Journal Archive

Journal of Advanced Marine Engineering and Technology - Vol. 44 , No. 4

[ Original Paper ]
Journal of Advanced Marine Engineering and Technology - Vol. 44, No. 4, pp. 311-317
Abbreviation: JAMET
ISSN: 2234-7925 (Print) 2234-8352 (Online)
Print publication date 31 Aug 2020
Received 07 Jul 2020 Revised 02 Aug 2020 Accepted 07 Aug 2020
DOI: https://doi.org/10.5916/jamet.2020.44.4.311

Preliminary study on the docking control of a large ship using tugboats
Jin-Kyu Choi
Associate Professor, Ocean Science & Technology School, Korea Maritime & Ocean University, 727, Taejong-ro, Yeongdo-gu, Busan 49112, South Korea, Tel: 051-410-4342 (jk-choi@kmou.ac.kr)

Correspondence to : Associate Professor, Ocean Science & Technology School, Korea Maritime & Ocean University, 727, Taejong-ro, Yeongdo-gu, Busan 49112, South Korea, E-mail: jk-choi@kmou.ac.kr, Tel: 051-410-4342


Copyright ⓒ The Korean Society of Marine Engineering
This is an Open Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License (http://creativecommons.org/licenses/by-nc/3.0), which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.
Funding Information ▼

Abstract

In this study, we aim to automate the docking operation by manipulating a (disabled) large ship using multiple tugboats. To this end, as a preliminary study, we attempt to incorporate the dynamics of the ship and tugboats into the manipulation control. Assuming that the contact between the ship and each tugboat is a point contact, the velocity constraints at the contact points yield a reduced dynamic equation that can be expressed using independent variables, e.g., ship variables (surge, sway, and yaw rates), and that excludes contact force terms; this allows us to treat the ship and multiple tugboats as a single system to control the independent variables. We formulate the reduced dynamic equation of a ship and n number of tugboats and present a controller to manipulate the ship to track a straight-line path while maintaining a desired heading angle. However, a separate controller is required when using a single tugboat because a singular point appears when the contact point is aligned with the two centers of mass of the ship and tugboat. We conduct simulations of cases of using one, two, and three tugboat(s) to verify the effectiveness of the proposed control method based on the reduced dynamics.


Keywords: Tugboat, Docking, Ship, Reduced dynamic equation, Path-tracking control

1. Introduction

Operation of a ship in harbors, including the docking operation, is a tense task for onboard members because of the risk of collisions with other ships and harbor facilities. In particular, a large ship typically has powerful propellers that have poor maneuverability at a low speed; thus, it becomes difficult for the ship to move with a delicate motion, despite having a dynamic positioning system. Therefore, tugboats play an active role in maneuvering such large ships in harbors [1].

Meanwhile, pushing or pulling motion enables us to move a large object easily in our daily activities, and it can be more effective on the sea because the frictional force in sea is less than that on land. A tugboat utilizes this advantage to move large floating objects, such as (disabled) large ships, barges, oil platforms, caissons, etc. Moreover, multiple tugboats can cooperatively perform more delicate tasks like docking a large ship in a narrow space. However, the radio communications by humans between tugboats often interrupt the smooth flow of the operation, and operators are required to have considerable skills; this has encouraged researchers to automate the maneuvering operation of a large ship through multiple tugboats.

The research group in [2]-[5] investigated the manipulation problem of a large (disabled) ship using swarms of autonomous marine surface vehicles (tugboats). Each tugboat in contact with the ship was considered to be a three-degree-of-freedom (3-DOF) azimuth thruster of the ship, and its influence was treated as a pushing force to the ship. Ship dynamics were used to develop control strategies. Robust control methods based on the aforementioned concept have also been studied by other researchers [6][7].

The manipulation of a large ship by multiple surface vehicles is worth investigating as indicated in [2] and there are few related works. However, in robotics, there is a significant amount of research on the manipulation of a large object by multi-robot cooperation [8]-[16]. Related works were reviewed in [8]. Approaches using pushing [9]-[12], formation [13], grasping [14], caging [15], and object closure [16] have been investigated; these robotics technologies can be applied to marine robotics.

This study involves pose (position and orientation) control of a large ship by multiple tugboats with the aim of automating docking operation. To this end, we first challenge to incorporate the dynamics of both the ship and tugboats as a preliminary study; the previous research in [2]-[7] considered only the ship dynamics. The contact between the ship and each tugboat is assumed to be a point contact. The velocity constraints at the contact points yield a reduced dynamic equation that is expressed by independent variables, e.g., ship variables (surge, sway, and yaw rates), and excludes contact force terms; this allows us to treat the ship and multiple tugboats as a single system to control the independent variables. Section 2 of this paper formulates the reduced dynamic equation inspired by multi-fingered robot hand dynamics [17]. In Section 3, a controller to manipulate the ship to track a straight-line path while maintaining a desired heading angle is presented; a separate controller is, however, required for a single tugboat because a singular point appears when the contact point is aligned with two centers of mass of the ship and tugboat. Section 3 further considers a controller for this case. Lastly, Section 4 conducts simulations to verify the effectiveness of the presented control method based on the reduced dynamics.


2. Reduced dynamic equation

Consider a (disabled) ship manipulated by n number of tugboats as shown in Figure 1. To formulate the reduced dynamic equation, we use the following notations:


Figure 1:  
Ship manipulated by n number of tugboats

  • 𝑥𝑜-𝑦𝑜 coordinate system: the fixed frame
  • 𝑥𝑠-𝑦𝑠 coordinate system: the ship frame
  • 𝑢𝑠: the surge rate of the ship
  • 𝑢𝑡𝑖 : the surge rate of the i-th tugboat
  • 𝑣𝑠: the sway rate of the ship
  • 𝑣𝑡𝑖 : the sway rate of the i-th tugboat
  • 𝜔𝑠 : the yaw rate of the ship
  • 𝜔𝑡𝑖 : the yaw rate of the i-th tugboat
  • 𝑠𝑢𝑡𝑖, 𝑠𝑣𝑡𝑖, 𝑠𝜔𝑡𝑖 : the rates expressed with respect to (w.r.t.) the ship frame
  • 𝑽𝑠 = (𝑢𝑠 𝑣𝑠 𝜔𝑠)𝑇 ∈ ℜ3
  • 𝑽𝑡𝑖 = (𝑢𝑡𝑖 𝑣𝑡𝑖 𝜔𝑡𝑖 )𝑇 ∈ ℜ3
  • 𝑠𝑽𝑡𝑖 = ( 𝑠𝑢𝑡𝑖𝑠𝑣𝑡𝑖𝑠𝜔𝑡𝑖 )𝑇 ∈ ℜ3
  • 𝜓𝑠 : the yaw angle of the ship w.r.t. the fixed frame
  • 𝜓𝑡𝑖 : the yaw angle of the i-th tugboat w.r.t. the fixed frame
  • 𝑠𝜓𝑡𝑖 : the yaw angle of the i-th tugboat w.r.t. the ship frame
  • 𝝉𝑠 ∈ ℜ3: the resultant forces and moment applied to the ship
  • 𝝉𝑡𝑖 ∈ ℜ3 : the forces and moment exerted by the i-th tugboat, which are expressed w.r.t. its body frame
  • 𝑠𝝉𝑡𝑖 ∈ ℜ3 : 𝝉𝑡𝑖 expressed w.r.t. the ship frame
  • 𝒇𝑖 ∈ ℜ2: the contact force at the contact point of the i-th tugboat
  • 𝒓𝑖 ∈ ℜ2: the position vector pointing from the center of mass of the ship to the contact point of the i-th tugboat
  • 𝒓𝑡𝑖 ∈ ℜ2: the position vector pointing from the center of mass of the i-th tugboat to the contact point with the ship, which is expressed w.r.t. its body frame
  • 𝑠𝒓𝑡𝑖 ∈ ℜ2: 𝒓𝑡𝑖 expressed w.r.t. the ship frame

From Figures 1 and 2, we obtain the dynamic equations of the ship and the i-th tugboat expressed w.r.t. the ship frame as follows:


Figure 2:  
Force and moment relationships between the ship and the i-th tugboat

MsV˙s+CsVsVs+DsVsVs=τs(1) 
 sMti sV˙ti+ sCtiVti sVti+ sDtiVti sVti= sτti(2) 
 sMti=RMtiRTR3×3(3) 
 sCtiVti=RCtiVti-MtiRTR˙RTR3×3(4) 
 sDtiVti=RDtiVtiRTR3×3(5) 
 sτti=RτtiR3(6) 
RR sψtiR3×3:the rotation matrix of sψti (7) 

where 𝑀𝑠 = 𝑀𝑠𝑅𝐵 + 𝑀𝑠𝐴 ∈ ℜ3×3, 𝑀𝑠𝑅𝐵 and 𝑀𝑠𝐴 are the rigid-body inertia matrix and the added inertia matrix, respectively, 𝐶𝑠(𝑽𝑠) = 𝐶𝑠𝑅𝐵(𝑽𝑠) + 𝐶𝑠𝐴(𝑽𝑠) ∈ ℜ3×3, 𝐶𝑠𝑅𝐵(𝑽𝑠) and 𝐶𝑠𝐴(𝑽𝑠) are the rigid-body Coriolis and centripetal matrix, as well as the hydrodynamic Coriolis and centripetal matrix, respectively, and 𝐷𝑠(𝑽𝑠) ∈ ℜ3×3 is the hydrodynamic damping matrix [18]; these are for the ship and 𝑀𝑡𝑖, 𝐶𝑡𝑖(𝑽𝑡𝑖), and 𝐷𝑡𝑖(𝑽𝑡𝑖) represent the same things associated with the tugboat.

Assuming that the contact between the ship and each tugboat is a point contact, and considering the influence of the contact force to the ship and tugboat, as shown in Figure 2, Equations (1) and (2) can be rewritten as

MsV˙s+CsVsVs+DsVsVs=BsTf(8) 
 sMti sV˙ti+ sCtiVti sVti+ sDtiVti sVti=Rτti- sBtiTfi(9) 
BsT=I2I2r1TETrnTETR3×2n(10) 
 sBtiT=I2 srtiTETR3×2(11) 
f=f1fnR2n(12) 
E=0-110(13) 

where 𝐼2 is a 2×2 identity matrix, and E is the orthogonal rotation matrix rotating an arbitrary vector 90o counterclockwise in a plane.

Combining Equations (8) and (9) for a ship and n number of tugboats, we obtain the following:

MV˙+CVV+DVV=Tτt+BTf(14) 
V=Vs sVt1 sVtnR3+3n(15) 
M=blockdiag(Ms,  sMt1,, sMtn)(16) 
C(V)=blockdiag(Cs(Vs),  sCt1(Vt1),, sCtn(Vtn))(17) 
D(V)=blockdiag(Ds(Vs),  sDt1(Vt1),, sDtn(Vtn))(18) 
M,CV,D(V)R3+3n×(3+3n)(19) 
T=O3×3nblockdiag(R sψt1,,R( sψtn))R3+3n×3n(20) 
τt=τt1τtnR3n(21) 
BT=BsT- sBtTR3+3n×2n(22) 
 sBtT=blockdiag( sBt1T,, sBtnT)R3n×2n(23) 

Let 𝒗𝑠 = (𝑢𝑠 𝑣𝑠)𝑇, 𝑠𝒗𝑡𝑖 = ( 𝑠𝑢𝑡𝑖𝑠𝑣𝑡𝑖)𝑇 , and 𝒗𝑖 be the velocity of the contact point with the i-th tugboat; moreover, let us assume that the contact point does not change during operation.

From Figure 3, the velocity constraint at the contact point can then be expressed as


Figure 3:  
Velocity constraint at the contact point between the ship and the i-th tugboat

vi=vs+Eriωs= svti+E srti sωti(24) 

Combining all the velocity constraints at n number of contact points, we have

BsVs= sBt sVtBV=0(25) 

Here, selecting 𝑽𝑠 as independent variables, we obtain the following:

V=NVs(26) 
V˙=NV˙s+N˙Vs(27) 
N=I3 sBt#BsR3+3n×3(28) 
 sBt#:the pseudoinverse of  sBt(29) 

Substituting Equations (26) and (27) into Equation (14) and eliminating the contact force terms because 𝐵𝑁 = 𝑂2𝑛×3 from Equation (25), the reduced dynamic equation is finally expressed as follows:

M-V˙s+C-VVs+D-VVs=NTTτt(30) 
M-=NTMNR3×3(31) 
C-V=NT(CVN+MN˙)R3×3(32) 
D-V=NTDVNR3×3(33) 

3. Path-tracking control

As a complex trajectory can be formed by combinations of straight lines, we discuss the straight-line path-tracking control of a ship using multiple tugboats while maintaining a certain yaw angle (heading angle) of the ship. Figure 4 shows notations for tracking a particular straight-line path. 𝒑𝑠, 𝒑𝑝, and 𝑝𝒑𝑠 = (𝜂𝑥 𝜂𝑦)𝑇 are the position vectors of the ship, the origin of the 𝑥𝑝-𝑦𝑝 frame (path frame) w.r.t. the fixed frame, and the ship w.r.t. the path frame, respectively, and 𝜓𝑝, α, and 𝜂𝜓 are the angle of the 𝑥𝑝-axis from the 𝑥𝑜-axis, the angle from the front side of the ship to the moving direction, and the angle of the moving direction from the 𝑥𝑝-axis, respectively.


Figure 4:  
Tracking a straight-line path. The tugboats are omitted from the figure

The control purpose ensures that 𝜂𝑦 and 𝜂𝜓 are zeros, and 𝜂𝑦 and 𝜂𝜓 are obtained by

 pps=ηxηy=RT(ψp)(ps-pp)(34) 
ηψ=ψs-α-ψp(35) 

Letting 𝑈𝑑 and 𝑽𝑠𝑑 be the desired moving speed and the desired value of 𝑽𝑠, respectively, we can set 𝑽𝑠𝑑 as

Vsd=Udcosα-Udsinα-kyηy-kψηψ(36) 

From the reduced dynamic equation in Equation (30), the forces and moments required to be exerted by the tugboats to achieve 𝑽𝑠 → 𝑽𝑠𝑑 are obtained by

τt=M~KVsd-Vs+C~VVs+D~VVs(37) 
M~=NTT-1M-R3n×3(38) 
C~(V)=NTT-1C-(V)R3n×3(39) 
D~(V)=NTT-1D-(V)R3n×3(40) 
K=diag(ku,kv,kω)R3×3(41) 

where 𝐾 is a control gain matrix. Substituting Equation (37) into Equation (30) yields V˙s=K(Vsd-Vs), and 𝑽𝑠 ≈ 𝑽𝑠𝑑 if the elements of K are sufficiently large.

Here, we should note that 𝝉𝑡 in Equation (37) can be calculated if 𝑁𝑇𝑇 is invertible; however, when using a single tugboat, i.e., n = 1, it cannot be invertible when the contact point is aligned with the two centers of mass of the ship and tugboat as shown in Figure 5 (a) (singular configuration); this situation usually occurs when a single tugboat pushes the ship to move. To address this problem, a kinematic control approach is written as follows:


Figure 5:  
(a) Singular configuration. (b) Pushing the ship to track a given path by a single tugboat

τt1=K(Vsd-Vs)R3(42) 
Vsd=Udcosα-Udsinα-kyηy-kψηψ-kϕϕR3(43) 

where a control variable 𝜙 was added to the desired value of the yaw rate in Equation (36), 𝜙 = (𝜓𝑠α) − 𝜓𝑡1 is the angle from the line passing through the center of mass of the tugboat and the contact point to the moving direction as shown in Figure 5 (b), and α depends on the contact point here.

For more than two tugboats, singular configurations can also occur, e.g., when the lines passing through the contact points and the centers of mass of the tugboats intersect at a point; however, such singular configurations are difficult to occur in actual operations and thus this paper does not consider this issue.


4. Simulation

To verify the effectiveness of the path-tracking control method based on the reduced dynamics, simulations were conducted. It was assumed that the ship did not use its propulsive forces for movement (we can consider another floating object, e.g., barge, instead of a ship) and the tugboat can exert forces and moment for 3-DOF motion on a horizontal plane. The contact points were chosen arbitrarily to prove the correct functioning of the reduced dynamics.

The parameter values of the ship and tugboats used for simulations are shown in Table 1, and they were determined for future lab-scale experiments. L, W, m and 𝐼𝑧 are the length, width, mass, and moment of inertia, respectively. Xu˙, Yν˙, Yω˙, Nν˙, and Nω˙ are the added mass and inertia derivatives, 𝑋𝑢, 𝑌𝑣, 𝑌𝜔, 𝑁𝑣, and 𝑁𝜔 are the coefficients of the linear skin friction, and 𝑋|𝑢|𝑢, 𝑌|𝑣|𝑣, and 𝑁|𝜔|𝜔 are the coefficients of the quadratic drag (refer to [18]). 𝐼𝑧, Xu˙, Yν˙, and Nω˙ were obtained through calculations, and others were chosen by referring to the ship parameters in [18] and through simulations. Three cases were examined: the use of two tugboats, three tugboats, and a single tugboat.

Table 1: 
Parameter values of the ship and tugboats used for simulations. It is assumed that all tugboats are identical.
Parameter Unit Value Parameter Unit Value
Ship Tugboat Ship Tugboat
L× W m 2 × 1 1 × 0.5 𝑋𝑢 kg/s −2 −1.2
m kg 40 20 𝑌𝑣 kg/s −20 −12
𝐼𝑧 kgm2 15.83 1.98 𝑌𝜔 kg/s −8.7 −5.2
Xu˙ kg −12.03 −6.02 𝑁𝑣 kgm2/s −32 −19.2
Yν˙ kg −24.98 −12.49 𝑁𝜔 kgm2/s −20 −12
Yω˙ kg −0.14 −0.02 𝑋|𝑢|𝑢 kg/m −3 −1.5
Nν˙ kgm2 −0.35 −0.04 𝑌|𝑣|𝑣 kg/m −30 −15
Nω˙ kgm2 −1.23 −0.15 𝑁|𝜔|𝜔 kgm2 −30 −15

First, we tested the case when two tugboats were used, i.e., n = 2. The two tugboats were initially in contact with the ship as shown in Figure 6. 𝒓1 = (−1, 0.5)𝑇 , 𝒓2 = (−1, −0.5)𝑇 , 𝒓𝑡1 = 𝒓𝑡2 = (0.5, 0)𝑇 , and the origin and angle of the path frame were set as (2, 0)𝑇 and 𝜓𝑝 = 30o, respectively. The desired moving speed was 𝑈𝑑 = 1 knot, α in Figure 4 and Equation (35) is set as 0o, and the control gains are 𝑘𝑦 = 1, 𝑘𝜓 = 1.2, 𝑘𝑢 = 𝑘𝑣 = 𝑘𝜔 = 1000, i.e., 𝐾 = diag(1000, 1000, 1000). As observed in the figure, the ship tracks the given path well while maintaining the heading angle specified (α = 0o). Figure 7 presents the case when a given path is tracked while maintaining α = 45o . The result proved the good performance in this case.


Figure 6:  
Tracking a given path using two tugboats while maintaining α = 0o


Figure 7:  
Tracking a given path using two tugboats while maintaining α = 45o

Figures 8 and 9 are the cases of using three tugboats, i.e., n = 3. First, the initial conditions for the case of Figure 8 are 𝒓1 = (1, 0.5)𝑇, 𝒓2 = (0, 0.5)𝑇, 𝒓3 = (−1, 0.5)𝑇, 𝒓𝑡1 = 𝒓𝑡2 = 𝒓𝑡3 = (0.5, 0)𝑇, 𝜓𝑠 = 90o, and α = 135o. The origin and angle of the path frame, the desired moving speed, and the control gains are the same as those used in the previous simulations. For Figure 9, 𝒓1 = (1, 0.5)𝑇, 𝒓2 = (0, 0.5)𝑇, 𝒓3 = (−1, 0)𝑇, 𝒓𝑡1 = 𝒓𝑡2 = 𝒓𝑡3 = (0.5, 0)𝑇, 𝜓𝑠 = 90o , and α = 45o . The two simulation results verified that the controller worked correctly. Figure 10 shows the changes of α in the simulations shown in Figures 7, 8, and 9.


Figure 8:  
Tracking a given path using three tugboats while maintaining α = 135o


Figure 9:  
Tracking a given path using three tugboats while maintaining α = 45o


Figure 10:  
Changes of α in Figures 7, 8, and 9

In addition, simulation was conducted using a single tugboat, as shown in Figure 11. 𝒓1 = (−1, 0)𝑇, 𝒓𝑡1 = (0.5, 0)𝑇, 𝜓𝑝 = 10o, and α = 0o, and the control gains are 𝑘𝑦 = 1, 𝑘𝜓 = 1.92, 𝑘𝜙 = 2.55 , and 𝐾 = diag(100, 100, 100) . The figure shows that the ship tracks the given path well.


Figure 11:  
Tracking a given path using a single tugboat

From these simulation results, it was confirmed that the reduced dynamics functioned correctly; however, this study did not consider the friction angle of the contact force that assures the pushing motion of the tugboat for manipulation of the ship. For future works, we aim to incorporate/combine the contact force constraints into/with the reduced dynamics.


5. Conclusion

In this study, the manipulation of a large ship was conducted using multiple tugboats. We formulated a reduced dynamic equation that incorporates the dynamics of the ship and n number of tugboats and excludes the contact force terms; this allowed us to treat the ship and n number of tugboats as a single system to manipulate the ship. A pose control method based on the reduced dynamics was presented, and its effectiveness was verified through simulations. Our future works will involve incorporating the contact force constraints, as mentioned in the last part of Section 4, and conducting lab-scale experiments.


Acknowledgments

This work was supported by the Korea Maritime And Ocean University Research Fund.

Author Contributions

Conceptualization, J.-K. Choi; Methodology, J.-K. Choi; Validation, J.-K. Choi; Formal Analysis, J.-K. Choi; Investigation, J.-K. Choi; Resources, J.-K. Choi; Data curation, J.-K. Choi; Writing-Original Draft Preparation, J.-K. Choi; Writing-Review & Editing, J.-K. Choi; Visualization, J.-K. Choi; Supervision, J.-K. Choi; Project Administration, J.-K. Choi; Funding Acquisition, J.-K. Choi.


References
1. Tugboat, https://en.wikipedia.org/wiki/Tugboat, Accessed July 7, 2020.
2. M. Feemster, J. Esposito and J. Nicholson, “Manipulation of large objects by swarms of autonomous marine vehicles: part 1 - rotation,” Proceedings of the 38th Southeastern Symposium on System Theory, pp. 255-259, 2006.
3. E. T. Smith, M. G. Feemster, and J. M. Esposito, “Swarm manipulation of an unactuated surface vessel,” Proceedings of the 39th Southeastern Symposium on System Theory, pp. 16-20, 2007.
4. D. Braganza, M. Feemster, and D. Dawson, “Positioning of large surface vessels using multiple tugboats,” Proceedings of American Control Conference, pp. 912-917, 2007.
5. J. Esposito, M. Feemster, and E. Smith, “Cooperative manipulation on the water using a swarm of autonomous tugboats,” Proceedings of IEEE International Conference on Robotics and Automation, pp. 1501-1506, 2008.
6. V. P. Bui, S. W. Ji, K. H. Choi, and Y. B. Kim, “Nonlinear observer and sliding mode control design for dynamic positioning of a surface vessel,” Proceedings of 12th International Conference on Control, Automation, and Systems, pp. 1900-1904, 2012.
7. B. Bidikli, E. Tatlicioglu, and E. Zergeroglu, “Robust control design for positioning of an unactuated surface vessel,” Proceedings of International Conference on Intelligent Robots and Systems, pp. 1071-1076, 2015.
8. Y. Cao, A. Fukunaga, and A. Kahng, “Cooperative mobile robotics: Antecedents and directions,” Autonomous Robots, vol. 4, pp. 7-27, 1997.
9. S. Akella and M. Mason, “Posing polygonal objects in the plane by pushing,” Proceedings of IEEE International Conference on Robotics and Automation, vol. 3, pp. 2255-2262, 1992.
10. K. Lynch and M. Mason, “Stable pushing: Mechanics, controllability, and planning,” International Journal of Robotics Research, vol. 15, no. 6, pp. 533-556, 1996.
11. M. Matarić, M. Nilsson, and K. Simsarin, “Cooperative multi-robot box-pushing,” Proceedings of IEEE/RSJ International Conference on Intelligent Robots and Systems, vol. 3, pp. 556-561, 1995.
12. D. Rus, B. Donald, and J. Jennings, “Moving furniture with teams of autonomous robots,” Proceedings of IEEE/RSJ International Conference on Intelligent Robots and Systems, pp. 235-242, 1995.
13. J. Spletzer, A. Das, R. Fierro, C. Taylor, V. Kumar, and J. Ostrowski, “Cooperative localization and control for multi-robot manipulation,” Proceedings of IEEE/RSJ International Conference on Intelligent Robots and Systems, vol. 2, pp. 631-636, 2001.
14. A. Sudsang and J. Ponce, “On grasping and manipulating polygonal objects with disc-shaped robots in the plane,” Proceedings of IEEE International Conference on Robotics and Automation, pp. 2740-2746, 1998.
15. E. Rimon and A. Blake, “Caging planar bodies by one-parameter two-fingered gripping systems,” International Journal of Robotics Research, vol. 18, no. 3, pp. 299-318, 1999.
16. Z. Wang and V. Kumar, “Object closure and manipulation by multiple cooperating mobile robots,” Proceedings of IEEE International Conference on Robotics and Automation, pp. 394-399, 2002.
17. R. Murray, Z. Li, and S. Sastry, A Mathematical Introduction to Robotic Manipulation, CRC Press, Inc., 1994.
18. T. Fossen, Guidance and Control of Ocean Vehicles, 1st ed., John Wiley & Sons Ltd., 1994.