in

Dynamic hovering for uncrewed underwater vehicles via an error-separation-based cooperative strategy


Abstract

Uncrewed underwater vehicles(UUVs) play an indispensable role in ocean resource efficiency utilization due to their security and cost-effectiveness. However, the significant uncertainties, time-varying disturbances, and unstructured subsea environments pose challenges for UUVs in achieving precise and efficient marine missions. To address these challenges, this study introduces a novel cooperative control framework for UUVs. The proposed framework minimizes the high control efforts of sliding mode control while preserving robustness, enabling efficiency and precision for long-duration dynamic hovering missions of UUVs. Specifically, a key innovation is the development of a deviation separation strategy, which, for the first time, decouples hovering deviations into task-specific and anti-disturbance components using an influence function. This enables real-time disturbance estimation without prior knowledge enabling adaptive disturbance compensation. By cooperating between LQR and SMC, the proposed method avoids the performance conflicts commonly observed in single-controller schemes. This structure improves compensation accuracy, robustness to disturbances, and energy efficiency across various operating conditions. The results demonstrate that the proposed cooperative control strategy effectively counters current perturbations by leveraging the real-time insights from the error segregation, while concurrently executing high-precision hovering tasks with low control costs. This work advances UUVs control, offering a versatile solution for complex underwater tasks.

Similar content being viewed by others

Detection and tracking of ocean layers using an AUV with UKF based extremum seeking control in the Baltic Sea

Research on the dynamic performance and motion control methods of deep-sea human occupied vehicles

Multicooperation of Turtle-inspired amphibious spherical robots

Introduction

The Earth’s surface is overwhelmingly aquatic with 71% submerged in oceans, rivers, and lakes, most of which remain unexplored. These underwater bodies host resource-rich ecosystems that significantly impact human life, both directly and indirectly1,2. As the depletion of land-based nonrenewable resources accelerates, the imperative to comprehensively explore and exploit these ocean resources is intensifying for human advancement. To unlock these ocean resources sustainably, researchers and scholars have developed advanced technologies such as underwater vehicles named “Jiaolong” equipped with intelligent systems3,4. However, the deep-sea environment is characterized by unstructured terrain, extreme hydrostatic pressures, and unpredictable disturbances, which inevitably compromise operational efficiency and pose significant safety risks for manned submersible vehicles. Consequently, uncrewed underwater vehicles (UUVs) have garnered growing attention owing to their enhanced safety and cost-efficiency in the exploration and sustainable utilization of underwater resources5,6.

Up to now, UUVs have been widely applied in many marine tasks such as ocean exploration, seafloor mapping, deep water mining, etc7,8. In the above-mentioned applications, control, planning, and localization are all important9,10. Among them, control is the core because no operators are inside the UUV to make adjustments11,12. In the above-mentioned applications, designing a reliable controller is essential for successful marine tasks due to the absence of operators inside the UUV to adjust the vehicles to arrive at the desired position. Some efforts have been made to integrate the advanced control strategies into UUVs13. For instance, a 6-degree-of-freedom (DOF) UUV has been successfully implemented with a classical proportional-integral-derivative (PID) control algorithm to accomplish precise trajectory tracking tasks14. Following a similar framework, researchers also proposed an enhanced PID algorithm and demonstrated its efficacy in depth-tracking missions for the hovering UUV15. Moreover, a novel PID approach coupled with computational fluid dynamics is developed for UUVs to enable precise trajectory tracking while ensuring constant velocity16. Despite the straightforward implementation of PID algorithms in UUVs, their broader application is hindered by the intricate challenges of parameter optimization and their inherent limitations in complex tasks.

To address these limitations, the linear quadratic regulator (LQR) is introduced for UUVs to solve the control inputs by optimizing a cost function, subject to the performance objectives of complex missions. The LQR method is proposed for UUVs, yielding a 32.04% improvement in tracking performance compared to the traditional PID algorithm17. Given the excessive optimization time per control step in complex nonlinear systems, a novel path-following LQR approach is proposed, utilizing a simplified deep deterministic policy gradient algorithm to enhance computational efficiency and control precision18. Furthermore, a prescribed-performance fuzzy LQ fault-tolerant control framework is presented for UUVs, tailored for three-dimensional trajectory tracking in complex marine environments19. The developed scheme effectively mitigates unknown actuator faults, thereby ensuring precise control performance under adverse operational conditions.

The primary focus of the discussed control strategies has been to maintain the vehicles as closely as possible to their desired trajectories under ideal conditions. However, UUVs often navigate in highly uncertain and unknown circumstances, which may reduce the efficacy of these approaches and even result in complete operational failure. In particular, thrusters’ performance variations, payload changes, and tether drag altering will lead to high degree uncertainties of UUVs’ parameters. Then, the effectiveness of control strategies, depending on accurate models, is drastically limited. On the other hand, UUVs inevitably suffer from the currents and waves caused by tidal energy and topographical undulations in the deep sea. These disturbances can also result in unregulated and unpredictable movements of UUVs20,21. Therefore, designing a high-performance controller to compensate for the influence of uncertainties and disturbances is crucial for UUVs motion.

To handle the above challenges, the sliding mode control (SMC) based control framework emerges as a highly suitable approach for high-precision missions of UUVs22. In SMC, the control law is designed and imposed through a sliding manifold23. The high-frequency switching input mechanism inherent in SMC enables rapid compensation for perturbations, ensuring system stability and consistent performance despite parametric variations and external disturbances. Given this advantage, SMC is widely applied in UUVs control24,25. Liu and Zhang proposed a terminal SMC by integrating a piecewise error transformation into the sliding manifold, effectively compensating for hydrodynamic disturbances in marine systems26. For the precise position regulation of remotely operated UUVs, an SMC scheme integrating time-delay estimation is presented to address hydrodynamic disturbances effectively27. Aiming to compensate for the bounded disturbances while minimizing control chattering, a hybrid control strategy combining integral sliding mode and super-twisting algorithm for precise 3D trajectory tracking of 4-DOF UUVs is developed28. Despite the demonstrated efficacy of SMC algorithms in UUVs, it requires high control efforts. This expensive control cost is primarily driven by the necessity for continuous high-gain switching to maintain system trajectories on the sliding manifold. Moreover, the tracking accuracy often experiences significant degradation under extreme hydrodynamic loading conditions due to the fixed feedback gain parameters inherent in SMC, which lack adaptability to dynamic environmental variations.

In fact, UUVs commonly need to perform ocean missions for several hours in different regions. Throughout this period, UUVs are subjected to a variety of operational environments, including disturbance from ocean currents, modifications in seabed topography, the tranquility of calm sea surfaces, etc. Each environment imposes different primary control objectives. For instance, strong robustness is required when the ocean current occurs, and high control accuracy and low control cost are preferable in the calm deep sea. However, it is difficult for a single control strategy to satisfy all the control objectives mentioned above during the whole mission period. Therefore, how integrating multiple control methodologies into a unified hybrid framework to preserve the strengths and complement its drawbacks could be a very promising solution for UUVs missions. This is the primary motivation of this study.

Recently, advances have been made to enhance robustness and adaptability under complex underwater conditions. For example, a distributed tube model predictive control (tube-MPC) is used to handle communication delays and disturbances of multiple UUVs under dynamic marine environments29,30. To address the low-visibility conditions in controlling heterogeneous multi-UUV systems, researchers used graph-based models and relative states to achieve robustness and real-time performance in simultaneous localization and mapping31,32. Recent papers also combine integrating event-triggered MPC, nonlinear Lyapunov control, and collision avoidance to enhance tracking accuracy, safety, and computational efficiency when facing current disturbance16,33.

Motivated by these developments, this study proposes a cooperative control (COC) framework that combines the SMC algorithm’s strong robustness and rapid responsiveness with the high accuracy and low control cost offered by LQR. Specifically, a deviation separation approach is developed for the first time to segregate the errors during UUVs mission. Based on the real-time error separation result, the deviation from the current disturbance is rejected by the SMC, and LQR achieves the precise hovering task. Through this approach, UUVs equipped with the proposed COC framework can better address varying operational challenges and fulfill diverse mission objectives. Contributions of this work are summarized as follows:

  1. (1)

    A novel COC protocol is developed for UUVs to adeptly maintain the superior robustness of SMC architecture while meticulously leveraging the efficacy of LQR to mitigate its inherent high control efforts and frequent switching. With the designed cooperative framework, the controlled UUVs can perform underwater missions efficiently and precisely with low control consumption for long operation duration.

  2. (2)

    A deviation separation strategy is introduced for the first time to address the challenge of coupling underwater missions. By leveraging the influence function, we establish an analytical relationship between variations in disturbances and corresponding changes in control performance, enabling the estimation of disturbances based on control outcomes. Consequently, the deviation in hovering within the marine environment is decoupled into two sub-deviations: one corresponding to the hovering task and the other to the anti-disturbance task. This strategy facilitates the design of a tailored controller for each decoupled task, thereby enhancing overall system performance.

  3. (3)

    This proposed controller can mitigate a wide range of disturbances without requiring prior knowledge, such as the size or source of the ocean disturbance. The disturbance is estimated at each point in real-time using the deviation separation strategy. Consequently, the controller can effectively compensate for the effects of the disturbance without the need for additional information. This approach ensures that the controlled UUVs remain stable, even in the event of a failure of one of the sub-controllers.

Results and discussions

The hovering of UUV plays a crucial role in various underwater missions. As shown in Fig. 1, the primary objective is to maintain the position (xyz) and orientation (ϕθψ) at a desired pose while performing tasks such as object retrieval, device connection, or underwater docking. During these processes, the UUV’s previous motion state and the ocean currents’ movement can significantly impact the pose in the next moment. As a result, two primary challenges emerge for UUV underwater operations: maintaining hovering at the desired pose and counteracting current disturbances. We address the coupled tasks of hovering and disturbance suppression. A variety of complex underwater missions, including military reconnaissance, hull inspection, scientific exploration, and fisheries, can be analyzed as different combinations of these fundamental control subtasks, with UUV hovering and anti-disturbance being a consistent requirement across them. This classification framework distinguishes itself from traditional, application-centric approaches by prioritizing the fundamental control of the UUV’s motion states.

Fig. 1: Overview of the UUV underwater hovering mission.

The vector [xyzϕθψ]T represents the pose of the UUV in the inertial reference frame, where xyz denote the position, and ϕθψ represent the pitch, roll, and yaw angles, respectively. The velocity vector [uvwpqr]T in the body-fixed frame is composed of the linear velocities uvw, and the angular velocities pqr. The total disturbance affecting the system consists of both current disturbance and thruster noise, which significantly impact the hovering task, causing the UUV to deviate from the desired pose.

Full size image

Current disturbances can be attributed to two primary sources. The first is wind-driven currents, generated by wind forces acting on the ocean surface. These currents typically weaken with increasing depth, with the affected depths extending several hundred meters. The second source of disturbance arises from changes in seawater temperature and salinity, which alter the density distribution of the ocean. These variations cause discrepancies between the ocean’s isobaric and equipotential surfaces, thus generating ocean currents. Such disturbances primarily affect the forces and moments acting on the UUV, leading to deviations in the control inputs generated by the thrusters. In addition to the current disturbance, random noise can also arise from various independent events, such as the collapse of tiny bubbles produced by the thrusters. Although this kind of noise typically has a minimal impact on the UUV’s overall performance, continuous correction is necessary to prevent the accumulation of errors over time.

Related research indicates that the abovementioned disturbances can be mitigated using SMC, which applies continuous high-gain thrust to counteract current disturbances. On the other hand, optimal strategies such as the LQR can effectively address the hovering tasks influenced by thruster noise. However, SMC treats all deviations as significant disturbances, applying high-gain thrust continuously, resulting in inefficient energy consumption. While LQR is more thrust-efficient for hovering tasks, it fails to suppress current disturbance. Therefore, in practical scenarios, relying on a single control strategy cannot simultaneously meet the multiple critical objectives of low power consumption, robust anti-disturbance performance, and high precision for UUV hovering tasks.

To address this, we propose a COC strategy, as illustrated in Fig. 2, which combines the robust and rapid responsiveness of the SMC algorithm with the energy efficiency of LQR. By employing a deviation separation mechanism, we decompose the complex task of maintaining the UUV’s desired pose in dynamic ocean environments into two distinct sub-tasks: the traditional hovering and anti-disturbance tasks. We then design two sub-controllers (LQR and SMC) tailored to each task based on their strengths. Finally, the thrust outputs from both sub-controllers are integrated and constrained through engineering techniques to form the cooperative controller, ensuring optimal performance while maintaining robustness.

Fig. 2: Cooperative control (COC) framework.

This framework operates through a workflow initiated by the central computational host, which broadcasts predefined pose information to the entire system. The deviation between the current and desired pose is quantified and decomposed into two sub-deviations using the deviation-separation algorithm. The sub-controllers, SMC and LQR, compute the thrust based on these sub-deviations. The resulting thrust inputs are integrated and modulated within predefined limits before being transmitted to the thruster system. This ensures the precise application of forces and moments to the UUV, thereby aligning its behavior with the intended dynamics.

Full size image

To evaluate the engineering feasibility of the proposed COC strategy, a computational analysis was performed on a standard desktop platform (Intel Core i7-13700F, 16GB RAM). All algorithms—including deviation separation, LQR, and SMC sub-controller computations—were implemented discretely with a fixed control cycle corresponding to a 1 Hz sampling frequency over 2000 steps. Under these conditions, the total execution time of the entire simulation was 20.687 s, while the cumulative execution time for all control loop computations was measured to be 0.124 s. This duration includes all control matrix operations, sliding surface evaluations, and separation matrix computations, amounting to less than 0.5994% of the total available time budget. These results confirm the strategy’s suitability for real-time implementation.

Control accuracy comparisons under different control policies

This section mainly considers the horizontal non-rotational ocean current during the UUVs hovering mission described in Fig. 1. Results and analysis under other interference conditions could be found in Sections S.4–S.6 of the supplementary file. Referred to reference34, the total disturbance is a combination of current disturbance (a sine function with an amplitude of 1 and a frequency of 5 Hz) and thruster noise (a standard Gaussian distribution signal). The initial pose of the UUV is set as [45, 45, 110, 0, 0, 0]T, representing its starting position and orientation in the global coordinate system, where the first three values denote the position in meters along the x, y, and z axes, and the last three values indicate roll, pitch, and yaw angles in radians. This initial pose simulates the UUV’s offset from the desired hovering location at the start of the mission. The target hovering pose is set as [50, 50, 100, 0, 0, 0]T., representing the desired location and orientation the UUV aims to reach and maintain during the hovering task. Assuming the current disturbance affects the system at the 1000th step. To illustrate the effectiveness and superiority of our proposed COC algorithm, the control performance of SMC and LQR under parameters R = I, Q = 100I, and z = 2 is compared. Detailed mathematical derivations of SMC and LQR algorithms can be found in Sections S.2–3 of the supplementary note.

Thrusts computed by LQR, SMC, and COC are applied to UUVs. Figure 3 shows the current and desired pose deviations. The embedded inset figures on certain phases provide a detailed view of a more precise comparison of tracking performance. To better analyze the results, we roughly divide the process into three stages: initial value convergence, stabilization, and anti-disturbance. In particular, during the initial value convergence stage (within 650 steps), the gray lines reach the desired point (zero in this experiment) faster and closer than the blue lines. The above trend indicates that the COC framework has a faster convergence speed and higher steady-state accuracy than LQR. This enhanced performance arises from integrating SMC’s advantages during the initial phase, where COC leverages significant, high-frequency thrust adjustments to rapidly drive the UUV’s posture toward the desired state. By combining the robustness of SMC with the precision of LQR, COC achieves superior dynamic response and stability.

Fig. 3: Deviation responses of UUV across all six degrees of freedom under different methods.

a illustrates the response of position x, while b depicts the response of position y and c presents the response of position z. d, e, f show the responses of orientation ϕ, θ, and ψ, respectively. The ocean current disturbance is introduced at the 1000th step in all subfigures. The blue, red, and gray lines represent the deviation responses under the LQR, SMC, and COC control strategies.

Full size image

In the stabilization phase (between 650 and 1000 steps), the blue lines in all subfigures converge to within at least 5% of the maximum deviation from the desired state. In contrast, red lines in all subfigures of Fig. 3 other than Fig. 3f remain closer to the desired state but oscillate within 1% of the maximum deviation. This phenomenon indicates a static error exists between the UUVs hovering pose and the desired pose under the control of LQR. Although SMC has no static error, Fig. 3f shows some situations in which the pose states oscillate. During the initial 1000 steps in Fig. 3e, the slow convergence of the blue lines prevents it from reaching a stable state, indicating that LQR struggles to stabilize the system rapidly in the presence of significant initial deviations or sudden disturbances. This limitation stems from LQR’s reliance on an accurate model.

The COC framework overcomes the drawbacks of both LQR and SMC, as evidenced by the gray lines’ steady trends in Fig. 3. These lines exhibit smoother trajectories, with reduced oscillation amplitude and frequency compared to SMC, while maintaining more minor pose deviations than LQR. These trends demonstrate that when the deviation state of COC achieves a high accuracy, it does not overreact to minor deviations, which ensures the system will not oscillate. Further, the cooperative mechanism ensures that, once the pose deviation converges within a desired level, the thrust input transfers from a severely oscillating one to an efficient and stable one. Therefore, the COC system only needs a small thrust input to stabilize the system, thereby minimizing control consumption while ensuring accuracy.

During the anti-disturbance phase (after 1000 steps), the gray and red lines remain closer to the desired state than the blue lines, demonstrating that both SMC and COC achieve higher steady-state accuracy than LQR. However, the lines of SMC exhibit occasional high peaks and persistent oscillations. For instance, between 1280 and 1360 steps in Fig. 3a, the red line peaks above the blue line, indicating that SMC can induce significant deviations in the pose state. Similarly, in Fig. 3f, the fluctuation of the red line reaches a peak of 70°, indicating that the roll angle of the UUV hull is approaching a critical lateral capsizing threshold. Moreover, the red line in this subfigure oscillates, failing to reach a steady state before subsequent disturbances occur. These issues arise because SMC treats all disturbances as maximum disturbances, often applying high-frequency and high-thrust corrections even for minor disturbances, leading to severe overshoot.

However, the gray lines in the same period do not exceed the blue lines, indicating that COC effectively avoids the overshoot problem associated with aggressive control inputs. By utilizing the deviation separation mechanism, the COC framework dynamically modulates the weight ratio between LQR and SMC, effectively mitigating overshoot in the presence of minor disturbances. Ensures robust and precise control, overcoming the inherent limitations of both SMC and LQR in dynamic and uncertain environments.

As discussed in the previous section, the effectiveness of the COC framework relies heavily on the performance of the deviation separation mechanism. To vividly illustrate the impact of this mechanism, Fig. 4 compares the separated sub-deviations. It delineates the relative contributions of thruster noise and current disturbance to the total deviation, offering critical insights into the system’s dynamic behavior. From the beginning to the 650th step, the purple region dominates, indicating that the sub-deviation caused by the current disturbance occupies most of the total deviation. This dominance arises because the initial values deviate from the desired state, similar to a sudden current disturbance affecting the UUV. As a result, the purple region predominates, and according to the collaborative control mechanism, the proposed COC algorithm is primarily governed by SMC. Due to SMC’s superior control accuracy and dynamic response, the COC framework achieves faster convergence and more minor steady-state deviations than LQR during this phase.

Fig. 4: Comparison of the proportion of sub-deviations from different origins.

af are the proportion comparison of position x, y, z and orientation ϕ, θ, and the ψ, respectively. The pink part represents the sub-deviation from thruster noise and the purple part represents that current disturbance.

Full size image

Upon focusing on the stabilization phase (from step 650 to 1000), the purple region steadily diminishes, signifying that the sub-deviations stemming from the initial pose have progressively converged. At this juncture, the COC strategy transitions to an LQR-dominated regimen, which persists until the onset of the current disturbance. As previously discussed, the aggressive thrust application of SMC is primarily responsible for the oscillations in the pose state. COC addresses this issue by integrating the thrust constraints of LQR. With the pose state that has converged and oscillated within a narrow range, COC achieves the best control accuracy with limited thrust and mitigates the oscillation cases. During the initial stabilization process, COC harnesses SMC’s rapid convergence and high precision while circumventing the drawbacks of substantial overshoot and protracted oscillations, offering a refined control solution for the UUV.

Then, focus on the post-1000 steps in Fig. 4. The total deviation increases primarily due to the growth of the pink segment, which represents the sub-deviation caused by thruster noise. This trend persists until approximately 1200 steps, when the purple area, the sub-deviation from the current disturbance, begins to emerge. The delayed onset of the purple area can be attributed to the fact that, in the early stages of current influence, the total pose deviation has not yet accumulated sufficiently to trigger adjustments in the COC’s collaborative strategy. During this phase, the COC remains predominantly governed by LQR, which is less effective in mitigating the immediate effects of current disturbances. As a result, the deviation accumulates until around 1200 steps, after which the purple area expands, and its contribution to the total deviation increases.

This observation reveals the dynamic interplay between the sub-deviations, where the dominance of one is swiftly mitigated, only for the other to become more pronounced. This cyclical process gradually diminishes the total deviation, enabling the UUV’s pose state to approximate the desired state continuously. The underlying mechanism driving this behavior can be explained as follows: when the COC transitions from an LQR-dominated to an SMC-dominated strategy, the impact of current disturbance and its associated sub-deviation begins to subside. As the sub-deviation from the current disturbance decreases, the relative contribution of the thruster-noise sub-deviation to the total deviation increases, prompting a shift back to an LQR-dominated strategy within the COC framework. This cyclical regulation results in the observed proportional fluctuations. Throughout this fluctuation period, the UUV’s pose accuracy is iteratively refined, and the system efficiently manages thruster consumption, thereby avoiding unnecessary power expenditure. This adaptive response underscores the effectiveness of the proposed COC in maintaining optimal control performance despite environmental disturbances.

Thrust and control performance comparisons under different control policies

In UUV hovering missions, thrust consumption is a critical factor, as is the system’s anti-disturbance performance and hovering error. In the above experiment, eight thrusters generate the needed force and moment, including two main thrusters, four vertical thrusters, and two lateral thrusters. This subsection presents a comparative analysis of the thrust curves under LQR, SMC, and COC control strategies. As illustrated in Fig. 5, during the initial phase of the convergence process (within 300 steps), the red lines, representing the SMC strategy, exhibit significant overshoot and oscillation. This behavior indicates that SMC utilizes an aggressive thrust output to mitigate deviations caused by initial conditions swiftly. However, this approach often leads to overshooting and thruster output saturation.

Fig. 5: Thrust variations of each thruster under the influence of current disturbance.

a Represents the thrust variation of the left-main thruster. b Illustrates the thrust variation of the right-main thruster. c Shows the thrust variation of the left-back vertical thruster. d Displays the thrust variation of the left-front vertical thruster. e Presents the thrust variation of the right-rear vertical thruster. f Demonstrates the thrust variation of the right-front vertical thruster. g Captures the thrust variation of the front-side thruster. h Delineates the thrust variation of the rear-side thruster. The thrust profiles are represented by blue, red, and gray lines corresponding to LQR, SMC, and COC, respectively.

Full size image

In contrast, COC uses a collaborative strategy that inherently constrains thrust output, mitigating excessively aggressive thrust responses at the expense of a slightly extended adjustment period. For instance, as illustrated in all six subfigures of Fig. 5, COC avoids the extreme thrust amplitudes ranging from  −300 N to 200 N observed in SMC. As a price, it achieves stabilization within 200 steps. During the latter phase of the convergence process, the red and gray lines exhibit high-frequency oscillations within the range of  −50 N–50 N. In contrast, the blue line, representing LQR, maintains smaller thrust values. This case indicates that SMC and COC continue applying thrust to suppress disturbance. In summary, COC effectively mitigates the issues of overshoot and thruster saturation associated with aggressive thrust outputs while maintaining precise pose control during the initial value convergence process.

After 650 steps, the lines of COC in Fig. 5 hold at a lower value and exhibit no oscillations compared to those of SMC. It indicates that, at this stage, COC achieves high control accuracy without relying on high-frequency thrust inputs. This phenomenon can be attributed to the cooperative strategy transitioning to an LQR-dominated approach once the deviations converge within a sufficiently small range. Given COC’s excellent steady-state accuracy, only minimal control inputs are required to correct residual deviations. As a result, the thrust outputs under COC are significantly reduced compared to SMC and are even lower than those of LQR, demonstrating the efficiency of the proposed control framework.

After 1000 steps, the red lines exhibit slightly more pronounced oscillations, while the gray lines resume oscillating after another 200 steps. Although COC and SMC achieve comparable control accuracy under current perturbations, the former requires a longer response time. This delay exists because the deviations are too small to prompt a shift in COC’s cooperative strategy. As deviations accumulate, the COC strategy progressively shifts towards SMC, increasing thrust outputs to counteract the effects of current disturbance quickly. Leveraging the coordination mechanism, COC dynamically adjusts the control weights between LQR and SMC in response to the disturbance, thereby maintaining high control accuracy with less consumption than only SMC.

Keep focusing on the distinct thrust variation trends observed between steps 1400 and 1500 in Fig. 5g to elucidate the interaction between the coordination and deviation separation mechanisms. The enlarged view within Fig. 5g reveals that during steps 1440–1450, the gray line remains close to zero. However, for the rest of the period between steps 1400 and 1500, the gray line oscillates within a narrow range of  −3 N to 3 N. The former case indicates that COC stabilizes the UUV with minimal thrust, while the latter suggests that COC still requires significant thrust input to maintain system stability. This cyclical behavior occurs because when the sub-deviations caused by current disturbances are significantly suppressed through an SMC-dominated strategy, the COC coordination mechanism automatically transitions to an LQR-dominated mode. Then, there is a reduction in thrust. As a result, the influence of the thruster-noise sub-deviation diminishes until the coordination mechanism reverts to an SMC-dominated approach. This self-regulating collaborative mechanism mitigates the drawbacks of SMC’s tendency to switch among significant, opposing thrusts and LQR’s overly conservative thrust application. Consequently, COC consistently selects a more minor, smoother thrust profile while maintaining control accuracy, thereby reducing thruster wear and enhancing precision to achieve superior overall control performance.

Finally, we present a comparative analysis of the three control strategies’ overall performance in Fig. 6. For each segment, deviations and thrusts are first expressed in quadratic form and then summed to derive the control cost, which is represented as a bar chart. Given the substantial control inputs associated with SMC, the logarithm ({log }_{10}) of all control costs is applied to better visualize differences across orders of magnitude. As depicted in Fig. 6, the control cost of COC in the initial two segments exceeds that of LQR but remains lower than that of SMC. This is attributed to COC’s strategy of stabilizing the UUV through high-frequency thrust adjustments with lower amplitude than SMC. In subsequent segments, where deviations have significantly converged after 1000 steps, the gray bars consistently remain the lowest, indicating that COC achieves minimal overall control costs. These findings further validate that COC attains superior control accuracy with reduced control effort through its collaborative control strategy.

Fig. 6: Histogram of control costs for different controllers under current disturbance.

The control process is discretized into ten segments, each comprising 200 time steps. The red, gray, and blue cuboids represent the logarithm of each controller’s cost function.

Full size image

In summary, the above subsections analyze the control performance of LQR, SMC and COC controllers under the influence of current disturbance. With the comparisons between three control strategies, it is shown that the COC can obtain an optimal performance while balancing the need for robustness and consumption. Due to space limitations, the further exploration of the control performance in other scenarios is presented in the supplementary file from Sections S.4–S.6, including stable ocean current disturbance, increasing current disturbance, and load changing disturbance.

Conclusion

In this study, we proposed a novel COC framework that synergistically integrates SMC’s robustness and rapid responsiveness with the precision and low control cost of LQR. By developing a real-time error separation strategy, we successfully decoupled the challenges of disturbance rejection and precise hovering, enabling UUVs to operate efficiently in complex underwater environments. The proposed framework demonstrated remarkable adaptability to current disturbance without requiring prior knowledge of the disturbance characteristics. Simulation results revealed that the developed COC framework effectively mitigates the impact of current disturbance while maintaining precise hovering performance. Specifically, while the COC cost function remains optimal, the state response of position and direction under COC control remains fast and accurate. This highlights the ability of the proposed framework to achieve high-precision control with low energy consumption, making it particularly suitable for long-duration underwater missions. Furthermore, the COC framework demonstrated exceptional robustness in scenarios where other controllers failed, ensuring system stability and preventing catastrophic outcomes.

These findings underscore the potential of the COC framework to revolutionize underwater robotics by addressing the dual challenges of robustness and precision in dynamic marine environments. Future research will extend this framework to multi-agent systems and explore its applicability in more complex and unstructured underwater scenarios. This work represents a significant advancement in UUV control, offering a versatile and reliable solution for the growing demand for underwater exploration and operations.

Methods

Uncrewed underwater vehicles model

Consider a UUV system in the form of (For the details of the model settings, please see Section S.1 of the supplementary file.) :

$$left{begin{array}{l}[{{bf{M}}}({{bf{v}}})+Delta {{bf{M}}}({{bf{v}}})]dot{{{bf{v}}}}+[{{bf{C}}}({{bf{v}}})+Delta {{bf{C}}}({{bf{v}}})]{{bf{v}}}+[{{bf{D}}}({{bf{v}}})+Delta {{bf{D}}}({{bf{v}}})]{{bf{v}}}+[{{bf{g}}}({{boldsymbol{eta }}})+Delta {{bf{g}}}({{boldsymbol{eta }}})]={{{boldsymbol{tau }}}}_{delta }+{{{boldsymbol{delta }}}}_{c},quad \ dot{{{boldsymbol{eta }}}}={{bf{J}}}({{boldsymbol{eta }}}){{bf{v}}},hfill end{array}right.$$
(1)

where η = [xyzϕθψ]T represents the pose vector of the UUV in the inertial reference frame, and v = [uwvpqr]T denotes the velocity vector, which includes both linear and angular velocities in the body-fixed reference frame. The control input, influenced by the current disturbance δc, is expressed as ({{{boldsymbol{tau }}}}_{delta }={[{F}_{x},{F}_{y},{F}_{z},{M}_{phi },{M}_{theta },{M}_{psi }]}^{{{rm{T}}}}), representing the resultant force F and moments M along the x, y, and z axes, as well as the orientation angles ϕ, θ, and ψ. The matrix M(v) denotes the inertia matrix (which includes the added mass), C(v) refers to the matrix accounting for the centripetal force and the Coriolis forces, D(v) represents the hydrodynamic damping matrix, and g(η) is the vector of hydrostatic forces. In addition, ΔM(v), ΔC(v), ΔD(v) and Δg(η) represent the model uncertainties on the above matrixes generated from the loading changing or inaccurate modeling situations. The influence of these model uncertainties can be viewed as an external disturbance term as δm, from where the above model (1) can be simplified to:

$$begin{array}{l}left{begin{array}{l}{{bf{M}}}({{bf{v}}})dot{{{bf{v}}}}+{{bf{C}}}({{bf{v}}}){{bf{v}}}+{{bf{D}}}({{bf{v}}}){{bf{v}}}+{{bf{g}}}({{boldsymbol{eta }}})={{{boldsymbol{tau }}}}_{delta }+{{{boldsymbol{delta }}}}_{c}+{{{boldsymbol{delta }}}}_{m}={{{boldsymbol{tau }}}}_{delta }+{{boldsymbol{delta }}},quad \ dot{{{boldsymbol{eta }}}}={{bf{J}}}({{boldsymbol{eta }}}){{bf{v}}},hfill end{array}right.\ {{{boldsymbol{delta }}}}_{m}=-Delta {{bf{M}}}({{bf{v}}})dot{{{bf{v}}}}-Delta {{bf{C}}}({{bf{v}}}){{bf{v}}}-Delta {{bf{D}}}({{bf{v}}}){{bf{v}}}-Delta {{bf{g}}}({{boldsymbol{eta }}}),hfillend{array}$$
(2)

where δ is the integrated disturbance with δ = δc + δm. The model presented in equation (2) is referred to as the continuous model in this paper, which provides an intuitive representation of the relationship between thrust, pose, and velocity. Table 1 introduces the common mathematical expressions used in this article.

Table 1 The general mathematical expression
Full size table

Given the widespread adoption of digital controllers, designing controllers directly in the discrete-time domain effectively mitigates the delay issues introduced by sampling devices, such as zero-order hold elements, thereby ensuring the stability of the control system. The discrete-time form of the model (2) is derived to accommodate the dynamics of digital systems more effectively:

$$left{begin{array}{l}{{{bf{v}}}}_{k+1}=-{{{bf{M}}}{({{{bf{v}}}}_{k})}^{-1}[{{bf{C}}}({{{bf{v}}}}_{k}){{{bf{v}}}}_{k}+{{bf{D}}}({{{bf{v}}}}_{k}){{{bf{v}}}}_{k}+{{bf{g}}}({{{boldsymbol{eta }}}}_{k})]+{{bf{M}}}{({{{bf{v}}}}_{k})}^{-1}({{{boldsymbol{tau }}}}_{{{boldsymbol{delta }}},k}+{{{boldsymbol{delta }}}}_{k})}dt+{{{bf{v}}}}_{k},quad \ {{{boldsymbol{eta }}}}_{k+1}=[{{bf{J}}}({{{boldsymbol{eta }}}}_{k}){{{bf{v}}}}_{k}]dt+{{{boldsymbol{eta }}}}_{k},hfill end{array}right.$$
(3)

where dt represents the sampling time, determined by the hardware requirements, and k denotes the discrete time index. vk, vk+1, and ηk, ηk+1 represent the velocity vector v and the orientation vector η at time steps k and k + 1, respectively. τδ,k and δk denote the control input and the current disturbance at time step k. The matrices M(vk), C(vk), D(vk), g(ηk), and J(vk) retain the same physical interpretation as in the continuous model (2). Still, they represent the discrete-time model’s corresponding functions of vk and ηk.

In practical applications, the hydrostatic force g(ηk) can be considered negligible and effectively treated as zero. To simplify the controller design, a Taylor expansion at the time step k = 0 is applied to the model above, thereby establishing a linear relationship between the vehicle’s pose and the forces and moments generated by the thrusters:

$${{{bf{x}}}}_{{{boldsymbol{delta }}},k+1}={{{bf{A}}}}_{k}{{{bf{x}}}}_{{{boldsymbol{delta }}},k}+{{{bf{B}}}}_{k}{{{boldsymbol{tau }}}}_{{{boldsymbol{delta }}},k}+{{{boldsymbol{delta }}}}_{k},$$
(4)

where ({{{bf{x}}}}_{{{boldsymbol{delta }}},k}={[{{{boldsymbol{eta }}}}_{k}-{{{boldsymbol{eta }}}}_{r,k},{{{bf{v}}}}_{k}-{{{bf{v}}}}_{r,k}]}^{{{rm{T}}}}) represents the deviation between the desired pose ηr,k and velocity vr,k, and the current pose and velocity under the influence of current disturbance only, which constitutes the system state. The following system matrix Ak and input matrix Bk are defined as:

$$begin{array}{rcl}{{{bf{A}}}}_{k}&=&left[begin{array}{cc}left(frac{partial {{bf{J}}}({{{boldsymbol{eta }}}}_{{{bf{0}}}}){{{bf{v}}}}_{0}}{partial {{{boldsymbol{eta }}}}_{{{bf{0}}}}}+{{bf{I}}}right)dt&{{bf{J}}}({{{boldsymbol{eta }}}}_{0})dt\ -frac{partial {{bf{M}}}{({{{bf{v}}}}_{{{bf{0}}}})}^{-1}{{bf{g}}}({{{boldsymbol{eta }}}}_{0})}{partial {{{boldsymbol{eta }}}}_{{{bf{0}}}}}dt&left[{{bf{I}}}-frac{partial {{bf{H}}}({{{bf{v}}}}_{{{bf{0}}}},{{{boldsymbol{tau }}}}_{{{bf{0}}}})}{partial {{{bf{v}}}}_{0}}right]dtend{array}right],{{{bf{B}}}}_{k}={left[begin{array}{cc}0&frac{partial {{bf{M}}}{({{{bf{v}}}}_{{{bf{0}}}})}^{-1}{{{boldsymbol{tau }}}}_{0}}{partial {{{boldsymbol{tau }}}}_{{{bf{0}}}}}dtend{array}right]}^{{{rm{T}}}},\ {{bf{H}}}({{{bf{v}}}}_{{{bf{0}}}},{{{boldsymbol{tau }}}}_{{{bf{0}}}})&=&{{bf{M}}}{({{{bf{v}}}}_{0})}^{-1}{{bf{C}}}({{{bf{v}}}}_{0}){{{bf{v}}}}_{0}+{{bf{M}}}{({{{bf{v}}}}_{0})}^{-1}{{bf{D}}}({{{bf{v}}}}_{0}){{{bf{v}}}}_{0}+{{bf{M}}}{({{{bf{v}}}}_{0})}^{-1}{{bf{g}}}({{{boldsymbol{eta }}}}_{0})-{{bf{M}}}{({{{bf{v}}}}_{0})}^{-1}{{{boldsymbol{tau }}}}_{0}.end{array}$$
(5)

η0, v0, and τ0 represent the values of η, v, and τδ at the time step k = 0, respectively. I is the identity matrix with the same dimensions as J(ηk). H(v0τ0) is an intermediate variable introduced to simplify the notation.

A COC is designed to ensure stability and high precision during UUV hovering under the influence of currents for the UUV system described above. As previously noted, the controller manages hovering and disturbance rejection functions. To illustrate the coupling between these two objectives, we present the mismatched UUV model, derived from the discrete model (3), as follows:

$${{{bf{x}}}}_{k+1}={{{bf{A}}}}_{k}{{{bf{x}}}}_{k}+{{{bf{B}}}}_{k}{{{boldsymbol{tau }}}}_{k}+{{{boldsymbol{delta }}}}_{k}+{{{bf{w}}}}_{k},$$
(6)

where xk and τk represent the system state and thrust input of the mismatched model, respectively. Their physical meanings are identical to those of xδ,k and τδ,k. The only distinction is that, in contrast to model (4), random noise ({{{bf{w}}}}_{k}in {{mathscr{N}}}(0,{{{bf{W}}}}_{k})) from the thrusters is introduced. The remaining terms are consistent with those in the model (4). The model presented in equation (6) represents the actual model in this paper, and the subsequent derivations are based on this model.

The second-order linear model adopted in this study is widely used in the control design of underwater vehicles, mainly when the vehicle operates within a limited envelope under hovering or low-speed conditions. This modeling assumption is well-suited for streamlined, low-damping UUVs with symmetric hull structures and 6 degrees of freedom (6-DOF) control realized through vector thrusters or distributed actuators. As supported by prior studies35,36, such reduced-order models preserve the dominant dynamics relevant to pose regulation and disturbance rejection while greatly simplifying controller design and real-time implementation. However, we acknowledge that more sophisticated nonlinear models may be required for UUVs performing high-speed cruising, exhibiting complex hydrodynamic interactions (e.g., crossflow effects), or undergoing large-angle maneuvers. In such cases, the proposed COC framework can be extended by incorporating nonlinear control strategies (e.g., gain scheduling, feedback linearization), as long as the core deviation separation mechanism is maintained.

In this paper, we assume the hydrostatic force g(ηk) is negligible and can be treated as zero. This assumption is valid for typical UUVs with slightly positive buoyancy operating at shallow to moderate depths, where the effect of hydrostatic restoring forces is minimal and can be safely ignored for pose control purposes. However, this assumption may no longer hold for deep-sea operations or missions involving heavy payloads due to significant variations in buoyancy and restoring torques. The term g(ηk) should be retained in the dynamic model. Nevertheless, doing so does not affect the structural validity of the proposed COC framework. The additional terms can be incorporated into the LQR and SMC components through model-based feed-forward compensation, thus compensating without affecting the cooperative framework.

Cooperative control framework for UUV

The previous section introduced the mismatched model, and the thrust required to stabilize the UUV system was calculated. Relying solely on SMC to address the coupled tasks would lead to inefficient utilization of limited power resources. In contrast, LQR alone is insufficient for handling current disturbances37. Therefore, a combined control strategy is required to exploit the strengths of both approaches.

However, to enable each sub-controller to perform its respective task, it is necessary to first decouple the mismatched model into two distinct models, each representing the hovering task and the disturbance rejection task, respectively. The fundamental framework of COC relies on feedback control, with the only available feedback being the deviation between the current pose and the desired one. Therefore, a deviation separation mechanism is introduced to partition the overall deviation into sub-deviations, each corresponding to the specific characteristics of the respective tasks. Mathematically, this analysis can be expressed as follows:

$${{{bf{x}}}}_{k+1}={{{bf{A}}}}_{k}{{{bf{x}}}}_{{{boldsymbol{delta }}},k}+{{{bf{B}}}}_{k}{{{boldsymbol{tau }}}}_{{{boldsymbol{delta }}},k}+{{{boldsymbol{delta }}}}_{k}+{{{bf{A}}}}_{k}{{{bf{x}}}}_{{{boldsymbol{varepsilon }}},k}\ +{{{bf{B}}}}_{k}{{{boldsymbol{tau }}}}_{{{boldsymbol{varepsilon }}},k}+{{{bf{w}}}}_{k}={{{bf{x}}}}_{{{boldsymbol{delta }}},k+1}+{{{bf{x}}}}_{{{boldsymbol{varepsilon }}},k+1},$$
(7)

where xε,k is the vector representing the separated sub-deviation from thruster noise, while xδ,k is the sub-deviation from current disturbance as introduced earlier. These two sub-deviations are related by the equation xk = xδ,k + xε,k. Specifically, xδ,k corresponds to the sub-deviation associated with the anti-disturbance task, whereas xε,k represents the sub-deviation related to the hovering task.

Recent researches show that the linear quadratic problem can be reformulated as an optimization problem involving probability distributions, which transforms the task of calculating optimal thrust into maximizing the posterior probability distribution of the deviation state xk38,39. Building on these findings, the effect of minor disturbances on the posterior distribution can be estimated by analyzing the influence function. Similarly, the anti-disturbance task can be viewed as a minor disturbance influencing the hovering task. By examining the influence function of the LQR, we can quantify the impact of the anti-disturbance task on the hovering task under the thrust calculated by the LQR. In this context, it corresponds to the influence of the current disturbance on the hovering task. Subsequently, the estimated sub-deviation caused by the current disturbance can be computed by inverting the influence function. The difference between the current deviation and the calculated sub-deviation represents the estimated sub-deviation of the hovering task, which forms the foundation for the separation mechanism.

Mathematically, the optimal thrust in LQR is computed by solving an optimization problem subject to the constraints of the nominal model, as described in Eq. (4). When current disturbances affect the UUVs, the dynamic model transitions from the nominal model (4) to the mismatched model (6). As a result, the optimization problem is modified as follows:

$$begin{array}{l}left{begin{array}{l}{J}_{k}^{star }={min }_{{{{boldsymbol{tau }}}}_{k}}left[({{{{bf{x}}}}_{{{boldsymbol{varepsilon }}},k}}^{{{rm{T}}}}{{bf{Q}}}{{{bf{x}}}}_{{{boldsymbol{varepsilon }}},k}+{{{{boldsymbol{tau }}}}_{k}}^{{{rm{T}}}}{{bf{R}}}{{{boldsymbol{tau }}}}_{k})+{{{{bf{x}}}}_{k+1}}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{x}}}}_{k+1}right],quad \ {{{boldsymbol{tau }}}}_{k}^{star }={min }_{{{{boldsymbol{tau }}}}_{k}}{{J}_{{{boldsymbol{varepsilon }}},k}+[{{{{bf{x}}}}_{k+1}}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{x}}}}_{k+1}-{{{{bf{x}}}}_{{{boldsymbol{varepsilon }}},k+1}}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{x}}}}_{{{boldsymbol{varepsilon }}},k+1}]}.quad end{array}right.\ {J}_{k}={{{{bf{x}}}}_{k+1}}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{x}}}}_{k+1},{J}_{{{boldsymbol{varepsilon }}},k}={{{{bf{x}}}}_{{{boldsymbol{varepsilon }}},k+1}}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{x}}}}_{{{boldsymbol{varepsilon }}},k+1},end{array}$$
(8)

where Q and R are two weighting matrices defined artificially, Jk and Jε,k represent the cost functions of the COC and LQR, respectively. ({{{boldsymbol{tau }}}}_{k}^{star }) is the optimal control thrust. ({J}_{k}^{star }) is the optimal cost function of the COC, minimized by ({{{boldsymbol{tau }}}}_{k}={{{boldsymbol{tau }}}}_{k}^{star }), which can also be expressed as ({J}_{k}^{star }={min }_{{{{boldsymbol{tau }}}}_{k}}{J}_{k}). Pk+1 is a symmetric matrix obtained through the Riccati equation (The detailed derivation process is shown in Section S.3 of the supplementary file). The bias between ({{{boldsymbol{tau }}}}_{k}^{star }) and τε,k is defined as ({tilde{{{boldsymbol{tau }}}}}_{k}). According to the maximum principle, the first derivative of Eq. (8) satisfies:

$${left.frac{partial {{J}_{{{boldsymbol{varepsilon }}},k}+[{{{{bf{x}}}}_{k+1}}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{x}}}}_{k+1}-{{{{bf{x}}}}_{{{boldsymbol{varepsilon }}},k+1}}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{x}}}}_{{{boldsymbol{varepsilon }}},k+1}]}}{partial {{{boldsymbol{tau }}}}_{k}}rightvert }_{{{{boldsymbol{tau }}}}_{k} = {{{boldsymbol{tau }}}}_{k}^{star }}=0.$$
(9)

For simplicity, let (f({{{boldsymbol{tau }}}}_{k}^{star })) denote the left-hand side of Eq. (9). By performing a second-order Taylor expansion of Eq. (9), we obtain:

$$ f({{{boldsymbol{tau }}}}_{{{boldsymbol{varepsilon }}},k})+frac{partial f({{{boldsymbol{tau }}}}_{k})}{partial {{{boldsymbol{tau }}}}_{k}}({{{boldsymbol{tau }}}}_{k}-{{{boldsymbol{tau }}}}_{{{boldsymbol{varepsilon }}},k})+{{mathscr{O}}}{({{{boldsymbol{tau }}}}_{k}-{{{boldsymbol{tau }}}}_{{{boldsymbol{varepsilon }}},k})}^{2}\ approx {left.left{frac{partial {J}_{{{boldsymbol{varepsilon }}},k}}{partial {{{boldsymbol{tau }}}}_{k}}+frac{partial ({J}_{k}-{J}_{{{boldsymbol{varepsilon }}},k})}{partial {{{boldsymbol{tau }}}}_{k}}right}right| }_{{{{boldsymbol{tau }}}}_{k} = {{{boldsymbol{tau }}}}_{{{boldsymbol{varepsilon }}},k}}+left{frac{{partial }^{2}[{J}_{{{boldsymbol{varepsilon }}},k}+{J}_{k}-{J}_{{{boldsymbol{varepsilon }}},k}]}{partial {{{{boldsymbol{tau }}}}_{k}}^{2}}right}({{{boldsymbol{tau }}}}_{k}-{{{boldsymbol{tau }}}}_{{{boldsymbol{varepsilon }}},k})$$
(10)

Substituting Eq. (9) into the above equations and noting that τε,k minimizes Jε,k, it follows that the partial derivative (frac{partial {J}_{{{boldsymbol{varepsilon }}},k}}{partial {{{boldsymbol{tau }}}}_{k}}{| }_{{{{boldsymbol{tau }}}}_{k} = {{{boldsymbol{tau }}}}_{{{boldsymbol{varepsilon }}},k}}) equals to zero. By substituting ({tilde{{{boldsymbol{tau }}}}}_{k}) for the corresponding terms in Eq. (10) and neglecting higher-order terms, Eq. (10) simplifies to:

$${left.frac{partial ({J}_{k}-{J}_{{{boldsymbol{varepsilon }}},k})}{partial {{{boldsymbol{tau }}}}_{k}}+left(frac{{partial }^{2}{J}_{k}}{partial {{{{boldsymbol{tau }}}}_{k}}^{2}}right){tilde{{{boldsymbol{tau }}}}}_{k}right| }_{{{{boldsymbol{tau }}}}_{k} = {{{boldsymbol{tau }}}}_{{{boldsymbol{varepsilon }}},k}}to {tilde{{{boldsymbol{tau }}}}}_{k}={left(frac{{partial }^{2}{J}_{k}}{partial {{{{boldsymbol{tau }}}}_{k}}^{2}}right)}^{-1}{left.frac{partial ({J}_{k}-{J}_{{{boldsymbol{varepsilon }}},k})}{partial {{{boldsymbol{tau }}}}_{k}}rightvert }_{{{{boldsymbol{tau }}}}_{k} = {{{boldsymbol{tau }}}}_{{{boldsymbol{varepsilon }}},k}}.$$
(11)

By substituting the cost function expression (8) into the above formula (11), the final expression for ({tilde{{{boldsymbol{tau }}}}}_{k}) is obtained as:

$${tilde{{{boldsymbol{tau }}}}}_{k}=-{({{bf{R}}}+{{{bf{B}}}}_{k}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{B}}}}_{k})}^{-1}{{{bf{B}}}}_{k}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{x}}}}_{{{boldsymbol{delta }}},k}.$$
(12)

Since τε,k is independent of xδ,k, the first-order partial derivative of ({tilde{{{boldsymbol{tau }}}}}_{k}) with respect to xδ,k can be expressed as:

$$frac{partial {tilde{{{boldsymbol{tau }}}}}_{k}}{partial {{{bf{x}}}}_{{{boldsymbol{delta }}},k}}=frac{partial ({{{boldsymbol{tau }}}}_{k}^{star }-{{{boldsymbol{tau }}}}_{{{boldsymbol{varepsilon }}},k})}{partial {{{bf{x}}}}_{{{boldsymbol{delta }}},k}}=frac{partial {{{boldsymbol{tau }}}}_{k}^{star }}{partial {{{bf{x}}}}_{{{boldsymbol{delta }}},k}}.$$
(13)

As mentioned earlier, xδ,k, which represents the sub-deviation of the anti-disturbance task, influences the cost function. This influence can be quantified by performing a partial derivative. Thus, to calculate the change in the cost function resulting from the sub-deviation xδ,k, we take the derivative of Jk along the xδ,k direction. The partial derivative mentioned above is denoted as Lk, and its expression is:

$${{{bf{L}}}}_{k}= frac{partial {J}_{k}}{partial {{{bf{x}}}}_{{{boldsymbol{delta }}},k}}={left.frac{partial {J}_{k}}{partial {{{boldsymbol{tau }}}}_{k}}frac{partial {{{boldsymbol{tau }}}}_{k}^{star }}{partial {{{bf{x}}}}_{{{boldsymbol{delta }}},k}}right| }_{{{{boldsymbol{tau }}}}_{ = }{{{boldsymbol{tau }}}}_{{{boldsymbol{varepsilon }}},k}}\ = -2{({{bf{R}}}{{{boldsymbol{tau }}}}_{k}+{{{bf{B}}}}_{k}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{x}}}}_{k+1})}^{{{rm{T}}}}{({{bf{R}}}+{{{bf{B}}}}_{k}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{B}}}}_{k})}^{-1}{{{bf{B}}}}_{k}^{{{rm{T}}}}{{{bf{P}}}}_{k+1},$$
(14)

Thus, we establish the relationship between the partial derivative and the sub-deviation we aim to estimate. By performing an inverse integration on the partial derivative, the expression for the sub-deviation can be written as ({{{bf{x}}}}_{{{boldsymbol{delta }}},k}={{{bf{L}}}}_{k}^{-1}({J}_{k}-{{boldsymbol{beta }}})), where ({{boldsymbol{beta }}}=min {J}_{k}| forall ({{{bf{x}}}}_{k},{{{boldsymbol{tau }}}}_{k})=0). The term β is close to zero due to the optimality of LQR, and since ({{{bf{L}}}}_{k}^{-1}) is bounded, the expression for xδ,k is approximated as ({{{bf{L}}}}_{k}^{-1}{J}_{k}). In the ocean environment, where current changes are slow, the UUV control problem can be modeled as a system with a fast control frequency and slow time-invariant disturbances. Therefore, we assume that xδ,m ≈ xδ,m+1 ≈  ≈ xδ,k, where m = k − M2 + 1 is a different time step and M2 represents the dimension of the sub-deviation vector xδ,k. Consequently, the full expression for xδ,k is given by ({{{bf{x}}}}_{{{boldsymbol{delta }}},k}={{{bf{J}}}}_{m,k}{{{bf{L}}}}_{m,k}^{-1}), where Jm,k = [JmJm+1, …, Jk] and Lm,k = [LmLm+1, …, Lk] are two augmented matrices of Jk and Lk, respectively.

In the previous subsections, the sub-deviation of the anti-disturbance task was separated from the total deviation. To express the COC more concisely in subsequent sections, we define the separation matrix Φk such that ({{{bf{x}}}}_{{{boldsymbol{delta }}},k}={{{mathbf{Phi }}}}_{k}{{{bf{x}}}}_{k}={{{bf{J}}}}_{m,k}{{{bf{L}}}}_{m,k}^{-1}). The analytical expression for Φk is derived as follows:

$$left{begin{array}{l}{{{mathbf{Phi }}}}_{k}=diag({{{bf{J}}}}_{m,k}{{{{bf{L}}}}_{m,k}}^{-1})./{{{bf{x}}}}_{k},quad \ {{{bf{x}}}}_{{{boldsymbol{delta }}},k}={{{mathbf{Phi }}}}_{k}{{{bf{x}}}}_{k},hfill \ {{{bf{x}}}}_{{{boldsymbol{varepsilon }}},k}=({{bf{I}}}-{{{mathbf{Phi }}}}_{k}){{{bf{x}}}}_{k},hfill end{array}right.$$
(15)

where ./ denotes element-wise division of the corresponding elements, and diag(…) represents a matrix with the vectors in (…) as its diagonal elements. If the system is stable at the desired pose or has little current disturbance, no sub-deviation is generated from the anti-disturbance task. Consequently, the value of Lk becomes very small, and its inverse grows exceedingly large. In order to avoid numerical instability when the matrix Lk is tiny, a threshold control strategy is introduced when constructing the separation matrix Φk. Specifically, a safety bound is set for the augmented vector Lm,k as Lm,k ≤ ε, where Lm,k is the determinant of Lm,k and ε is an artificial defined scalar. When the strategy is satisfied, the deviation caused by the disturbance is almost negligible, and the LQR controller is sufficient to stabilize the system without the need for the SMC compensation term, so the separation matrix Φk is set to zero.

Since the sub-deviation has been obtained, the decoupled sub-deviations are introduced into the LQR and SMC controllers separately. For the SMC sub-controller, the thrust is calculated as follows:

$$left{begin{array}{l}{{{bf{s}}}}_{{{boldsymbol{delta }}},k}={{bf{c}}}{{{bf{v}}}}_{{{boldsymbol{delta }}},k}+{{{boldsymbol{eta }}}}_{{{boldsymbol{delta }}},k},hfill \ {tau }_{{{boldsymbol{delta }}},k}={{bf{C}}}({{{bf{v}}}}_{k}){{{bf{v}}}}_{{{boldsymbol{delta }}},k}+{{bf{D}}}({{{bf{v}}}}_{k}){{{bf{v}}}}_{{{boldsymbol{delta }}},k}+{{bf{g}}}({{{boldsymbol{eta }}}}_{k})-{{bf{M}}}({{{bf{v}}}}_{k}){{{bf{c}}}}^{-1}{{bf{J}}}({{{boldsymbol{eta }}}}_{k}){{{bf{v}}}}_{{{boldsymbol{delta }}},k}-{{bf{M}}}({{{bf{v}}}}_{k}){{{bf{c}}}}^{-1}rho sgn({{{bf{s}}}}_{{{boldsymbol{delta }}},k})quad end{array}right.$$
(16)

where ηδ,k = xδ,k(1: 6) and vδ,k = xδ,k(7: 12) are the pose vector and velocity vector of the sub-deviation, respectively. sδ,k represents the sliding mode surface, and c is the sliding mode gain, both of which are similar to the SMC. ρ is the switching gain, and sgn is the sign function, with their respective expressions are available in Section S.2 of the supplementary file. From a mechanical perspective, the thrust τδ,k can be viewed as a combination of two components. The second part,  −M(vk)c−1ρsgn(sδ,k), is a discontinuous switching term that drives the UUV’s dynamics toward the sliding surface sδ,k. This sliding surface is inherently stabilized by the feedforward term C(vk)vδ,k + D(vk)vδ,k + g(ηk) − M(vk)c−1J(ηk)vδ,k, which ensures the system converges to the desired pose without requiring additional thrust.

For the LQR sub-controller, its thrust is calculated by:

$${{{boldsymbol{tau }}}}_{{{boldsymbol{varepsilon }}},k}={{{bf{K}}}}_{varepsilon }{{{bf{x}}}}_{{{boldsymbol{varepsilon }}},k}={{{bf{K}}}}_{varepsilon }({{bf{I}}}-{{{mathbf{Phi }}}}_{k}){{{bf{x}}}}_{k},$$
(17)

where Kε is the control gain of LQR. It follows the same feedback control strategy as the traditional LQR from Section S.3 of the supplementary file, with the only difference being that the deviation here corresponds to the sub-deviation of the hovering task. By combining the two control inputs, the thrust of the COC is:

$${{{boldsymbol{tau }}}}_{k}={tau }_{delta ,k}+{{{boldsymbol{tau }}}}_{{{boldsymbol{varepsilon }}},k}.$$
(18)

From the thrust of COC in Eq. (18), it can be observed that the thrust is a combination of the thrusts from the two sub-controllers, including the LQR and SMC. By integrating both a feedforward term and a compensated switching term, the proposed COC strategy introduces SMC’s robust disturbance suppression capabilities into the COC framework from two key perspectives. First, the feedforward term ensures that the system reaches the SMC’s sliding surface within a finite time. Second, the switching term guarantees the inherent stability of the sliding surface, thereby enhancing the system’s resilience to uncertainties and disturbances. Coupled with the deviation separation mechanism, this collaborative framework allows the proposed COC approach to effectively balance robustness, precision, and energy efficiency in dynamic and uncertain environments. The overall framework is listed as Algorithm 1.

Algorithm 1

The framework of the cooperative controller

Require: vk, ηk, M(vk), C(vk), D(vk), g(ηk), Q, R, c, ρ, dt

Ensure: τk

1: ({{bf{H}}}({{{bf{v}}}}_{{{bf{0}}}},{{{boldsymbol{tau }}}}_{{{bf{0}}}})={{bf{M}}}{({{{bf{v}}}}_{0})}^{-1}{{bf{C}}}({{{bf{v}}}}_{0}){{{bf{v}}}}_{0}+{{bf{M}}}{({{{bf{v}}}}_{0})}^{-1}{{bf{D}}}({{{bf{v}}}}_{0}){{{bf{v}}}}_{0}+{{bf{M}}}{({{{bf{v}}}}_{0})}^{-1}{{bf{g}}}({{{boldsymbol{eta }}}}_{0})-{{bf{M}}}{({{{bf{v}}}}_{0})}^{-1}{{{boldsymbol{tau }}}}_{0}),               initialization process

2: ({{{bf{A}}}}_{k}=left[begin{array}{c}(frac{partial {{bf{J}}}({{{boldsymbol{eta }}}}_{{{bf{0}}}}){{{bf{v}}}}_{0}}{partial {{{boldsymbol{eta }}}}_{{{bf{0}}}}}+{{bf{I}}})dt{{bf{J}}}({{{boldsymbol{eta }}}}_{0})dt\ -frac{partial {{bf{M}}}{({{{bf{v}}}}_{{{bf{0}}}})}^{-1}{{bf{g}}}({{{boldsymbol{eta }}}}_{0})}{partial {{{boldsymbol{eta }}}}_{{{bf{0}}}}}dt[{{bf{I}}}-frac{partial {{bf{H}}}({{{bf{v}}}}_{{{bf{0}}}},{{{boldsymbol{tau }}}}_{{{bf{0}}}})}{partial {{{bf{v}}}}_{0}}]dtend{array}right],{{{bf{B}}}}_{k}={left[0frac{partial {{bf{M}}}{({{{bf{v}}}}_{{{bf{0}}}})}^{-1}{{{boldsymbol{tau }}}}_{0}}{partial {{{boldsymbol{tau }}}}_{{{bf{0}}}}}dtright]}^{{{rm{T}}}}),

3: ({{{bf{P}}}}_{k}={{{bf{A}}}}_{k}^{T}{{{bf{P}}}}_{k+1}{{{bf{A}}}}_{k}-{{{bf{A}}}}_{k}^{T}{{{bf{P}}}}_{k+1}{{{bf{B}}}}_{k}{({{bf{R}}}+{{{bf{B}}}}_{k}^{T}{{{bf{P}}}}_{k+1}{{{bf{B}}}}_{k})}^{-1}{{{bf{B}}}}_{k}^{T}{{{bf{P}}}}_{k+1}{{{bf{A}}}}_{k}+{{bf{Q}}}quad triangleright) This equation is convergent. Considering the infinite time domain, then Pk = Pk+1 = P.

4: ({{{bf{K}}}}_{varepsilon }=-{({{bf{R}}}+{{{bf{B}}}}_{k}^{T}{{bf{P}}}{{{bf{B}}}}_{k})}^{-1}{{{bf{B}}}}_{k}^{T}{{bf{P}}}{{{bf{A}}}}_{k})

5: loop                                                  main loop

6:  ({{{bf{L}}}}_{k}=-2{({{bf{R}}}{{{boldsymbol{tau }}}}_{k}+{{{bf{B}}}}_{k}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{x}}}}_{k+1})}^{{{rm{T}}}}{({{bf{R}}}+{{{bf{B}}}}_{k}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{B}}}}_{k})}^{-1}{{{bf{B}}}}_{k}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}), ({J}_{k}=({{{{bf{x}}}}_{k}}^{{{rm{T}}}}{{bf{Q}}}{{{bf{x}}}}_{k}+{{{{boldsymbol{tau }}}}_{k}}^{{{rm{T}}}}{{bf{R}}}{{{boldsymbol{tau }}}}_{k})+{{{{bf{x}}}}_{k+1}}^{{{rm{T}}}}{{{bf{P}}}}_{k+1}{{{bf{x}}}}_{k+1})

7:  Jm,k = [Jk−5Jk−4, …, Jk], Lm,k = [Lk−5Lk−4, …, Lk]

8:  ({{{mathbf{Phi }}}}_{k}=diag({{{bf{J}}}}_{m,k}{{{{bf{L}}}}_{m,k}}^{-1})./{{{bf{x}}}}_{k}), xδ,k = Φkxk, xε,k = (I − Φk)xk

9:  τε,k = −Kεxε,k                                       LQR subcontroller

10: ηδ,k = xδ,k(1: 6), vδ,k = xδ,k(7: 12), sδ,k = cvδ,k + ηδ,k

11: τδ,k = C(vk)vδ,k + D(vk)vδ,k + g(ηk) − M(vk)c−1J(ηk)vδ,k − M(vk)c−1ρsgn(sδ,k)             SMC subcontroller

12: τk = τδ,k + τε,k                                     Cooperative controller

13: return τk

14: end loop

This algorithm also raises a natural question: since the final expression of the proposed controller is a weighted form, why not directly calculate the final thrust in a weighted manner? The reason is that simply combining these two control inputs does not guarantee that their advantages are maximized or the disadvantages mitigated. Using LQR alone causes improper handling of the current disturbance, leading to divergence of the control system. On the other hand, using SMC only results in overreaction to minor deviations caused by thruster noise, leading to excessive control power consumption and potentially causing oscillations in the control system.

Thus, as depicted in Fig. 7, instead of directly combining the two controllers as done previously, we propose decomposing the deviation into two parts, each tailored to LQR and SMC. The proposed deviation separation algorithm divides the mixed deviation into two distinct parts. LQR is then used to address the sub-deviation caused by thruster noise, while SMC handles the sub-deviation arising from the current disturbance. Compared to the mixed controller, the COC allows the two controllers to complement each other’s strengths, leading to improved performance.

Fig. 7: The basic framework of COC.

The COC optimizes performance by separating deviations and coordinating LQR (for thruster noise) and SMC (for current disturbances), leveraging their complementary strengths.

Full size image

Data availability

The data supporting this funding are available at the Supplementary Data 1 or through https://github.com/DanielGrowl/Co-controller-for-UUV.git.

Code availability

The code for replicating our results is available at https://github.com/DanielGrowl/Co-controller-for-UUV.git.

References

  1. Sahoo, A., Dwivedy, S. K. & Robi, P. Advancements in the field of autonomous underwater vehicle. Ocean Eng. 181, 145 (2019).

    Article 

    Google Scholar 

  2. Tijjani, A. S., Chemori, A. & Creuze, V. A survey on tracking control of unmanned underwater vehicles: experiments-based approach. Annu. Rev. Control 54, 125 (2022).

    Article 
    MathSciNet 

    Google Scholar 

  3. Liu, F., Cui, W. & Li, X. China’s first deep manned submersible, JIAOLONG. Sci. China Earth Sci. 53, 1407 (2010).

    Article 

    Google Scholar 

  4. Cui, W. Development of the Jiaolong deep manned submersible. Mar. Technol. Soc. J. 47, 37 (2013).

    Article 

    Google Scholar 

  5. Zhao, X. et al. Bayesian learning for the robust verification of autonomous robots. Commun. Eng. 3, 18 (2024).

    Article 

    Google Scholar 

  6. Liu, K. et al. A maneuverable underwater vehicle for near-seabed observation. Nat. Commun. 15, 10284 (2024).

    Article 

    Google Scholar 

  7. Marra, G. et al. Optical interferometry–based array of seafloor environmental sensors using a transoceanic submarine cable. Science 376, 874 (2022).

    Article 

    Google Scholar 

  8. Huertas-Cerdeira, C. & Gharib, M. A 3-DOF caudal fin for precise maneuvering of thunniform-inspired unmanned underwater vehicles. Sci. Rep. 14, 17000 (2024).

    Article 

    Google Scholar 

  9. Qi, Y. et al. AirFormer: learning-based object detection for Mars helicopter. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 18, 100–111 (2025).

    Article 

    Google Scholar 

  10. Zhao, H. et al. M2CS: a multimodal and campus-scapes dataset for dynamic SLAM and moving object perception. J. Field Robot. 42, 787–805 (2025).

    Article 

    Google Scholar 

  11. Shamshad, N. et al. Advanced KNN-based cost-efficient algorithm for precision localization and energy optimization in dynamic underwater sensor networks. Sci. Rep. 15, 2182 (2025).

    Article 

    Google Scholar 

  12. Karimi, H. R. & Lu, Y. Guidance and control methodologies for marine vehicles: a survey. Control Eng. Pract. 111, 104785 (2021).

    Article 

    Google Scholar 

  13. Luo, Y., Yao, M. & Xiao, X. GCNT: Graph-Based Transformer Policies for Morphology-Agnostic Reinforcement Learning. In Proceedings of the Thirty-Fourth International Joint Conference on Artificial Intelligence (IJCAI-25), 8741–8749 (International Joint Conferences on Artificial Intelligence Organization, 2025).

  14. Tijjani, A. S., Chemori, A. & Roman, G. From non-model-based to adaptive model-based tracking control of low-inertia underwater vehicles. In Underwater vehicles: Design and applications, 35 (Nova Science Publishers, 2021).

  15. Mohd Tumari, M. Z. et al. PSO fine-tuned model-free PID controller with derivative filter for depth control of hovering autonomous underwater vehicle. In Proc. 10th National Technical Seminar on Underwater System Technology 2018: NUSYS’18, 3 (Springer, 2019).

  16. Hou, Y. et al. Velocity and trajectory tracking control model for underactuated UUVs through coupling of direct CFD and PID control algorithm. Ocean Eng. 314, 119775 (2024).

    Article 

    Google Scholar 

  17. Dong, B., Shi, Y., Xie, W., Chen, W. & Zhang, W. Learning-based robust optimal tracking controller design for unmanned underwater vehicles with full-state and input constraints. Ocean Eng. 271, 113757 (2023).

    Article 

    Google Scholar 

  18. Wang, Z., Li, Y., Ma, C., Yan, X. & Jiang, D. Path-following optimal control of autonomous underwater vehicle based on deep reinforcement learning. Ocean Eng. 268, 113407 (2023).

    Article 

    Google Scholar 

  19. Gong, H., Er, M. J. & Liu, Y. Fuzzy optimal fault-tolerant trajectory tracking control of underactuated AUVs with prescribed performance in 3-d space. IEEE Trans. Syst. Man. Cybern. Syst. 55, 170 (2025).

    Article 

    Google Scholar 

  20. Chu, Z., Wang, F., Lei, T. & Luo, C. Path planning based on deep reinforcement learning for autonomous underwater vehicles under ocean current disturbance. IEEE Trans. Intell. Veh. 8, 108 (2022).

    Article 

    Google Scholar 

  21. Gong, P., Yan, Z., Zhang, W. & Tang, J. Lyapunov-based model predictive control trajectory tracking for an autonomous underwater vehicle with external disturbances. Ocean Eng. 232, 109010 (2021).

    Article 

    Google Scholar 

  22. Yu, X., Feng, Y. & Man, Z. Terminal sliding mode control–an overview. IEEE Open J. Ind. Electron. Soc. 2, 36 (2020).

    Article 

    Google Scholar 

  23. Sun, J., Yao, M., Xiao, X., Xie, Z. & Zheng, B. Co-optimization of morphology and behavior of modular robots via hierarchical deep reinforcement learning. In Robotics: Science and Systems (2023).

  24. Guerrero, J., Chemori, A., Torres, J. & Creuze, V. Time-delay high-order sliding mode control for trajectory tracking of autonomous underwater vehicles under disturbances. Ocean Eng. 268, 113375 (2023).

    Article 

    Google Scholar 

  25. Wahyuadnyana, K. D., Indriawati, K., Darwito, P. A., Aufa, A. N. & Tnunay, H. Performance analysis of PID and SMC control algorithms on AUV under the influence of internal solitary wave in the bali deep sea. J. Robot. Control 5, 1957 (2024).

    Article 

    Google Scholar 

  26. Liu, X., Zhang, M., Rogers, E., Wang, Y. & Yao, F. Terminal sliding mode-based tracking control with error transformation for underwater vehicles. Int. J. Robust. Nonlinear Control 31, 7186 (2021).

    Article 
    MathSciNet 

    Google Scholar 

  27. Tugal, H. et al. Sliding mode controller for positioning of an underwater vehicle subject to disturbances and time delays. In Proc. International Conference on Robotics and Automation (ICRA), 3034–3039 (IEEE, 2022).

  28. Manzanilla, A. et al. Super-twisting integral sliding mode control for trajectory tracking of an unmanned underwater vehicle. Ocean Eng. 234, 109164 (2021).

    Article 

    Google Scholar 

  29. Yan, Z., Yan, J., Nan, F., Cai, S. & Hou, S. Distributed tmpc formation trajectory tracking of multi-UUV with time-varying communication delay. Ocean Eng. 297, 117091 (2024).

    Article 

    Google Scholar 

  30. Huang, C., Shi, Q., Ding, W., Mei, P. & Karimi, H. R. A robust MPC approach for platooning control of automated vehicles with constraints on acceleration. Control Eng. Pract. 139, 105648 (2023).

    Article 

    Google Scholar 

  31. Li, Z., Zhang, W., Zhang, Y. & Wu, W. Coordinated control strategies for heterogeneous multi-UUV recovery system in low-visibility underwater implicit interaction scenarios. Ocean Eng. 297, 117076 (2024).

    Article 

    Google Scholar 

  32. Fu, B., Zhu, D., Chen, M. & Yang, S. X. Formation control of unmanned underwater vehicles with unknown absolute position using rigid graph-based mpc. IEEE Internet Things J. 12, 29061 (2025).

    Article 

    Google Scholar 

  33. Long, C., Hu, M., Bian, Y. & Chang, D. Sredftc: a safe, robust and efficient distributed formation tracking controller for nonlinear multi-UUV systems. Nonlinear Dyn. 1, 1 (2025).

    Google Scholar 

  34. Yang, X. et al. National geodatabase of ocean current power resource in USA. Renew. Sustain. Energy Rev. 44, 496 (2015).

    Article 

    Google Scholar 

  35. Hodgkinson, C. Wealth and happiness: an analysis and some implications for education. Can. J. Educ./Rev. Can. de l’edu. 7, 1–14 (1982).

  36. Mitchell, D. G. & Hoh, R. H. Low-order approaches to high-order systems: problems and promises. J. Guidance Control Dyn. 5, 482 (1982).

    Article 

    Google Scholar 

  37. Anderson, B. D. & Moore, J. B. Optimal Control: Linear Quadratic Methods (Courier Corporation, 2007).

  38. Levine, S. Reinforcement learning and control as probabilistic inference: tutorial and review. arXiv preprint arXiv:1805.00909 (2018).

  39. Cook, R. D. Assessment of local influence. J. R. Stat. Soc. Ser. B Stat. Methodol. 48, 133 (1986).

    Article 
    MathSciNet 

    Google Scholar 

Download references

Acknowledgements

This work was supported in part by the National Natural Science Foundation of China under Grant 62403215, the Natural Science Foundation of Jiangsu Province (BK20241607), the Fundamental Research Funds for the Central Universities China (JUSRP123064) and the Postgraduate Research Practice Innovation Program of Jiangsu Province KYCX25-2661.

Author information

Authors and Affiliations

Authors

Contributions

Xiaoli Luan contributed to the study’s conceptualization and methodology. Shenhan Yu drafted the manuscript, developed the software, and created visualizations. Haiying Wan conducted the investigation, performed validation, and contributed to the writing process. Fei Liu conducted some coordination and optimization work.

Corresponding author

Correspondence to
Haiying Wan.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Communications Engineering thanks Meibao Yao and the other, anonymous, reviewers for their contribution to the peer review of this work. Primary Handling Editors: [Zhuo Zhang] and [Rosamund Daw].

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Supplementary information

Description of Additional Supplementary Files

Supplementary Data 1

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.

Reprints and permissions

About this article

Cite this article

Luan, X., Yu, S., Wan, H. et al. Dynamic hovering for uncrewed underwater vehicles via an error-separation-based cooperative strategy.
Commun Eng 4, 193 (2025). https://doi.org/10.1038/s44172-025-00528-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Version of record:

  • DOI: https://doi.org/10.1038/s44172-025-00528-w


Source: Resources - nature.com

Scalable brine treatment using 3D-printed multichannel thermodiffusion

Publisher Correction: Predicting water scarcity in northern Bangladesh using deep learning and climate data

Back to Top