3.1. Entropy for Bell States, Product States, and Mixed States
The numerical calculations allow us to calculate the most fundamental property of the system, defined by the von Neumann entropy
. The von Neumann entropy (also defined as quantum entropy) quantifies the uncertainty of a quantum state represented by the density operator
, defined as:
Following
Figure 1, both Bell states
and
initially display an entropy of approximately zero, indicating that the system starts in a pure state. Over time, the entropy increases to about 1.85 bits, suggesting a process of mixing, likely due to decoherence. Notably, the identical evolution of these states implies that the dynamics are phase-insensitive, meaning that the system’s evolution does not depend on the specific phase relationship between the states. For the product states
and
, the initial entropy of the full system is also zero, but the reduced entropies are approximately 0.72 bits and 0.97 bits, respectively. As the system evolves, the entropy increases, reaching values of 1.3 bits and 0.8 bits. This growth in entropy suggests that entanglement is generated over time, transforming the initially separable states into partially entangled states. The differences in the entropy values indicate that the two product states undergo different degrees of entanglement formation, possibly due to variations in their initial superposition coefficients.
For the mixed states and , the initial reduced entropies are 1 bit and approximately 0.65 bits. The entropy of fluctuates between 0.75 and 1.75 bits before stabilizing at 1.6 bits, while reaches a peak of 1.9 bits and settles there. This suggests more complex dynamical behavior compared with the Bell and product states, possibly due to the probabilistic mixture of pure states in their initial configurations. The long-term stabilization of entropy values indicates that the system reaches a steady state, reflecting a balance between entanglement generation and dissipation effects. Overall, all simulations generate steady states with a maximal value of , it is, however, noteworthy that the interval represents a transition phase with decoherence and coherence, which generate one-hump camel-like entropy behavior, particularly of the product and mixed states. The Bell states display a lesser undulation in this interval, which, however, is more pronounced in the entanglement entropy seen in the next section.
Regarding the entropy plots (see
Figure 1), we observe a striking similarity to the spin-up electron fraction measured for a single phosphorus donor atom in silicon subjected to microwave radiation, as reported by Pla et al. [
20], however, with a considerable damping effect. Specifically, their results show Rabi oscillations of the spin-up fraction, which we find mirrored in the non-monotonic time evolution of the von Neumann entropy in our system for both product states (
,
) and mixed states (
,
). For the product states in
Figure 1, particularly
and
, we observe the entropy behavior as “camel-like” due to the characteristic rise, dip, and subsequent increase, resembling a double-humped structure in the entropy curves. Indeed, for the mixed states in
Figure 1, both
and
exhibit a rise-dip pattern, with
(blue curve) showing a more pronounced “camel-like” shape, 1.75 bits, while
(red curve) reaches a higher entropy of 1.85 (as the Bell states) bits with a subtler dip. This non-monotonic pattern in both cases can be attributed to coherent dynamics driven by noncommuting terms in the Hamiltonian, which induce damped Rabi oscillations [
9,
19]. These oscillations cause the qubit’s state to evolve coherently, leading to a transient oscillatory profile in the entropy when traced over a subsystem; however, weak dissipation, as introduced by the dissipation term
, damps these oscillations after a single cycle, resulting in the observed “camel-like” shape before the entropy stabilizes at a steady value.
We add here an additional analysis of the evolution of the entropy for the Bell states using a parameterized coefficient of
. We parameterize
on the domain
. By this, we obtain a map of the effect of the relaxation factor
on the unit interval, providing a smooth variation for the entropy evolution analysis, which is shown in
Figure 2.
We find an interesting result in the plot of the parameterized under the framework given in [
2] with relaxation factor
(
Figure 2), where the maximal number of undulations in the entropy (between 1 and 2 bits) occurs for
values approximately between
radians and
radians. The special behavior of the entropy observed in
Figure 2 is aligned with the damped Rabi oscillations detected for spin-fractions of single silicon atoms studied by Pla et al. [
20]. In this study, several different spin oscillations were observed after various input powers were applied to a single silicon atom device. The plots of these different spin oscillations result as highly similar plots to the various plots represented in
Figure 2, indicating that the Khrennikov setup of the GKSL equation may be suitable for the calculation of properties of such spin-fractions from single-atom systems and entropy properties therein.
The Rabi oscillations we observe in
Figure 2 signify intermittent entanglement degradation and revival, arising from the competition between the Hamiltonian’s coherent dynamics and the dissipative action of the
operator. Furthemore, the undulations in the entropy evolution for
under
are signatures of non-Markovian dynamics [
27], violating the Markovian monotonicity condition
where these oscillations lead to an increasing number of metastable entanglement revival via Liouvillian exceptional points [
28]. These points are degeneracies in quantum open systems (and classical systems as well) and have significant relevance to physics and optics [
28]. Such exceptional points define an entropy evolution where two or more eigenvalues coalesce [
29] and thus revoke that
-type Bell states could be optimally protected against external influence (i.e., decoherence channels) when
= 0.1–0.3, making them promising for quantum devices to reduce noise and error.
3.2. Entanglement Entropy
The entanglement entropy is defined by the von Neumann entropy of the reduced density matrix and is employed by being computed for one of the states, working as a key metric for understanding the coherence and mixedness of quantum states over time. Following the top plot in
Figure 3, we focus on the Bell states
and
, both initially maximally entangled with
(due to normalization, scaling to a maximum value of 1 for clarity). As time progresses, the entropy (represented by a single red curve, as both states evolve identically) dips slightly around
, reflecting the dissipative influence of the raising operator, which nudges the system toward the ground state
. However, the entropy stabilizes near 1, significantly above zero, indicating robust residual entanglement. This persistence suggests that the Hamiltonian
H, by facilitating transitions between
and
, effectively counteracts the disentangling effects of dissipation, preserving quantum correlations in the long-time limit.
For the product states
and
, the entanglement entropy dynamics reveal three distinct phases. Initially, both states show zero entanglement entropy (as expected for separable states), followed by a sharp increase to approximately 0.8–0.9 bits, indicating significant entanglement generation under the system’s dynamics. The subsequent plateau around 1 bits demonstrates stable entanglement preservation, with the states remaining highly (though not maximally) entangled. The close similarity between both curves suggests the dynamics affect different product states in qualitatively similar ways, while the slight difference in their steady-state values may reflect varying degrees of correlation in their initial configurations. The mixed states
and
(
Figure 3) exhibit more complex behavior, following a damped Rabi oscillation mode, beginning with zero entanglement despite their mixed nature, which highlights their preparation using mixtures of product states. Both states rapidly develop entanglement, with both reaching maximal entanglement (1 bit) and also dipping like the Bell states in the interval
. This non-monotonic evolution suggests competing dynamics with initial entanglement generation through unitary evolution followed by partial disentanglement, due to decoherence or specific Hamiltonian terms. We pay particular attention to the interval
, which aligns with the definition of one-hump in Definition 1, and study it further by the rate of entanglement entropy.
Relevance with Established Literature
Our findings can be placed in the broader context of quantum many-body out-of-equilibrium physics, where entanglement entropy is routinely employed as a diagnostic of dynamical phases and information spreading. In particular, global and local quenches in one-dimensional lattice and conformal field theory models produce an initially linear growth of entanglement followed by saturation in finite systems [
30,
31]. The non-monotonic (“camel-like”) profiles and revivals we observe under GKSL dynamics have analogues in finite spin chains and quenched CFTs, where coherent quasiparticle propagation and finite-size reflections generate oscillations of the entropy [
32]. Furthermore, the entropy oscillations and transient entanglement revivals reported here are consistent with signatures of non-Markovian dynamics and information backflow identified in open-system studies [
27,
33]. Taken together, these observations indicate that the camel-like GKSL framework offers a complementary perspective to the large body of work on quenched lattice/QFT models: it bridges master-equation approaches to entropy dynamics with the many-body literature on quenches, revivals, and memory effects, and thus may help to translate phenomena observed in closed interacting systems into experimentally relevant, dissipative settings.
3.3. Maximal Rate of Entanglement Entropy as a Marker of Information Exchange Transitions
Extended results on the evolution of mutual information (see
Appendix A) for these quantum systems show that the Lindblad framework described in
Section 1.3 generally drives systems toward classical-like information exchange. Nevertheless, in all quantum correlations we report here, we identify a key transition state for
,
,
, and
, which corresponds to an interval where the rate of change in entanglement entropy is maximized. This rate is given by
where
A represents the reduced state of
,
,
, or
. The maximal rate of information exchange is thus defined as
This period of maximal rate corresponds to the largest increase in entanglement entropy, contrasting with the overall decrease driven by dissipation. The plot of
is shown in
Figure 4 and highlights this transition. Notably, as we report in an extended version of this study in [
13], the mutual information also peaks around
, which is a period contained in the one-hump interval defined in Definition 1, and defined in Definition 1. Thus, we observe that the transition state (
) from Definition 1 corresponds to the maximum of the first derivative of the entanglement entropy for all states. This is particularly evident in the rate of change in the information exchange for the Bell states (
Figure 4), which exhibit only a minor “hump” in the entanglement entropy within the time interval
(see
Figure 3). This feature is reflected in
Figure 4 as an immediate maximum followed by a minimum in the rate of change in the entanglement entropy.
The maximum in the derivative highlights the period
as the interval of the most rapid rate of information exchange across all simulated systems. This suggests that the maximum rate of change in the entanglement entropy serves as a key indicator for identifying whether the information exchange is transitioning toward classical or quantum-like behavior: if it is classical, the rate of change decreases, while a quantum-like exchange is characterized by a maximum in this rate. This behavior is further supported by the mutual information, which decreases more slowly for the Bell states during this period (see calculation of the mutual information in the
Appendix A). Thus, calculating the rate of change in entanglement entropy provides a critical method for predicting whether a system is evolving from classical to quantum information exchange, or vice versa.
Notably, quantum correlations alone often fail to determine the direction of this transition; instead, they primarily indicate whether the system currently exhibits classical or quantum-like behavior. In contrast, when the entanglement entropy can be resolved, its rate of change robustly predicts the system’s evolution toward either regime. In
Figure 4, we observe that the transition state
from Definition 1 corresponds to the maximum of the first derivative of the entanglement entropy for all states. This is especially evident in the Bell states, which display a minor “hump” in the entanglement entropy within
(see
Figure 3). Correspondingly,
Figure 4 shows an immediate maximum followed by a minimum in the rate of change. These findings are consistent with recent studies [
34] demonstrating that quantum-to-classical transitions arise from the dynamical properties of systems in Hilbert space, rather than from operator non-commutativity, as reflected in our calculations.
3.4. Quantum Discord
Quantum discord gives a measure of the level of “quantumness” of a system, hence the degree to which the behavior of the qubits is classical or quantistic. This is important to be determined, since the relationship between “quantumness” and quantum relations is not always as expected. For instance, a system can behave classically, but still be entangled, which indicates that quantum discord is an important property of systems and can give better classifications of their behavior in an open quantum system simulation. We calculate and plot the quantum discord of our model states, the Bell states, the product states in (
19), (
21), and the mixed states in (
23), (
24) by the Khrennikov-picture, using the Hamiltonians and dissipation operators given by (
16) (
Figure 5). In this figure, we see that we obtain a stabilization of the quantum discord of the Bell states to a stable steady state after experiencing a small “hump” at the beginning of the simulation. The fact that the Bell states start from quantum discord of
D = 1 is correct as quantum discord measures non-classical correlations in mixed states, which Bell states initially are not. Clearly, the Bell states lose coherence and their correlation turns into a classical correlation, to a higher extent, by attaining a steady state with a discord of
D = 0.1. It is noteworthy that the decrease in discord is showing a “hump” during the transition state period of
, as well as for the previous correlations and as defined by Definition 1.
The quantum discord evolution in
Figure 5 reveals distinct dynamics for the entangled (
) and separable (
) states. The entangled state displays initial discord
bits, confirming to non-classical correlations, which suddenly increase to
bits in the interval
, and by
, it stabilizes to 0.05 bits. In contrast, the separable state maintains
throughout, as expected for a classically correlated system. Notably, the non-zero steady-state discord for
implies persistent quantum correlations despite environmental interaction.
The mixed states
and
behave similarly to the product states; however, with a stronger synchronicity during the evolution. Their synchronous behavior rests in their higher purity than the product states; however, their complete loss of quantum discord into the classical realm indicates that, however pure, their combination generates classical relations for their substates (
Figure 5).
From
Figure 5, we see a further confirmation that the period
is the critical transition, which delineates that the system is moving towards quantum-like information exchange by increasing the quantum discord for the product states and inducing a slower rate of negative change in the discord for the Bell state and the mixed state
. Indeed, this is equally well confirmed by the rate of change in the entanglement entropy of all the states, as an indicator of whether a system moves towards quantum-like information exchange or classical information exchange.
3.5. Entropy and Entanglement Entropy by Parametrized Coefficients of Mixed States
The goal of this final part of the numerical analysis of the GKSL equation is to study the camel-hump-landscape defined by Definition 1 for the entropy and entanglement entropy for mixed quantum states parameterized by
under Lindblad evolution. As shown in the Methods section, mixed states are constructed as convex combinations of pure states, where the coefficients
and
are defined by trigonometric functions of
(see Method section). The mixed state that is considered for this final calculation is defined as
where the pure states
and
are given by
So the convex combination of these two pure states forms the mixed state as
These states evolve according to the Lindblad master equation, governed by the camel-like evolution Hamiltonian as given by [
2]. The von Neumann entropy,
, is calculated at each time step to quantify the uncertainty or mixedness of the state. By varying
, the dependence of steady-state entropy on the parameterization is analyzed, providing insight into how the structure of the coefficients influences the quantum system’s dynamics.
From
Figure 6 (left), we see that the entropy dynamics observed in the simulations reveal a periodic behavior with a period of
determined by the trigonometric parametrization of the coefficients
and
using
and
. This periodicity, inherent to the symmetry of the parameterization, is reflected in the steady-state entropy values, which exhibit maxima at
and
, and minima at
and
, accordingly with the nature of
and
by the normalization condition
. These points correspond to configurations where the quantum system achieves the highest and lowest levels of mixedness, respectively (
Figure 6).
Figure 6 (left) shows that the periodic high-entropy states yield an increased uncertainty in the system, while the low-entropy states imply closer alignment to pure states (compare with
Figure 2). However, a key insight is that the camel-like landscape gradually vanishes as the entropy grows by the periodic change in the value of
(
Figure 6, left). By considering Definition 1, we see in
Figure 6 (left) that as the value of the coefficients evolves towards
and
as the transition state. This implies that the transition state is no longer a transition state followed by a relaxed steady state but a step towards a steady state of higher entropy than the transition state itself. By this, we can conclude that the camel-like behavior of the entropy given by [
2] given in
Figure 2 is sensitive to the parametrization of the coefficients of the quantum states.
Nevertheless, we find an equally interesting result in the evolution of the entanglement entropy (
Figure 6, left). The parametrization of the quantum state coefficients by
has no effect on the entropy of the reduced density matrices. This is expected, as entanglement entropy depends only on the structure of the total state and not on any direct interaction between
and
. As a result, the entanglement entropy for the mixed states follows the same transition-state-based landscape as when the coefficients are fixed (
Figure 3).
We can now construct the following theorem, relating the behavior of the entropy, entanglement entropy, the rate of information exchange, and the quantum discord in
Figure 2,
Figure 3,
Figure 4 and
Figure 5.
Theorem 1 (Classical versus non-classical evolution). Let be the density matrix of a quantum system at time t, and let denote the entanglement entropy of a bipartite subsystem of the system, where and is the reduced density matrix of subsystem A. Let furthermore J be the entire period of the evolution of the states under a completely positive trace-preserving (CPTP) map, and let .
Assume the following:
- 1.
The system evolves under a completely positive trace-preserving (CPTP) map (i.e., the GKSL equation), which ensures the physicality of the evolution.
- 2.
The entanglement entropy is differentiable with respect to time t.
Then, it follows:
- 1.
If for all t in some interval I, the system’s information exchange evolves towards classical-like behavior in I. This implies a reduction in quantum correlations, such as entanglement, and a shift towards classical correlations, where the system behaves more classically.
- 2.
If for all t in some interval I, the system’s information exchange evolves towards quantum-like behavior in I. This implies an increase in quantum correlations, such as entanglement, which enables non-classical phenomena like quantum teleportation.
- 3.
The measured substate shifts the quantum discord toward 0 by adding classical contributions, while the unmeasured substate shifts it toward 1 if , and conversely if .
Proof. Since the system evolves under a CPTP map (completely positive trace-preserving map), the density matrix
evolves as
, where
is a CPTP map. This guarantees that
remains a valid density matrix. Furthermore, we base our proof on the universal relation:
The entanglement entropy
is defined as the von Neumann entropy of the reduced density matrix
:
By assumption,
is differentiable, and we analyze the implications of
and prove
points 1 and 2.
Case 1: If for all t in some interval I, this implies that the entanglement entropy is decreasing over time. Since entanglement entropy measures quantum correlations, its decrease suggests that quantum information is being lost, likely due to decoherence. This results in a reduction in quantum correlations, leaving only classical correlations dominant, signifying classical-like behavior.
Case 2: If for all t in some interval I, then entanglement entropy is increasing. Since increasing entropy indicates an increase in quantum correlations, the system is evolving towards more non-classical correlations, including entanglement and quantum discord (see relevant sections). This suggests that the system’s information exchange is dominated by quantum effects, signifying quantum-like behavior.
For point 3, recall that the quantum discord
quantifies the difference between total and classical correlations after a measurement on subsystem
B. Measurements on
B tend to remove quantum correlations associated with
B, driving discord toward zero by increasing classical contributions. Conversely, the unmeasured subsystem
A may retain or increase quantum correlations when
, pushing discord toward 1. This behavior follows from the monotonicity of quantum mutual information under the local GKSL equation (see
Appendix A), which ensures that local measurements reduce quantum correlations in the measured subsystem. Hence, the sign of
governs the direction of discord change depending on which substate is measured. □
Remark 1.
- 1.
The theorem assumes that entanglement entropy is a valid measure of quantum entanglement, which holds for pure and product states, by considering the results from this study. The theorem also holds for mixed states since the parametrization by ϕ of the coefficients of the subsystems does not affect the entanglement entropy landscape.
- 2.
The theorem applies to bipartite systems. For multipartite systems, the behavior of entanglement entropy can be more complex, and additional considerations may be necessary.
- 3.
The theorem assumes differentiable entanglement entropy, which may not hold in all cases (e.g., during quantum phase transitions or in open quantum systems with non-Markovian dynamics or under discrete Markovian quantum dynamics).
3.7. Topological Analysis on the Bloch Sphere
The set
in (
48) can be classified topologically on the Bloch sphere
. We establish coordinates on this manifold using the standard spherical parameterization:
This allows us to combine topologically all stable solutions (which decohere slowly under the camel-like framework) as the complement
, forming a non-compact topological space homeomorphic to a 2-sphere with 5 punctures, where each puncture corresponds to one of the highly unstable states in
in (
48) (recall two of the six states are the same on the Bloch spheres’ South Pole giving a total of 5 punctures). The set
of unstable states in (
48) can be interpreted as the zero-set of a map
where
f vanishes precisely at these six specific points. We can express this map as
where
captures the pole structure (with zeros at
), and
represents the angular dependence (with zeros at specific
values depending on
). Here, we can form the function:
This function vanishes precisely at the eight points of set
in (
48). The component
captures the latitude dependence, vanishing at
, while
encodes the longitude constraints, vanishing at the South Pole and at four specific longitudes
when
. The factor
in
ensures it vanishes at the South Pole regardless of
. The reason we construct this function which vanishes for the unstable points selected from
Figure 7 is based on forming a topological analysis of where the stable states arise, which are states that satisfy a gradual decoherence and dephasing under the camel-like framework. These decoherence-instable states in the set (
48) are indeed still valid as solutions to the GKSL equation; however, as they decohere immediately, they form a special case of states that pose specific consequences for quantum computation purposes, as we briefly mentioned above. Therefore, isolating these states on the sphere can give us a topologically intuitive view of how states develop stability and instability to decoherence effects by the dissipation operator in the GKSL equation.
Thus, the function (
53) characterizes the set
of the unstable states as the zero set, where
, while (more) stable states correspond to points where
. We can use this framework on the Bloch sphere to start by performing a gradient flow analysis to study the evolution of states as they approach the highly unstable states in
on the Bloch sphere.
3.7.1. Gradient Flow Analysis on the Bloch Sphere
The function
where
,
are defined in (
53) is constructed as a scalar potential function (or a Lyapunov-like function) on the Bloch sphere,
, whose global minima are precisely the five unstable points in the set
in (
48). These minima occur at the four points on the three-quarter circle (
) with
, and at the South Pole (
). By computing the negative gradient,
, we define a gradient dynamical system
The integral curves (trajectories or flow lines) of this vector field represent paths of steepest descent on the potential surface
h. Plotting these trajectories visualizes the basins of attraction for each of the five special points under this gradient flow. This analysis provides topological insights into the phase portrait of the original system by partitioning the Bloch sphere according to the influence of these unstable points and their effect on solving the system using the ODE45 method. By visualizing the trajectories that follow
on the Bloch sphere, we can identify the resulting gradient field structure, where the set
of unstable points forms a
symmetry in the
direction that is preserved by the gradient flow due to the periodic structure of
. This partitioning reveals how the Bloch sphere is divided into the two distinct regions, Northern and Southern hemisphere, where quantum states in each region are attracted to their corresponding decoherence hotspot, either the South Pole, or the four concentric points on the Braiding ring (
(see
Figure 9)).
The flow lines in
Figure 9 indicate that unstable points near
act as attractors for certain states descending from the North Pole. As shown in the figure, the gradient structure around the South Pole reveals a complex flow pattern towards the attractor at the South Pole. The deep red region near the North Pole, corresponding to the highest gradient values, shows that decoherence and dephasing drive states as a source toward the South Pole, while the gradient flow (white arrows) reveals this as the dominant evolution pathway. The gradient lines evolve globally away from the third quarter circle (
), exhibiting bidirectional flow, toward the South Pole (dominant pathway), and back toward the North-West and North-East near four specific points on the third quarter circle. This behavior, clearly visible in
Figure 9 (South Pole image), reveals that the third quarter circle acts as a topological transition region, by the specific Lyapunov function. The gradients around both the South Pole and this quarter circle show bifurcating flows, creating what may be described as an open barrier that introduces disorder in the evolution of states.
These observations demonstrate a topological tendency of the solutions of the GKSL equation where the third quarter-circle at
acts as a transition region and the states can evolve both slightly Northward and due Southward. The system thus exhibits a topological transition containing unstable states and trajectories, with gradient-driven evolution creating complex flow patterns around these selected unstable points in (
48).
It is worth noting that these critical points simply emerge from the global shape of
h and do not affect the form of the gradient field itself. The “camel-like” geometry of the entropy thus arises solely from the interplay of the Hamiltonian and dissipative terms encoded in
h, rather than from any special, localized influence of the unstable points. This implies that the gradient flow trajectories around the unstable point are also real in a physical experiment, as long as we consider the points
in (
48) as “special” or “different”, and hence, are evaluated in
h by being outside the desired stability against decoherence of states under the camel-like framework.
Finally, we note that the trajectories on the Bloch sphere represent a multitude of different transitions between pure states under an entirely unitary evolution, as they remain on the surface of the Bloch sphere. In a physical case, this would be represented by a series of diffractometers for beamed photons, which preserves the purity while changing their polarization. Conclusively, in this section, we have thus constructed a Lyapunov function based on five highly unstable quantum states under the camel-like framework by their entropy plot in
Figure 7, in order to design an trajectory path for the evolution of states under a completely unitary and real experiment in quantum information, so that we can predict the evolution of quantum states under the camel-like framework. It is also worth noting that the North Pole represents the most stable point in terms of decoherence resistance, and the South Pole represents the most unstable. By the gradient field we have designed here, the instability of states is thus not related to instability/stability towards decoherence but stability and instability based on the gradient flow values that
attains on the Bloch sphere, which are useful for the mapping we have presented in
Figure 9.
3.7.2. Topological Constraints on the Bloch Sphere
With the notion that the North Pole acts as a source and the South Pole as an attractor by considering the gradient field dynamics, we can form a basic basin map and analyze the evolution of the stability of the quantum states, subdivided into topological basins. We develop this analysis from the quantum system’s fundamental functions given in (
53). The gradient flow dynamics are derived from the potential function
, with the evolution equations
which we derive from the Riemann flow on the
sphere [
35]. Following these dynamics, we classify the quantum state space into distinct basins:
The boundaries of each basin are defined as a manifold with boundary and can thus be viewed as compact topological sets on the Bloch sphere, which can be assigned well-defined closures, and where any calculation will converge within the established boundaries. We can thus define these basins topologically as
where
and
. Here,
represents the
-neighborhood around the point
, and
are distinct basins. The unstable points (in terms of decoherence) of the system given in (
48) are characterized by:
We plot the subdivision of basins on the Bloch sphere in
Figure 10 based on this Riemannian flow geometry, where the topological ridge sinks toward the third quarter circle on periodic longitudes, where we have the Braiding ring, composed of unstable states. This gives two basins, displayed in red and blue, where field lines move towards the South Pole from the North Pole, north of
.
In
Section 3.7.4, we shall discuss the physical implications of this topological manifold. Meanwhile, we form a theorem summarizing the results from topological analysis of the solutions of the GKSL equation under the camel-like framework.
Theorem 2 (Global Convergence on the Bloch Sphere)
. Consider the gradient flow of the function on , whereThen, the following:- 1.
The unstable points in terms of rapid decoherence consist of the critical points by the South Pole , and four saddle points at . Also, the North Pole is a critical point, however, it is stable in terms of decoherence.
- 2.
The South Pole is a global attractor: for all initial conditions , the flow converges to .
Proof. (1) Critical points satisfy . Since , we have . At : but , and , giving isolated critical points. At : both due to the factor . At : and when or , yielding four saddle points.
(2) Since
and
for all
, the South Pole is the unique global minimum. Furthermore, along any trajectory of the gradient flow
, the function
h decreases:
This means that
h always decreases (or stays constant) along trajectories, which can be seen in the gradient plot. Since the sphere is bounded and
everywhere,
h must approach some limiting value. The only points where trajectories can stop are where
(the critical points). At the North Pole and saddle points on the third quarter circle, small perturbations cause trajectories to move away (since
h decreases from these points). However, at the South Pole, where
(the minimum), trajectories cannot decrease
h further, so they must stop. Therefore, all trajectories eventually reach the South Pole, making it the global attractor. □
Remark 2. It is important to note that the notion of instability described by the gradient flow differs from the instability observed in the von Neumann entropy plots in Figure 7. The gradient flow instability arises from the rate at which the Lyapunov-like function changes over the Bloch sphere, whereas the decoherence-related instability of the states in the set is due to their susceptibility to the dissipation operator in the GKSL equation under the camel-like entropy framework, which we precisely observe in Figure 7. The gradient field of the Lyapunov function thus illustrates a descent from decoherence-stable states (North Pole) toward decoherence-sensitive ones (South Pole). 3.7.3. Evolution by Purity in the Bloch Ball
In this section, we investigate the evolution of pure states on the Bloch ball (where the entire space inside
is considered as a vector space of solutions), following the properties of the GKSL equation under the camel-like framework. The evolution starts from pure states (on the Bloch sphere—
) to mixed states in the center of the Bloch ball (a dense globular manifold). The evolution on the Bloch ball is tracked to simulate a pair of entangled particles that are passed through a series of diffractometers and electromagnetic fields, rearranging their polarity and reducing their state of purity into a mixed state. Naturally, this process is entirely irreversible due to the uncertainty principle and evolves from the Bloch surface to the center, following the paths allowed by the GKSL equation under a camel-like framework. These paths should not be confused with the gradient paths in the
manifold from
Figure 9, where the gradient field lines determine the evolution paths from pure metastable states to stable pure states (stability in the sense of being an admissible solution to the GKSL equation).
The universal steady state, the great attractor of all states under the GKSL equation in the camel-like framework, is computed in
Python and presented as the density matrix shown in
Figure 11.
3.7.4. Relevance for Physics
The Lyapunov function with four singularities on an unstable equilibrium circle represents a geometric quantum control framework that actively avoids the most decoherence-prone states on the Bloch sphere. These four points at
are fully admissible pure quantum states, but they exhibit the fastest decoherence rates in the system, making them the most fragile states to maintain coherence by the notion of stability of states in quantum information experiments [
36,
37]. By strategically placing singularities of
at these four maximally decohering points, the gradient flow creates a control landscape that steers quantum trajectories away from these vulnerable regions, effectively implementing a decoherence-avoiding quantum control protocol [
38]. This approach differs from traditional geometric phase schemes by explicitly incorporating decoherence information into the Lyapunov function design, where the unstable manifold at
represents the locus of states with maximal environmental coupling [
39]. The resulting flow pattern, with its characteristic basin structure, ensures that quantum states rapidly escape the high-decoherence region and converge to the more stable states, thereby minimizing exposure to decoherence throughout the evolution [
36,
38]. This framework can be experimentally realized using polarization qubits passing through engineered diffractometers, where polarization-dependent diffraction losses implement the Lyapunov function
. Specifically, using a first diffractometer, we create the
dependence through selective polarization, while a second holographic diffractometer implements
by creating null diffraction points (the unstable states we call "singularities") at the four decoherence-prone states [
40,
41]. The measurement backaction from photons lost to higher diffraction orders drives the remaining zero-order photons along the gradient flow trajectories, effectively steering them away from the fragile states at
toward the decoherence-protected South Pole [
42].