Open access
Technical Papers
Oct 19, 2015

Coordinated Motion Control of Distributed Spacecraft with Relative State Estimation

Publication: Journal of Aerospace Engineering
Volume 29, Issue 3

Abstract

This paper addresses the coordination problem of distributed spacecraft systems with time-invariant communication topologies. In order to estimate the neighbor spacecraft’ relative states which are not able to be measured directly, estimating processes are established on each spacecraft by utilizing an unscented Kalman filtering approach with measuring principles of both the pseudorange and the carrier phase. Based on such estimations, a distributed consensus protocol is proposed to address the coordinated motion problem of distributed spacecraft. The effectiveness of the proposed approach is demonstrated through numerical simulations, with a low-Earth-orbit satellites flying scenario.

Introduction

The coordination problem of distributed spacecraft systems has attracted much attention in recent years. Compared to solo systems, additional benefits such as high flexibility and great efficiency can be obtained by having a group of distributed spacecraft work cooperatively.
A fundamental approach to achieve such coordination is consensus, which enables all agents of a system to reach an agreement on certain quantities of interest. Consensus problems have been widely studied in recent years. Jadbabaie et al. (2003) analyzed the coordination behavior of a group of autonomous agents and provided a theoretical explanation for such behavior, and their results indicated that proper control models based only on neighbor information could motivate systems to achieve consensus. Ren et al. (2007) made an overview on information consensus of multiple vehicles by not only reviewing different consensus control laws but also describing several potential specific applications, such as rendezvous, formation stabilization, formation maneuvering, and flocking. Some more challenging results were obtained in Ren (2008) by extending consensus algorithms from first-order models to second-order ones, which made consensus approaches applicable for systems with second-order dynamics. Some practical conditions such as the actuator saturations and the lacks of relative velocity measurements were also discussed. In the context of spacecraft coordination, Ren (2007) considered consensus problems and showed the concept of coordinated motion. By utilizing a consensus control law with local information exchanges, the given formation configuration was maintained during the maneuvers, and all states of spacecraft were assumed available with respect to the Earth Centered Inertial (ECI) frame. In order to obtain accurate formation configurations, orbital reference frames were used to represent relative motions of spacecraft instead of the ECI frame (Chung et al. 2009; Li et al. 2010), and all relative states of spacecraft were also assumed available in this work.
However in practice, different from absolute states (i.e., positions and velocities), relative states (i.e., relative positions and relative velocities) are generally difficult to directly measure due to the limitation of sensors. Therefore, additional work should be done to obtain these relative quantities. Among varieties of methods to do so, although the direct method that simply subtracts between two irrelevant absolute states could be used to calculate the relative state, the results would be quite rough. In contrast, methods based on relative sensing are preferred. Carrier-phase-based differential global positioning system navigation is one such method that is applicable only for missions near the Earth [e.g., Mitchell (2004)]. For more-general missions where the global positioning system may not be available, such as deep-space missions, several alternative sensing technologies have been developed in recent years. Some common such technologies are based on lasers, charge coupled devices, and high-speed cameras. However, laser-based sensing is of high directionality, and image-based sensing requires complex image processing and target identification—with associated failures. Gunnam et al. (2002) developed a new vision-based navigation (VISNAV) sensor at Texas A&M University (College Station, Texas) for relative navigation. VISNAV uses at least four active beacons, which are at known positions in the body frame of one spacecraft, as light sources. It also uses a position-sensitive diode sensor, which is at a known position in the body frame of the other spacecraft, as the light receiver. Thus after gathering all the line-of-sight vectors generated by light sources and receiver, relative states between the two spacecraft can be estimated. Lau et al. (1996) provided an autonomous formation flying sensor (AFF) for relative navigation, which is built around commercially-available transmitter/receiver antennas at known positions in body frames of each spacecraft. By dealing with the measured pseudorange or phase observables, relative states can be solved to extremely high precision by the estimation. In the context of distributed spacecraft systems, since the vision-based method may be blocked by other spacecraft due to the complexity and flexibility of the formations, AFF is considered as the sensor whose sensing range can be from several meters to several kilometers.
This paper considers the coordinated motion problem of distributed spacecraft systems that consist of one free-flying spacecraft and several maneuvering spacecraft. All spacecraft are equipped with the AFF sensors, and besides, only the free-flying spacecraft has the ability to sense its absolute position, while the maneuvering spacecraft do not. The organization of this paper proceeds as follows. First, some background on algebraic graph theory, definition of coordinates, and AFF sensors are given. Then the measurement equations of the AFF sensor, which will be used for filtering, are derived in terms of both pseudorange and phase. Next, the equation of relative motion of spacecraft is reviewed and presented, and a set of unscented kalman filters are designed for relative estimation in a distributed way. During the design, the initial values of filters are given not by a priori assumption but by the solution provided by a nonlinear minimization algorithm. Next, a consensus control law is designed for the orbital coordinative motion. In the simulation, a low-Earth-orbit satellites flying scenario is taken as an instantiation, and subsequent results and analyses are presented. Finally, concluding remarks are made in the last section.

Backgrounds

Algebraic Graph Theory

The bidirectional information exchange among spacecraft can be modeled by undirected graphs. A weighted undirected graph G=(N,E) describes the communication links of a finite nonempty set of nodes N={1,2,,n} with EN×N being the edge set. In a weighted undirected graph, an edge (j,i)E denotes that spacecraft i and j can obtain information from each other, and they are called neighbors. The weighted adjacency matrix A=[aij]Rn×n of a weighted undirected graph is defined such that aij is a positive weight if (j,i)E, while aij=0 if (j,i)E (without loss of generality, it is considered that aij has been turned into the range of [0,1] in this paper). As (j,i)E implies (i,j)E, it follows that aij=aji, ij. The Laplacian matrix L=[lij]Rn×n associated with the graph is defined such that lii=j=1,jinaij, while lij=aij, ij. It is clear that A is symmetric, and L is symmetric, positive, and semidefinite when the undirected graph is connected [see e.g., Ren et al. (2007) for more].

Coordinates

In this paper, the free-flying spacecraft is treated as a reference spacecraft. To describe the absolute position of the reference spacecraft, an ECI frame named J2000 is used, while to describe the relative motion between every two spacecraft, an orbital reference frame named the local vertical/local horizontal (LVLH) frame is used. As shown in Fig. 1, OEXIYIZI denotes the ECI frame with the origin OE attached to the mass center of the Earth; ORXRYRZR denotes the LVLH frame with the origin OR attached to the mass center of the reference spacecraft; Point Oi denotes the mass center of the ith maneuvering spacecraft; and vector ρi denotes the relative position of the ith maneuvering spacecraft with respect to the LVLH frame. The LVLH frame is defined such that the XR axis is outward along the radial (local vertical), YR is perpendicular to XR in the orbit plane in the direction of motion (local horizontal), and ZR is along the orbit normal.
Fig. 1. Definition of the coordinates

AFF Sensor

AFF sensor was first devised for the Deep Space Mission 3 (DS-3) (Lau et al. 1996), where the stabilization of a fixed triangle formation is strictly required. In this mission, since the three spacecraft always face towards each other, AFF is configured with only a 2π steradian coverage for the purpose of saving cost and mass. There is also an alternative full configuration for more flexible formations, and in this paper, this full configuration is employed as shown in Fig. 2 for a full 4π steradian coverage. For simplicity, the shape of each spacecraft is simplified to a cube, and coordinate OiXiYiZi is the body fixed frame of spacecraft i (BF-i). Each spacecraft is equipped with the AFF consisting of eight antennas, of which six are receivers and two are transmitters at known positions in the BF-i frame. Symbol B is borrowed to indicate the receivers, while Symbol C indicates the transmitters. Superscripts are identifiers of distinct transmitters/receivers, and subscripts indicates the owner spacecraft of these antennas (e.g., Bi3 means Receiver 3 on spacecraft i, and Cj1 means the Transmitter 1 on spacecraft j). In a single-epoch measurement, ranging codes and carrier phases are generated upon each spacecraft, sent to the neighboring spacecraft, and processed in sequence.
Fig. 2. AFF sensor configuration

Relative Estimation

Measurement Equations

For convenience, the maneuvering spacecraft is labeled with number i=1,2,,n in turn, and to achieve a unified format, the reference spacecraft is labeled with number 0. Now consider the transmit–receive relations between two such distinct spacecraft i and j (i,j{0,1,2,,n} and ij) that can communicate with each other. Since there are two transmit antennas mounted on opposing corners on one spacecraft, there will always exist at least one transmit antenna that can be received by the other spacecraft. And in the meanwhile, there will always exist at least three active receive antennas on one spacecraft that can receive signals from the other spacecraft. Therefore, the directed transmit–receive paths between spacecraft i and j are at least six, among which only six will be selected (can be randomly) to take part in the measurement. Such six transmit–receive paths are illustrated in Fig. 3, where the arrows represent the directions of the information flows.
Fig. 3. Illustration of six transmit–receive paths
Pseudorange measurement equations will now be derived. As shown in Fig. 4, without loss of generality, focusing on one general path where the active receive antenna on spacecraft i is denoted as Bir and the active transmit antenna on spacecraft j is denoted as Cjt, r{1,2,,6}, t{1,2}. According to geometric vector addition, one obtains
CjtBirI=OiBirIOiOjIOjCjtI
(1)
where the superscripts of vectors indicate the coordinate frames in which the vectors are expressed, and I, L, i and j denote the ECI, LVLH, BF-i, and BF-j frames, respectively.
Fig. 4. Pseudorange measurement
Denoting the transformation matrix from frame F1 to F2 as AF2F1, Eq. (1) can be rewritten as
CjtBirI=AIi·OiBiriAIL·OiOjLAIj·OjCjtj
(2)
Then, taking measurement noises and clock errors into consideration, pseudorange measurement equations can be obtained as
CjtBirI=AIi·OiBiriAIL·OiOjLAIj·OjCjtj+nij+ε
(3)
where · = 2-norm operator; nij=ninj = clock error between spacecraft i and j, which is converted to the form of distance; and ε = measuring noise, which can be modeled to follow the normal distribution. Also note that Eq. (3) represents three equations corresponding to the three paths from spacecraft j to i.
The paths from spacecraft i to j are similar, and equations can be obtained by simply switching the variables i and j in Eq. (3) as
CitBjrI=AIj·OjBjrjAIL·OiOjLAIi·OiCiti+nji+ε
(4)
where nji=nij. In Eqs. (3) and (4), OiBiri, OjCjtj, OjBjrj, and OiCiti are known vectors, AIi=AiI1 and AIi=AiI1 with AiI and AjI are known as the attitude matrices of spacecraft i and j, AIL is the known transformation matrix from LVLH frame to ECI frame, while OiOjL and nij are unknown quantities to be solved.
As shown in Fig. 5, the differential phase measurement is abstracted. Denote Bir1 and Bir2 as the two receivers on spacecraft i that receive signals from the same transmitter Cjt on spacecraft j. Then difference the two received signals, the differential phase Δϕ can be measured. Define line-of-sight vectors as
los1=CjBir1CjBir1,los2=CjBir2CjBir2
(5)
Fig. 5. Differential phase measurement
Then the differential phase measurement equations can be expressed as
Δϕ=Bir1Bir2·los2+β+ηδ
(6)
where β = integer ambiguity; η = measurement noise; and δ = small distance, which can be calculated as
δ=CjBir1CjBir1·cosα=CjBir1(1los1·los2)
(7)
Since three receive antennas can only make two combinations, there are four differential phase measurements in a single epoch, and hence the equations are four. Combining six pseudorange measurement equations and four differential phase measurement equations yields the final measurement equations.

Relative Motion Equations

As defined previously, the LVLH frame attached to the free flying spacecraft is used as the orbital reference frame in this paper. Although the relative motion can be also described in similar orbital frames attached to any one of the maneuvering spacecraft, the maneuvering spacecraft will be maneuvering in general so each of them would not likely move in a fixed Keplerian orbit. Therefore, it is more accurate to use the LVLH frame attached to the free-flying spacecraft as opposed to the maneuvering spacecraft. Besides, although the relative motion is being studied, the rough orbital absolute information is needed due to the orbital characteristics, so in this paper it is assumed that the free-flying spacecraft is equipped with an absolute position sensor (e.g., the global positioning system) to provide a reference absolute position for the spacecraft cluster.
For simplicity, in this paper, the free-flying spacecraft is assumed to move in a nearly circular orbit. Using Kepler’s third law, the orbit angular velocity can be given by
ωo=μeRo3
(8)
where ωo = orbit angular velocity; μe = geocentric gravitational constant (398, 600.4418km3s2); and Ro = orbit radius. Denote the relative position of the ith maneuvering spacecraft with respect to the reference spacecraft in the LVLH frame as ρi=[xi,yi,zi]T. When the relative position between a maneuvering spacecraft i and the reference spacecraft are small compared to the orbit radius, the relative dynamics, known as the CW equations, can be derived as [see, Clohessy and Wiltshire (1960) and Lee et al. (2014) for more]
x¨i2ωoy˙i3ωo2xi=uix+dixmiy¨i+2ωox˙i=uiy+diymiz¨i+ωo2zi=uiz+dizmi
(9)
where mi = mass of the ith maneuvering spacecraft; uix, uiy, and uiz = control forces; and dix, diy, and diz = disturbance forces. Although Eq. (9) is an approximation and is not applicable for long-term orbital motion propagation, it can be used to predict the relative state in the next moment with enough precision. Furthermore, motion within the orbital plane (i.e., x and y axes) are coupled, while the motion perpendicular to the orbital plane (i.e., z axis) is independent from that within the orbital plane, and actually it is periodic vibratory. Denote vi=ρ˙i=[x˙i,y˙i,z˙i]T and define Xi=[ρiT,viT,niT]T, then the equation of Xi can be rewritten in the matrix form as
X˙i=FXi+GUi+GDi
(10)
with Fig. 6
G=[03×3I3×301×3]
(11)
Ui=1mi[uix,uiy,uiz]T,Di=1mi[dix,diy,diz]T
(12)
where 03×3 is a 3×3 zero matrix, and I3×3 is a 3×3 identity matrix. Furthermore, by defining Xij=XiXj, Uij=UiUj, and Dij=DiDj, the relative equations between two distinct maneuvering spacecraft i and j (i,j{1,2,,n} and ij) can be written as
X˙ij=FXij+GUij+GDij
(13)
Fig. 6. The system matrix F in the state differential equation Eq. (10)

Estimation

In this subsection, the method for relative estimations among multiple spacecraft will be presented. First, how the estimation process works at the system level will be shown. Since the spacecraft system is distributed and a host spacecraft does not exist, the estimation could be processed by many small estimators working together in a distributed way. It is clear that the number of small estimators is equal to that of undirected edges in the topology. To describe the allocation of these estimators, the sign matrix S=[sij]Rn×n is defined such that sij=1 if an estimator for Xij is running on spacecraft i, while sij=0 otherwise. Define si=k=0nsik, then the problem of allocating the estimators evenly is equivalent to the problem of trying to average si, i=0,1,2,,n. Hence a method to get an averaging decision is proposed as follows:
1.
For each spacecraft i{0,1,2,,n}, initialize sij=0, j=0,1,2,,n;
2.
During one communication, for each pair of neighbor spacecraft i and j, a random decision is made such that either sij=1, sji=0 or sij=0, sji=1;
3.
During a next communication, for each pair of neighbor spacecraft i and j, an improved decision is made by
sij(k)=sij(k1)1n+1aij[si(k1)sj(k1)]sij(k)=1,ifsij(k)>1sij(k)=0,ifsij(k)<0
(14)
where (k) = moment;
4.
Repeat Step. 3 for a given while. And then by rounding sij to the nearest integers, one obtains the final decision.
It can be seen from Eq. (14) that if spacecrafts i and j are neighbors, there comes aij>0, so that when si>sj (which means the work of spacecraft i is heavier than that of spacecraft j), sij will decrease (meanwhile sji will increase) to offload a little work from i to j, and when si<sj, then sij will increase to help share some work from j to i. Thus after some time, the amount of computations would be well offloaded to every spacecraft. Eq. (14) also constrains that 0sij1, so after rounding towards the nearest integers, the final decision can be obtained. The gain “1/n+1” was chosen to guarantee the convergence of the averaging iteration. Actually, by denoting the gain as κ, then the gain can be chosen as any values satisfying 0<κmin0in(1/j=0naij). This choosing principle can be proved as follows. Replace the gain with κ, and the first equation of Eq. (14) can be written as
sij(k)=sij(k1)aij·κ·[si(k1)sj(k1)]
(15)
One can obtain j=0nsij(k)=j=0n{sij(k1)aij·κ·[si(k1)sj(k1)]} by summing Eq. (15) together. Integrate these equations and get the matrix form as
sk=(Iκ·L)·sk1
(16)
where sk=[s0(k),s1(k),,sn(k)]T, I is the n+1-order identity matrix. Denote by λi the ith eigenvalue of (κ·L) in an increasing sequence and by μi the ith eigenvalue of Θ(Iκ·L), and it is easy to know that μi=1λi. Choose the gain following the principle of 0<κ1/j=0naij and the graph is connected; then it follows from Ren et al. (2007) that there is a single 0 in the eigenvalues of (κ·L) while all the other eigenvalues are positive but no more than 1, or in formula form, 0=λ0<λ1λn1. By doing so, the eigenvalues of Θ satisfy 0μnμ1<μ0=1, which implies the spectral radius of Θ is 1. It follows from Horn and Johnson (2012) that, limkΘk=1·ξT, where ξT is the left eigenvector of Θ corresponding to the eigenvalue μ0. Therefore after enough loops, sk converges following limksk=limkΘks0=1·ξT·s0=1·s¯, where s¯ denotes a constant. It means s1=s2==sn after convergence, which indicates that all the spacecraft finally reach consensus on the division of work.
After offloading computations to every spacecraft by the averaging decision, each spacecraft will then activate a corresponding estimator according to sij. These estimators are designed based on unscented Kalman filters as follows:
1.
Initialization: At moment 0, initialize the estimation of relative state X^ij(0), process noise matrix Qij(0), measurement noise matrix Rij(0), and error covariance matrix Pij(0). Further analyses and principles that accomplish the initialization will be given following the design.
2.
Sampling weights assignment: The weight assignment follows the minimal skew sampling method (Julier and Uhlmann 2002). Since Xij contains seven quantities, the number of sigma points should be 7+2=9. The weights are chosen as
Wm={1W027,m=1,22m1W1,m=3,,8
(17)
with W0 being a initial nonnegative weight, and in this paper W0=0.6.
3.
Unbiased sigma points construction of u: Since the covariance is already reflected in matrices Pij and Qij, the sigma points should be unbiased. Denote the nine sigma points as Sm7, m=0,,8 respectively, then their values can be given by following iterations
Sij={[S0j10],i=0[Sij112Wj+1],i=1,,j[0(j1)×112Wj+1],i=j+1
(18)
with three priori values
S01=0,S11=12W1,S21=12W1
(19)
4.
Global propagation: At moment k1, the purpose of the global propagation is to predict X^ij(k) roughly based on X^ij(k1), and the prediction is denoted as X^ij(k,k1). Then following Eq. (13), the prediction is given by
X^ij(k,k1)=Φ·X^ij(k1)+ΔT·GUij
(20)
where ΔT = time interval between estimations; and Φ = state transition matrix. In practice, it is of sufficient precision to preserve the first moment only, and hence one obtains
ΦI7×7+F·ΔT
(21)
5.
Error propagation: The purpose of error propagation is to evaluate the errors caused by uncertainties of X^ij(k1) during the global propagation. Since Eq. (20) is linear, the prediction itself can be omitted in this process. Based on the unbiased sigma points, perturbations before global propagation can be constructed as
χm(k1)=Pij(k1)+Qij(k1)·Sm7,m=0,,8
(22)
which will generate propagation errors as
χm(k,k1)=Φ·χm(k1),m=0,,8
(23)
6.
Measurement prediction: The measurement equations Eqs. (3), (4), and (6) are all functions of relative state X^, hence for clarity, they are treated as a whole, which is referred to as the following function
y=f(X^)
(24)
Therefore, the predictions of measurements are given by
Zm(k,k1)=f[X^ij(k,k1)+χm(k,k1)]
(25)
7.
Covariance prediction: Covariances are predicted as
Pij(k,k1)=m=08Wm[χm(k,k1)δX^ij(k,k1)]·[χm(k,k1)δX^ij(k,k1)]T+Qij(k1)Pijyy(k,k1)=m=08Wm[Zm(k,k1)Zij(k,k1)]·[Zm(k,k1)Zij(k,k1)]TPijxy(k,k1)=m=08Wm[χm(k,k1)δX^ij(k,k1)]·[Zm(k,k1)Zij(k,k1)]T
(26)
where
δX^ij(k,k1)=m=08Wmχm(k,k1)Zij(k,k1)=m=08WmZm(k,k1)
(27)
8.
Update: Once the true measurements Zij(k) occur at moment k, the estimation will be updated as
X^ij(k)=X^ij(k,k1)+Kij[Zij(k)Zij(k,k1)]
(28)
where
Kij=Pijxy(k,k1)[Pijyy(k,k1)+Rij(k)]1
(29)
9.
Assignment for next moment: After the estimation at moment k, the following assignments will be done to get ready for the estimation at moment k+1, which are
Pij(k)=Pij(k,k1)Kij[Pijyy(k,k1)+Rij(k)]KijTQij(k)=Qij(k1)Rij(k)=Rij(k1)
(30)
The initializations of estimators are assumed to be priori known in general studies, however, this assumption is difficult to realize in practice. In this paper, practical principles of initialization will be presented. X^ij contains three quantities which are ρ^ij, v^ij, and n^ij respectively. Regarding n^ij, the work done by Lau et al. (1996) show that clock error could be measured by a built-in approach with the precision of 0.1 ns, which is equivalent 0.03 m after being multiplied by the speed of light. Therefore, n^ij=0 can be initialized with the bound σn=0.03m. Regarding v^ij, the cluster of distributed spacecraft are assumed have been already launched approximately to a common orbit such that they can communicate with neighbors, and hence, their initial relative velocities are bounded. The bound can be approximately estimated simply. Since vij(ρij/Ro)Vo with Ro bigger than the Earth’s radius, Vo smaller than the first cosmic velocity, and ρij no bigger than several kilometers (in fact, the maximum effective distance of AFF is commonly 2 km) in general, there comes the inequality that vij<(2/6400)×7900<3m/s. Therefore, v^ij=[0,0,0]T can be initialized with the bound σv=3m/s. Regarding ρ^ij, it can be solved roughly based on Eqs. (3) and (4) ignoring the small nij. Thus the initialization of ρ^ij becomes the optimization problem optimizing ρij to make the results of Eqs. (3) and (4) best fit the pseudorange measurements. Since the number of equations (i.e., 6) is more than the number of variables (i.e., 3), this problem is a determined system. Therefore, many nonlinear minimization methods can be used for help. In this paper, the trust-region method is used to make a large-scale optimization, and then ρ^ij can be initialized to the optimization result with a confidence bound σρ. Detailed explanations about the algorithm procedure are omitted here due to its universal applicability, which can be found in many good articles, e.g., Coleman and Li (1996) and Conn et al. (2000). So X^ij(0) can be initialized until now, and subsequently Pij(0), Qij(0) and Rij(0) will be initialized as
Pij(0)=2[σρ2I3×3σv2I3×3σn2]Qij(0)=2ΔT·G·[σdx2σdy2σdz2]·GTRij(0)=2[σε2I6×6ση2I4×4]
(31)
where σdx, σdy, and σdz = confidence bound of the disturbance forces respectively; σε = confidence bound of pseudorange measurement noises; ση = confidence bound of differential phase measurement noises; and the constant factor 2 here means that corresponding covariances are doubled when differencing Xi and Xj.

Distributed Control Based on Estimations

Analyses on Design of Control Law

In this section, a distributed control law will be proposed for relative orbital motion. Suppose a formation configuration is desired for the spacecraft cluster, and it is described by a set of θij, i, j{0,1,2,,n}, which is to say, θij denotes the desired relative position between spacecraft i and j. Disturbance forces are generally considered as external disturbances, and hence the system dynamics used in design of the control law can be obtained from the governing dynamic Eq. (9) by ignoring the disturbance terms as
ρ˙i=viv˙i=Ui+M1ρi+M2vi,i=1,,n
(32)
where
M1=[3ωo20000000ωo2],M2=[02ωo02ωo00000]
(33)
It can be seen that the dynamics presented by Eq. (32) contain the two orbital terms M1ρi and M2vi, which are generated by uncontrolled orbital motion. And hence, he effects caused should be eliminated by these terms as much as possible. However, although the differential relative states ρij and vij can be obtained by estimation, ρi and vi themselves are not direct available to spacecraft i due to the fact spacecraft i will only communicate with its neighbors and could not know the global configuration. Hence, prior to the design of control law, a distributed method is proposed, which is inspired by Cao et al. (2010) and Meng et al. (2010), to help in obtaining the estimation of ρi.

Observing ρi Based on ρij

The observing method is proposed as
r^˙i=γrsgn[j=1naij(r^ir^jρ^ij)],i=1,,n
(34)
where sgn(·) = signum function; r^i = estimation of ρi; ρ^ij = estimation of ρij and is already known after the process shown in the previous section; and γr = positive gain satisfying γr>max1invi. The theoretical proof of the method is as follows. Denote r˜i=r^iρi, i=1,,n, and then Eq. (34) can be rewritten as
r˜˙i=ρ˙iγrsgn[j=1naij(r˜ir˜jρ^ij+ρiρj)]=ρ˙iγrsgn[j=1naij(r˜ir˜j)],i=1,,n
(35)
where the fact that ρ^ij=ρij=ρiρj is used to develop the equality. Defining r˜=[r˜1T,r˜2T,,r˜nT]T and ρ=[ρ1T,ρ2T,,ρnT]T, then Eq. (35) can be expressed in the matrix form as
r˜˙=γrsgn[(LI3×3)r˜]ρ˙
(36)
where is the Kronecker product operator. The subsequent proof can follow the process given by Meng et al. (2010). Since L is positive and semidefinite, consider the Lyapunov function candidate
V=12r˜T(LI3×3)r˜
(37)
Then the derivation of V along the trajectories of Eq. (36) is
V˙=r˜T(LI3×3)r˜˙=r˜T(LI3×3){γrsgn[(LI3×3)r˜]ρ˙}γr(LI3×3)r˜1+ρ˙·(LI3×3)r˜1=(γrρ˙)·(LI3×3)r˜10
(38)
where the Holders inequality is used to obtain the first Holder’s inequality and the fact γr>max1invi=ρ˙ is used to obtain the last inequality. It then follows from Lyapunov stability theory that Eq. (35) is asymptotically stable and r˜0 as t, and by Lasalle’s Invariance principle, it follows that r^iρi, i=1,,n as t. So there must exist such a time t¯ that satisfies r^i=ρi, i=1,,n with the margin of errors very small for all tt¯. In fact, it can be concluded from Cao and Ren (2012) that t¯<r˜T(0)(LI3×3)r˜(0)λmax(L)/(γrρ˙)λmin(L).

Design of Control Law

The control law is proposed as follows:
Ui={03×1,tt¯j=0naijγρtanh(ρ^ijθij)j=0naijγvtanh(v^ij)M1r^i,t>t¯,i=1,,n
(39)
where γρ = positive gain of the position term; γv = positive gain of the velocity term; and tanh(·) = hyperbolic tangent function and is used here to limit the output since the control forces are bounded in practice [e.g., Meng et al. (2013) and Yang et al. (2014)]. Using control law Eq. (40) for system dynamics Eq. (33), then the system is asymptotically stable. The proof is shown in the following. For tt¯, it is clear that system is always bounded, and for t>t¯, by combining Eqs. (32) and (39), one obtains
ρ˙i=viv˙i=j=0naijγρtanh(ρ^ijθij)j=0naijγvtanh(v^ij)+M2vi,i=1,,n
(40)
where term M1r^i counteracts the term M1ρi since r^i=ρi for tt¯. Consider the Lyapunov function candidate
V=12i=1nj=1n{aijγρ13Tln[cosh(ρijθij)]}+12i=1nviTvi+i=1n{ai0γρ13Tln[cosh(ρi0θi0)]}
(41)
where ln(·) = natural logarithm function; and cosh(·) = hyperbolic cosine function. Then the derivation of V along the trajectories of Eq. (40) is
V˙=12i=1nj=1n[aijγρvijTtanh(ρijθij)]+i=1nviT[j=0naijγρtanh(ρijθij)j=0naijγvtanh(vij)+M2vi]+i=1n[ai0γρvi0Ttanh(ρi0θi0)]=12i=1nj=1n[aijγρvijTtanh(ρijθij)]i=1nj=1n[aijγρviTtanh(ρijθij)]i=1n[viTai0γρtanh(ρi0θi0)]i=1nj=1n[aijγρviTtanh(vij)]i=1n[viTai0tanh(vi)]+i=1n(viTM2vi)+i=1n[ai0γρvi0Ttanh(ρi0θi0)]
(42)
By switching the dummy variables i and j in the summation function, one obtains
12i=1nj=1n[aijγρvijTtanh(ρijθij)]=i=1nj=1n[aijγρviTtanh(ρijθij)]
(43)
i=1nj=1n[aijγρviTtanh(vij)]=12i=1nj=1n[aijγρvijTtanh(vij)]
(44)
also, it can be seen from the definition that vi0vi, then Eq. (42) can be developed as
V˙=12i=1nj=1n[aijγρvijTtanh(vij)]i=1n[viTai0tanh(vi)]+i=1n(viTM2vi)
(45)
As viTM2vi0 and tanh(·) are odd functions, and then one can obtain
V˙=12i=1nj=1n[aijγρvijTtanh(vij)]i=1n[viTai0tanh(vi)]0
(46)
So the system is asymptotically stable and the equilibrium is achieved when ρij=θij and vij=03×1, i=1,,n as t, and thus the formation configuration of the spacecraft system will be formed.

Simulation

A spacecraft system that contains one free-flying spacecraft and four maneuvering spacecraft is taken as an example. A general communication topology is set for the system as shown in Fig. 7, which is fixed, undirected, and connected with communication link edges weighted by 0.8. The five spacecraft are set such that they move in a line initially in a same nearly-circular orbit with orbital semi-major axis 7378.137 km, eccentricity 2.41881×1016, inclination 60°, and right ascension of ascending node 0°. The desired formation offsets and other parameters are given as shown in Table 1. Denote the time when estimation is started as time origin, then the simulation is scheduled as follows. The establishment of communication links and allocation of estimators are processed during t[60,0]s, all estimators and all observing processes for rough ρi work during t[0,6000]s, and the distributed control law works during t[60,6000]s. Simulation results are as follows.
Fig. 7. Communication topology of the system
Table 1. Simulation Parameters
ParameterValue
Desired offset between Spacecrafts 0 and 1, θ10θ10=[300,1000,0]Tm
Desired offset between Spacecrafts 1 and 2, θ21θ21=[200,0,0]Tm
Desired offset between Spacecrafts 2 and 3, θ32θ32=[0,200,0]Tm
Desired offset between Spacecrafts 3 and 4, θ43θ43=[200,0,0]Tm
Communication and estimation time interval, ΔTΔT=1s
Spacecraft masses, mimi=100kg, i=0,,5
Disturbance forces, σx, σy, σzσx=σy=σz=106m/s2
Pseudorange measurement noises, σε3.33×103m
Differential phase measurement noises, σηση=3.33×106m
Initial estimation errors of position, σρσρ=40m
Initial estimation errors of position, σvσv=3m/s
Initial estimation clock errors, σnσn=3×102m
Gains in Eq. (34), γrγr=max{1,(600e0.03t200)}
Gains in Eq. (39), γρ, γvγρ=0.06, γv=0.3
At time t=60s, a random decision is made such that S=[0000010100000000010100000]. Following the definition of matrix S, it can be seen that Spacecrafts 1, 3, and 5 are idle, while Spacecrafts 2 and 4 both carry out two estimators. At time t=0s after the allocation, final decision is obtained and S=[0100000100000000010000010], which shows that the four estimators are successfully distributed evenly with Spacecraft 0 estimating X01, Spacecraft 1 estimating X12, Spacecraft 3 estimating X32, and Spacecraft 4 estimating X43.
Fig. 8 shows the 2D projections of trajectories of the four maneuvering spacecraft onto the x-y plane of the LVLH frame, where the points of the trajectories represent the mass centers of the spacecraft. Fig. 9 shows the detailed 3D position data. From both Figs. 8 and 9, it can be seen that the four maneuvering spacecraft have been successfully controlled to the desired formation configuration. Furthermore, the configuration errors are no more than 1 cm. Controlling is based on estimation, and the estimation results are as follows. During time t[0,2]s, the position initializations of the four estimators are obtained by the trust-region method, and results as shown in Table 2. Although the optimization processing does not take clock errors and measurement errors into consideration, and hence the results are rough, it still provide a good initialization (because errors are bounded by σρ) for further estimation. During time t[2,6000], the relative-position estimation errors and relative-velocity estimation errors are shown in Figs. 10 and 11 respectively, from which the following conclusion can be drawn. The errors of relative-position estimation are no more than 0.2 m throughout all the time after convergence and are much smaller when relative motion becomes slight, and the errors of relative-velocity estimation are no more than 0.005m/s. The results of the estimation for clock errors are similar, and they are omitted to save space. The estimation errors are very small, showing that the estimation is effective and helpful to controlling. Meanwhile during time t[0,6000], there are also four observers trying to observe ρi, i=1,,4, and the errors of the observing results are shown in Fig. 12, from which it can be seen that the errors have been decreased to no more than 10 m before t=60s. In the control law (40) r^i will be multiplied by matrix M1 and meanwhile the elements of M1 are small, then it is clear r^i and ρi make little difference to the control law.
Fig. 8. 2D projections of motion trajectories of the four maneuvering spacecraft onto the x-y plane
Fig. 9. 3D positions of the four spacecraft during maneuvering
Table 2. Optimization for Position Initialization
Relative positionSeeking originOptimizationTrue value
ρ10[0,0,0]Tm[6.43,734.68,33.28]Tm[0.04,735.46,0.45]Tm
ρ21[0,0,0]Tm[11.36,735.13,18.65]Tm[0.11,735.46,0.45]Tm
ρ32[0,0,0]Tm[26.69,734.85,14.70]Tm[0.18,735.46,0.45]Tm
ρ43[0,0,0]Tm[25.69,734.99,3.25]Tm[0.26,735.46,0.45]Tm
Fig. 10. Relative-position estimation errors
Fig. 11. Relative-velocity estimation errors
Fig. 12. Observation errors

Conclusion

In this paper, coordinated motion problem of distributed spacecraft systems that consist of one free-flying spacecraft and several maneuvering spacecraft is considered, and a relatively complete processing framework is proposed for the coordinated motion. In the upper level, a distributed algorithm is proposed to allocate the estimators, helping offloading computations to every spacecraft. In the lower level, estimators, observers, and controllers are all designed in a distributed way. Among them, estimators are designed based on Kalman fiters and solutions of initialization problems are also presented, observers work for rough positions of maneuvering spacecraft in order to provide good approximations for controlling, and controllers work based on consensus and only use the results of estimators and observers to achieve coordinative motion. Finally, simulations are carried out where a low-Earth-orbit satellite flying scenario is taken as an instantiation. Simulation results show that estimators, observers, and controllers work well with fixed, connected, and undirected communication topologies and configuration can be formed with centimeter-level accuracy, having demonstrated the effectiveness of proposed approach.

Acknowledgments

The authors would like to thank the support by the National High Technology Research and Development Program of China (Grant No. 2013AA122601).

References

Cao, Y., and Ren, W. (2012). “Distributed coordinated tracking with reduced interaction via a variable structure approach.” IEEE Trans. Automat. Contr., 57(1), 33–48.
Cao, Y., Ren, W., and Meng, Z. (2010). “Decentralized finite-time sliding mode estimators and their applications in decentralized finite-time formation tracking.” Syst. Control Lett., 59(9), 522–529.
Chung, S. J., Ahsun, U., and Slotine, J. J. E. (2009). “Application of synchronization to formation flying spacecraft: Lagrangian approach.” J. Guid. Control Dyn., 32(2), 512–526.
Clohessy, W., and Wiltshire, R. (1960). “Terminal guidance system for satellite rendezvous.” J. Aerosp. Sci., 27(9), 653–658.
Coleman, T. F., and Li, Y. (1996). “An interior trust region approach for nonlinear minimization subject to bounds.” SIAM J. Optim., 6(2), 418–445.
Conn, A. R., Gould, N. I., and Toint, P. L. (2000). Trust region methods, Vol. 1, SIAM, Philadelphia.
Gunnam, K. K., Hughes, D. C., Junkins, J. L., and Kehtarnavaz, N. (2002). “A vision-based DSP embedded navigation sensor.” Sens. J. IEEE, 2(5), 428–442.
Horn, R. A., and Johnson, C. R. (2012). Matrix analysis, Cambridge University Press, New York.
Jadbabaie, A., Lin, J., and Morse, A. S. (2003). “Coordination of groups of mobile autonomous agents using nearest neighbor rules.” IEEE Trans. Automat. Contr., 48(6), 988–1001.
Julier, S. J., and Uhlmann, J. K. (2002). “Reduced sigma point filters for the propagation of means and covariances through nonlinear transformations.” Proc., American Control Conf., IEEE, 887–892.
Lau, K., Lichten, S., Young, L., and Haines, B. (1996). “An innovative deep space application of gps technology for formation flying spacecraft.” Proc., AIAA GN&C Conf., AIAA, Reston, VA, 1–9.
Lee, D., Bang, H., Butcher, E. A., and Sanyal, A. K. (2014). “Kinematically coupled relative spacecraft motion control using the state-dependent riccati equation method.” J. Aerosp. Eng., 04014099.
Li, Z., Duan, Z., and Chen, G. (2010). “Consensus of multiagent systems and synchronization of complex networks: A unified viewpoint.” IEEE Trans. Circuits Syst., 57(1), 213–224.
Meng, Z., Ren, W., and You, Z. (2010). “Distributed finite-time attitude containment control for multiple rigid bodies.” Automatica, 46(12), 2092–2099.
Meng, Z., Zhao, Z., and Lin, Z. (2013). “On global leader-following consensus of identical linear dynamic systems subject to actuator saturation.” Syst. Control Lett., 62(2), 132–142.
Mitchell, M. L. (2004). “CDGPS-based relative navigation for multiple spacecraft.” M.S. thesis, Univ. of Texas at Austin, Austin, TX.
Ren, W. (2007). “Formation keeping and attitude alignment for multiple spacecraft through local interactions.” J. Guid. Control Dyn., 30(2), 633–638.
Ren, W. (2008). “On consensus algorithms for double-integrator dynamics.” IEEE Trans. Automat. Contr., 53(6), 1503–1509.
Ren, W., Beard, R. W., and Atkins, E. M. (2007). “Information consensus in multivehicle cooperative control: Collective group behavior through local interaction.” IEEE Contr. Syst. Mag., 27(2), 71–82.
Yang, T., Meng, Z., Dimarogonas, D. V., and Johansson, K. H. (2014). “Global consensus for discrete-time multi-agent systems with input saturation constraints.” Automatica, 50(2), 499–506.

Information & Authors

Information

Published In

Go to Journal of Aerospace Engineering
Journal of Aerospace Engineering
Volume 29Issue 3May 2016

History

Received: Sep 26, 2013
Accepted: Jul 29, 2015
Published online: Oct 19, 2015
Discussion open until: Mar 19, 2016
Published in print: May 1, 2016

Authors

Affiliations

Xiaochu Wang, Ph.D. [email protected]
Assistant Research Scientist, Qian Xuesen Laboratory of Space Technology, China Academy of Space Technology, No. 104 Youyi Rd., Beijing 100094, China; formerly, State Key Laboratory of Precision Measurement Technology and Instruments, Dept. of Precision Instrument, Tsinghua Univ., Beijing 100084, China (corresponding author). E-mail: [email protected]; [email protected]
Kaichun Zhao, Ph.D. [email protected]
Assistant Professor, State Key Laboratory of Precision Measurement Technology and Instruments, Dept. of Precision Instrument, Tsinghua Univ., Beijing 100084, China. E-mail: [email protected]
Zheng You, Ph.D. [email protected]
Professor, State Key Laboratory of Precision Measurement Technology and Instruments, Dept. of Precision Instrument, Tsinghua Univ., Beijing 100084, China. E-mail: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share