model that enables evaluation of how the introduction of uncertainty or disturbances alters the performance metric. In line with previous work [8, 9, 16] we consider real parametric uncertainty to the Hamiltonian represented as
(3) |
is the nominal controlled Hamiltonian, and represents a small, real perturbation strength. is the structure associated with the uncertainty indexed by . is a factor that accounts for scaling by the control field strength for those that index uncertainty in a control channel; otherwise (e.g., for uncertainty in the couplings) .
II-C Bloch Formulation
To put the analysis in the framework of a real vector space, we use the Bloch formulation to describe the perturbed dynamics [17]. Consider the most general description of the Schrödinger dynamics where the density operator for the pure state is given by . The perturbed dynamics generated by (3) are given by the von-Neumann equation
(4) |
Taking the adjoint representation of this equation with respect to an orthonormal basis of the Hermitian matrices, such as the generalized Gell-Mann basis [18], yields the mapping with elements
(5) |
where is the matrix commutator [17]. By linearity of the commutator and trace, and so that . As and are skew-Hermitian, it follows from the expansion (5) that and are skew-symmetric. The state maps to with elements so that , , and the state equation is
(6) |
with solution . The fidelity at the read-out time is . The fidelity error is . Finally, we define as the state transition matrix at and nominal and define its perturbed counterpart as .
II-D Differential Sensitivity
With this framework, the un-normalized differential sensitivity of to a perturbation structured as is given by [9, 19]. Given the spectral decomposition of (at its nominal value), ,
(7) |
where , , , and denotes the Hadamard product. Denoting as the th diagonal element of , the entries of are
(8) |
Alternatively, we may define the operator such that and
(9) |
III Geometric Interpretation of the Sensitivity
III-A Derivation of the Model
We now establish an analytic dependence of , the differential sensitivity to an uncertainty indexed by , on the fidelity error , or equivalently, the fidelity . We take a geometric approach and consider the operators and as elements of , the space of bounded operators on [20]. With the Hilbert-Schmidt inner product for , with compatible definition of the norm . The orthognal projection of in the direction of is given by . Setting , we have .
Let be the two-dimensional subspace of spanned by and .
Lemma 1
The operators and are orthogonal.
Proof:
By definition of the inner product and the differential sensitivity, we have
(10) | ||||
(11) |
Now (5) implies , being real and skew-symmetric, so for any allowed structure that retains unitary dynamics. ∎
Define the operator as . With this and , and the projection of on the subspace can be written as
(12) |
The projection given in (12) is orthogonal as . Based on the projection definition, this is equivalent to
(13) |
where follows from . Defining as the angle subtending and and as the angle subtending and gives
(14) |
Figure 1 displays the relation of the operators in the subspace . It follows that and . By Lemma 1, we have that and are orthogonal so that either as in Case or as illustrated in Case , and . Thus we have following relation between the magnitude of the differential sensitivity and the fidelity
(15a) | ||||
(15b) |
Eqs. (12)-(13) appear to be the quantum equivalent of the classical relation , expressed in terms of the fidelity rather than the error . At the less conceptual level of (15b), holding and constant, and decrease simultaneously, in disagreement with the classical limitations mandating a trade-off between error and sensitivity. What undermines the simplistic formulation that and are concordant is that and are not independent quantities; however, the first and crucial quantity can be bounded.
III-B The Relation between and Eigenstructure
To determine how depends on the eigenstructure of the controller , we employ (7), . It follows immediately that , and with (8)
(16) |
so that
(17) |
with the definition . We now identify the maximum of and establish strict positivity.
Lemma 2
is non-zero and bounded above by .
Proof:
To see that is not zero it suffices to show that is never the zero matrix. Consider the integral expression of from (9). Pulling out of and and using , we are left with . Noting that and , shows that the integrand is the adjoint action of the special orthogonal group on its Lie algebra for any [21]. This Lie group conjugation is described by the mapping . If the kernel of the integral of is empty, it follows that is not the zero matrix for non-trivial . The spectrum of given by determines the eigenvalues of , which are for indexed from to . The eigenvalues of retain the dependence of the integration variable (equivalently ), as can be verified by an explicit spectral decomposition of . Integrating over the interval yields eigenvalues for of for and otherwise. As these eigenvalues are never zero, the kernel of the integral of is empty. It follows that, for a non-trivial uncertainty structure , is always non-zero. To establish the upper bound, note that when all terms in (17) assume the maximum of , then . The upper bound of is only achieved if for all pairs , in which case, , which would provide a fidelity of zero for orthogonal states. ∎
Replacing by its upper bound in (15a) yields an upper bound on , which decreases as decreases. But this behavior on the bound does not rule out an increase in as decreases [9]. This indicates that any quantum performance limitation is not as straightforward as in classical control and that and could be concordant [8, 10] or discordant [9].
III-C Sufficient Condition for Vanishing Sensitivity
Lemma 3
For unitary evolution, the differential sensitivity vanishes for any physically allowable perturbation if the controller induces perfect state transfer.
Proof:
As a physically realizable perturbation structure, is Hermitian, and its Bloch representation is skew-symmetric. For non-zero and , is given by
From Lemma 2, the term in the integral is the adjoint action of on for any value of . Thus , and is secured if the product is symmetric. We thus require . Symmetry of this dyadic product requires for some real . Since and is orthogonal, it follows that . For , , which is the condition for perfect state transfer from input state to . Thus, for , a controller inducing perfect state transfer is sufficient for . For , we would have . Observing that with , it follows that , and would be associated with a nonpositive definite density, which is absurd. ∎
III-D Properties of
The following observations demonstrate bounds on .
Observation 1
Observation 2
For any fidelity, is bounded below by . This follows directly from (15b), which requires for the sensitivity to be real.
Though Observation 1 only provides the upper bound on in the exceptional case of perfect state transfer, we find that this holds as an upper bound in general. Fig. 2 for the case of a -ring and coupling uncertainty shows that remains below for all controllers.
In keeping with the geometric picture, we have where measures the minimum angle between and any other elements of . We thus have and deduce that for a given uncertainty structure the magnitude of the sensitivity is given as .
IV Vanishing Differential Sensitivity
We now extend the result of [15, Th. 3] for sufficient conditions on vanishing sensitivity in spintronic networks. We use our geometric model to show that perfect state transfer is not only sufficient but necessary for vanishing sensitivity.
Theorem 1
For unitary evolution and controllers that yield non-zero fidelity, the differential sensitivity vanishes for any physically allowable perturbation or uncertainty structure, if and only if the controller induces perfect state transfer.
Proof:
Sufficiency is established by Lemma 3. For necessity, we refer to (15a). For any physically realizable controller, both and are non-zero. A non-trivial sensitivity thus requires non-zero , , and . From Lemma 2 we have the is always non-zero. From (13) it follows that the component of in the direction is only zero if the fidelity is zero. So unless the fidelity vanishes. Then requires . But this is precisely the configuration of the operators in for perfect state transfer. We thus conclude that the differential sensitivity vanishes if and only if the controller yields perfect state transfer. ∎
V Examples
To demonstrate the utility of this model, we examine the fidelity error and sensitivity relationship for two state-transfer examples from taken from the data set in [10].
V-A Differing Fidelity Error-Sensitivity Profiles
We begin with the case of state-transfer from spin to spin in a -ring with control mediated by static, time-invariant controls. We index the possible perturbations to the controls (i.e., perturbations to the diagonal elements of the SES Hamiltonian) by though , and to the uncertainty in the couplings (i.e. uncertainty in the entries and of the SES Hamiltonian) by though . Uncertainty to the entries and is indexed by . For all cases aside from and , and display a strong linear correlation with a Pearson greater than . For the remaining two cases the correlation coefficient is less than . To investigate the origin of this differing behavior, we examine the differences between the and uncertainty cases as depicted in Figure 3. The strong linear relationship for the case is borne-out by a Pearson of , while the much flatter trend for the case can be verified by the Pearson of .
We explain these different trends with the geometric model of Section III and (15a). Consider the visualization of the components of shown in Figure 4. Note that both cases correspond to coupling uncertainty, so . For both cases, is nearly constant at across all controllers with a maximum deviation below this value on the order of . This data is not displayed in Fig. 4. shows some deviation across the controllers, however these deviations are unlikely to induce the different trends observed in Figure 3. In particular for , has a mean of and variance of , while for the statistics for are and . However, as seen from Figure 4, the behavior of is markedly different for the controllers in the versus cases. The positive rank correlation between and is borne out by a Kendall of . Conversely, the lack of rank correlation between metrics for the case is evident in the weak Kendall of .
This difference in behavior for has the geometric interpretation that for the structure, the change in the fidelity (error) is more attributable to the increase in the angle defining the inclination of with the subspace than to an increase of within the subspace. This minimal change of with the fidelity error manifests as the absence of a clear trend between and as the fidelity error increases. Conversely, for the case, an increase in the fidelity error is strongly correlated with, and nearly proportional to, an increase of the angle within the subspace . This suggests that those uncertainty structures that generate a subspace spanned by and such that a change in the fidelity corresponds to an change in , are more sensitive (less robust) to variations in the fidelity error. To account for the read-out time note that this the same for both uncertainty cases and thus may be ruled out as the cause of the differing trends observed in Figure 3.
V-B Large Variation in Sensitivity for Nearly Equal Error
We now study a case where there is no trend between and across controllers for the same uncertain parameter. We consider state transfer from spin to spin in a -ring with perturbation structure (perturbation to the control addressing spin ).
Figure 5 depicts the sensitivity versus error profile, which does not reveal any visual trend between the two metrics. Rather, we see that the controllers with the lowest absolute value of sensitivity fall in the range of . We focus on the controller indices , , and (the red circled data points in Figure 5) which have an error in the range , while the differential sensitivity for these controllers spans orders of magnitude from for controller up to for controller . We provide insight into this vast change in sensitivity by considering the geometric factors encoded in the size and orientation of the related matrix operators along with the physical parameters and as captured in (15a). Table I provides a summary of the relevant factors contributing to for each controller. Referring to this data, we see that unlike in the previous example, the behavoir cannot be attributed simply to the effect of , but rather to a combination of the factors , , and . The extremely small sensitivity of controller is most attributable to a small angle between and and small control amplitude. While controller demonstrates a only an order of magnitude larger than the previous controller and a shorter read-out time, the large value of the control field contributes to an increase of over two orders of magnitude in . Finally, while controller admits a smaller control amplitude than its predecessor, the larger read-out time and , manifest as the largest sensitivity of the trio. This brief analysis justifies the premium placed on minimizing transfer times to increase robustness beyond the desire to minimize the impact of decoherence in open systems in general [22]. It also supports efforts to limit control amplitudes beyond energy considerations. Finally, while optimizing directly for small may not be possible, minimizing the fidelity error necessarily minimizes with a concomitant reduction in sensitivity.
VI Conclusion
We developed a geometric model of the fidelity versus sensitivity to parametric uncertainty embodied in Eq. (12) more descriptive than the analytical formula (15b). We employed this model to provide insight into results of previous work based on statistical analysis of the sensitivity versus fidelity error [8]. With this geometric model we expanded the scope of another previous work [15, Th. 3] by proving that perfect fidelity is not only sufficient, but necessary, for vanishing sensitivity. Future work should focus on relating and to the eigenstructure of the controller in order to inform synthesis methods that account for these factors’ impact on sensitivity.
References
- [1] Christiane P. Koch et al., “Quantum optimal control in quantum technologies. Strategic report on current status, visions and goals for research in Europe,” EPJ Quantum Technology, vol. 9, 7 2022.
- [2] Steffen J. Glaser et al., “Training Schrödinger’s cat: quantum optimal control,” EJP D, vol. 69, 12 2015.
- [3] S. Wang, C. Ding, Q. Fang, and Y. Wang, “Quantum robust optimal control for linear complex quantum systems with uncertainties,” IEEE Trans. Autom. Control, vol. 68, no. 11, pp. 6967–6974, 2023.
- [4] M. R. James, H. I. Nurdin, and I. R. Petersen, “ control of linear quantum stochastic systems,” IEEE Trans. Autom. Control, vol. 53, no. 8, pp. 1787–1803, 2008.
- [5] I. Khalid, C. A. Weidner, E. A. Jonckheere, S. G. Shermer, and F. C. Langbein, “Statistically characterizing robustness and fidelity of quantum controls and quantum control algorithms,” Phys. Rev. A, vol. 107, Mar. 2023.
- [6] A. Koswara, V. Bhutoria, and R. Chakrabarti, “Robust control of quantum dynamics under input and parameter uncertainty,” Phys. Rev. A, vol. 104, p. 053118, Nov 2021.
- [7] R. L. Kosut, M. D. Grace, and C. Brif, “Robust control of quantum gates via sequential convex programming,” Phys. Rev. A, vol. 88, nov 2013.
- [8] E. Jonckheere, S. Schirmer, and F. Langbein, “Jonckheere‐Terpstra test for nonclassical error versus log‐sensitivity relationship of quantum spin network controllers,” Int. J. Robust Nonlinear Control., vol. 28, p. 2383–2403, Jan. 2018.
- [9] S. O’Neil, S. Schirmer, F. C. Langbein, C. A. Weidner, and E. A. Jonckheere, “Time-domain sensitivity of the tracking error,” IEEE Trans. Autom. Control, vol. 69, no. 4, pp. 2340–2351, 2024.
- [10] S. P. O’Neil, F. C. Langbein, E. Jonckheere, and S. Shermer, “Robustness of energy landscape controllers for spin rings under coherent excitation transport,” Research Directions: Quantum Technologies, vol. 1, p. e12, 2023.
- [11] E. A. Jonckheere, S. G. Schirmer, and F. C. Langbein, “Structured singular value analysis for spintronics network information transfer control,” IEEE Trans. Autom. Control, vol. 62, no. 12, pp. 6568–6574, 2017.
- [12] S. O’Neil, C. Weidner, E. Jonckheere, F. Langbein, and S. Schirmer, “Robustness of dynamic quantum control: Differential sensitivity bounds,” AVS Quantum Science, vol. 6, no. 3, 2024.
- [13] A. Donovan, V. Beltrani, and H. A. Rabitz, “Quantum control by means of hamiltonian structure manipulation.,” Phys. Chem. Chem. Phys., vol. 13 16, pp. 7348–62, 2011.
- [14] R.-W. Zhang, C. Cui, R. Li, J. Duan, L. Li, Z.-M. Yu, and Y. Yao, “Predictable gate-field control of spin in altermagnets with spin-layer coupling,” Phys. Rev. Lett., vol. 133, p. 056401, Aug 2024.
- [15] S. G. Schirmer, E. A. Jonckheere, and F. C. Langbein, “Design of feedback control laws for information transfer in spintronics networks,” IEEE Trans. Autom. Control, vol. 63, no. 8, pp. 2523–2536, 2018.
- [16] S. P. O’Neil, I. Khalid, A. A. Rompokos, C. A. Weidner, F. C. Langbein, S. Schirmer, and E. A. Jonckheere, “Analyzing and unifying robustness measures for excitation transfer control in spin networks,” IEEE Control Syst. Lett., vol. 7, pp. 1783–1788, 2023.
- [17] F. F. Floether, P. de Fouquieres, and S. G. Schirmer, “Robust quantum gates for open systems via optimal control: Markovian versus non-Markovian dynamics,” New J. Phys., vol. 14, p. 073023, jul 2012.
- [18] R. A. Bertlmann and P. Krammer, “Bloch vectors for qudits,” J. Phys. A-Math., vol. 41, p. 235303, may 2008.
- [19] I. Najfeld and T. Havel, “Derivatives of the matrix exponential and their computation,” Adv. Appl. Math., vol. 16, no. 3, pp. 321–375, 1995.
- [20] J. Siewert, “On orthogonal bases in the Hilbert-Schmidt space of matrices,” J. Phys. Commun., vol. 6, p. 055014, May 2022.
- [21] D. Elliott, Bilinear Control Systems: Matrices in Action. Springer Publishing Company, Inc., 1st ed., 2009.
- [22] C. P. Koch, “Controlling open quantum systems: tools, achievements, and limitations,” J. Condens. Matter Phys., vol. 28, p. 213001, 2016.