logo资料库

fastslam算法.pdf

第1页 / 共6页
第2页 / 共6页
第3页 / 共6页
第4页 / 共6页
第5页 / 共6页
第6页 / 共6页
资料共6页,全文预览结束
From: AAAI-02 Proceedings. Copyright © 2002, AAAI (www.aaai.org). All rights reserved. FastSLAM: A Factored Solution to the Simultaneous Localization and Mapping Problem Michael Montemerlo and Sebastian Thrun Daphne Koller and Ben Wegbreit School of Computer Science Carnegie Mellon University Pittsburgh, PA 15213 Computer Science Department Stanford University Stanford, CA 94305-9010 mmde@cs.cmu.edu, thrun@cs.cmu.edu koller@cs.stanford.edu, ben@wegbreit.com Abstract The ability to simultaneously localize a robot and ac- curately map its surroundings is considered by many to be a key prerequisite of truly autonomous robots. How- ever, few approaches to this problem scale up to handle the very large number of landmarks present in real envi- ronments. Kalman filter-based algorithms, for example, require time quadratic in the number of landmarks to in- corporate each sensor observation. This paper presents FastSLAM, an algorithm that recursively estimates the full posterior distribution over robot pose and landmark locations, yet scales logarithmically with the number of landmarks in the map. This algorithm is based on an ex- act factorization of the posterior into a product of con- ditional landmark distributions and a distribution over robot paths. The algorithm has been run successfully on as many as 50,000 landmarks, environments far be- yond the reach of previous approaches. Experimental results demonstrate the advantages and limitations of the FastSLAM algorithm on both simulated and real- world data. Introduction The problem of simultaneous localization and mapping, also known as SLAM, has attracted immense attention in the mo- bile robotics literature. SLAM addresses the problem of building a map of an environment from a sequence of land- mark measurements obtained from a moving robot. Since robot motion is subject to error, the mapping problem neces- sarily induces a robot localization problem—hence the name SLAM. The ability to simultaneously localize a robot and accurately map its environment is considered by many to be a key prerequisite of truly autonomous robots [3, 7, 17]. The dominant approach to the SLAM problem was in- troduced in a seminal paper by Smith, Self, and Cheese- man [16]. This paper proposed the use of the extended Kalman filter (EKF) for incrementally estimating the poste- rior distribution over robot pose along with the positions of the landmarks. In the last decade, this approach has found widespread acceptance in field robotics, as a recent tutorial paper [2] documents. Recent research has focused on scal- ing this approach to larger environments with more than a Copyright c 2002, American Association for Artificial Intelli- gence (www.aaai.org). All rights reserved. few hundred landmarks [6, 8, 9] and to algorithms for han- dling data association problems [18]. A key limitation of EKF-based approaches is their compu- tational complexity. Sensor updates require time quadratic in the number of landmarks K to compute. This complex- ity stems from the fact that the covariance matrix maintained by the Kalman filters has O(K 2) elements, all of which must be updated even if just a single landmark is observed. The quadratic complexity limits the number of landmarks that can be handled by this approach to only a few hundred— whereas natural environment models frequently contain mil- lions of features. This shortcoming has long been recog- nized by the research community [6, 8, 15]. In this paper we approach the SLAM problem from a Bayesian point of view. Figure 1 illustrates a generative probabilistic model (dynamic Bayes network) that underlies the rich corpus of SLAM literature. In particular, the robot poses, denoted s1, s2, . . . , st, evolve over time as a function of the robot controls, denoted u1, . . . , ut. Each of the land- mark measurements, denoted z1, . . . , zt, is a function of the position θk of the landmark measured and of the robot pose at the time the measurement was taken. From this diagram it is evident that the SLAM problem exhibits important condi- tional independences. In particular, knowledge of the robot’s path s1, s2, . . . , st renders the individual landmark measure- ments independent. So for example, if an oracle provided us with the exact path of the robot, the problem of determin- ing the landmark locations could be decoupled into K inde- pendent estimation problems, one for each landmark. This observation was made previously by Murphy [13], who de- veloped an efficient particle filtering algorithm for learning grid maps. Based on this observation, this paper describes an efficient SLAM algorithm called FastSLAM. FastSLAM decomposes the SLAM problem into a robot localization problem, and a collection of landmark estimation problems that are con- ditioned on the robot pose estimate. As remarked in [13], this factored representation is exact, due to the natural con- ditional independences in the SLAM problem. FastSLAM uses a modified particle filter for estimating the posterior over robot paths. Each particle possesses K Kalman fil- ters that estimate the K landmark locations conditioned on the path estimate. The resulting algorithm is an instance of the Rao-Blackwellized particle filter [5, 14]. A naive im- plementation of this idea leads to an algorithm that requires AAAI-02 593
1 s1 s s . . . s 2 Figure 1: The SLAM problem: The robot moves from pose s1 through a sequence of controls, u1, u2, . . . , ut. As it moves, it observes nearby landmarks. At time t = 1, it observes landmark θ1 out of two landmarks, fθ1, θ2g. The measurement is denoted z1 (range and bearing). At time t = 1, it observes the other landmark, θ2, and at time t = 3, it observes θ1 again. The SLAM problem is concerned with estimating the locations of the landmarks and the robot’s path from the controls u and the measurements z. The gray shading illustrates a conditional independence relation. O(M K) time, where M is the number of particles in the particle filter and K is the number of landmarks. We de- velop a tree-based data structure that reduces the running time of FastSLAM to O(M log K), making it significantly faster than existing EKF-based SLAM algorithms. We also extend the FastSLAM algorithm to situations with unknown data association and unknown number of landmarks, show- ing that our approach can be extended to the full range of SLAM problems discussed in the literature. Experimental results using a physical robot and a robot simulator illustrate that the FastSLAM algorithm can han- dle orders of magnitude more landmarks than present day approaches. We also find that in certain situations, an in- creased number of landmarks K leads to a mild reduction of the number of particles M needed to generate accurate maps—whereas in others the number of particles required for accurate mapping may be prohibitively large. SLAM Problem Definition The SLAM problem, as defined in the rich body of litera- ture on SLAM, is best described as a probabilistic Markov chain. The robot’s pose at time t will be denoted st. For robots operating in the plane—which is the case in all of our experiments—poses are comprised of a robot’s x-y coordi- nate in the plane and its heading direction. ferred to as the motion model: Poses evolve according to a probabilistic law, often re- p(st j ut, st−1) (1) Thus, st is a probabilistic function of the robot control ut and the previous pose st−1. In mobile robotics, the motion model is usually a time-invariant probabilistic generalization of robot kinematics [1]. The robot’s environment possesses K immobile land- marks. Each landmark is characterized by its location in space, denoted θk for k = 1, . . . , K. Without loss of gen- erality, we will think of landmarks as points in the plane, so that locations are specified by two numerical values. To map its environment, the robot can sense landmarks. For example, it may be able to measure range and bearing to 594 AAAI-02 a landmark, relative to its local coordinate frame. The mea- surement at time t will be denoted zt. While robots can often sense more than one landmark at a time, we follow com- monplace notation by assuming that sensor measurements correspond to exactly one landmark [2]. This convention is adopted solely for mathematical convenience. It poses no restriction, as multiple landmark sightings at a single time step can be processed sequentially. often referred to as the measurement model: Sensor measurements are governed by a probabilistic law, p(zt j st, θ, nt) (2) Here θ = fθ1, . . . , θkg is the set of all landmarks, and nt 2 f1, . . . , Kg is the index of the landmark perceived at time t. For example, in Figure 1, we have n1 = 1, n2 = 2, and n3 = 1, since the robot first observes landmark θ1, then landmark θ2, and finally landmark θ1 for a second time. Many measurement models in the literature assume that the robot can measure range and bearing to landmarks, con- founded by measurement noise. The variable nt is often referred to as correspondence. Most theoretical work in the literature assumes knowledge of the correspondence or, put differently, that landmarks are uniquely identifiable. Practi- cal implementations use maximum likelihood estimators for estimating the correspondence on-the-fly, which work well if landmarks are spaced sufficiently far apart. In large parts of this paper we will simply assume that landmarks are iden- tifiable, but we will also discuss an extension that estimates the correspondences from data. We are now ready to formulate the SLAM problem. Most generally, SLAM is the problem of determining the location of all landmarks θ and robot poses st from measurements zt = z1, . . . , zt and controls ut = u1, . . . , ut. In probabilis- tic terms, this is expressed by the posterior p(st, θ j zt, ut), where we use the superscript t to refer to a set of variables from time 1 to time t. If the correspondences are known, the SLAM problem is simpler: p(st, θ j zt, ut, nt) (3) As discussed in the introduction, all individual landmark es- timation problems are independent if one knew the robot’s path st and the correspondence variables nt. This condi- tional independence is the basis of the FastSLAM algorithm described in the next section. FastSLAM with Known Correspondences We begin our consideration with the important case where the correspondences nt = n1, . . . , nt are known, and so is the number of landmarks K observed thus far. Factored Representation The conditional independence property of the SLAM prob- lem implies that the posterior (3) can be factored as follows: p(st, θ j zt, ut, nt) = p(st j zt, ut, nt) p(θk j st, zt, ut, nt) (4) k Put verbally, the problem can be decomposed into K+1 esti- mation problems, one problem of estimating a posterior over robot paths st, and K problems of estimating the locations q q
of the K landmarks conditioned on the path estimate. This factorization is exact and always applicable in the SLAM problem, as previously argued in [13]. The FastSLAM algorithm implements the path estimator p(st j zt, ut, nt) using a modified particle filter [4]. As we argue further below, this filter can sample efficiently from this space, providing a good approximation of the poste- rior even under non-linear motion kinematics. The land- mark pose estimators p(θk j st, zt, ut, nt) are realized by Kalman filters, using separate filters for different landmarks. Because the landmark estimates are conditioned on the path estimate, each particle in the particle filter has its own, lo- cal landmark estimates. Thus, for M particles and K land- marks, there will be a total of KM Kalman filters, each of dimension 2 (for the two landmark coordinates). This repre- sentation will now be discussed in detail. Particle Filter Path Estimation FastSLAM employs a particle filter for estimating the path posterior p(st j zt, ut, nt) in (4), using a filter that is similar (but not identical) to the Monte Carlo localization (MCL) algorithm [1]. MCL is an application of particle filter to the problem of robot pose estimation (localization). At each point in time, both algorithms maintain a set of particles rep- resenting the posterior p(st j zt, ut, nt), denoted St. Each particle st,[m] 2 St represents a “guess” of the robot’s path: (5) We use the superscript notation [m] to refer to the m-th par- ticle in the set. The particle set St is calculated incrementally, from the set St−1 at time t−1, a robot control ut, and a measurement zt. First, each particle st,[m] in St−1 is used to generate a probabilistic guess of the robot’s pose at time t St = fst,[m]gm = fs[m] t gm , . . . , s[m] , s[m] 1 2 s[m] t p(st j ut, s[m] t−1), (6) obtained by sampling from the probabilistic motion model. This estimate is then added to a temporary set of parti- cles, along with the path st−1,[m]. Under the assumption that the set of particles in St−1 is distributed according to p(st−1 j zt−1, ut−1, nt−1) (which is an asymptotically cor- rect approximation), the new particle is distributed accord- ing to p(st j zt−1, ut, nt−1). This distribution is commonly referred to as the proposal distribution of particle filtering. After generating M particles in this way, the new set St is obtained by sampling from the temporary particle set. Each particle st,[m] is drawn (with replacement) with a probability proportional to a so-called importance factor w[m] , which is calculated as follows [10]: t p(st,[m] j zt, ut, nt) target distribution w[m] = t = proposal distribution p(st,[m] j zt−1, ut, nt−1) The exact calculation of (7) will be discussed further below. The resulting sample set St is distributed according to an ap- proximation to the desired pose posterior p(st j zt, ut, nt), an approximation which is correct as the number of particles M goes to infinity. We also notice that only the most recent robot pose estimate s[m] t−1 is used when generating the parti- cle set St. This will allows us to silently “forget” all other (7) pose estimates, rendering the size of each particle indepen- dent of the time index t. Landmark Location Estimation FastSLAM represents the conditional landmark estimates p(θk j st, zt, ut, nt) in (4) by Kalman filters. Since this estimate is conditioned on the robot pose, the Kalman filters are attached to individual pose particles in St. More specifi- cally, the full posterior over paths and landmark positions in the FastSLAM algorithm is represented by the sample set St = fst,[m], µ[m] 1 , Σ[m] 1 , . . . , µ[m] K , Σ[m] K gm (8) k k k (9) and Σ[m] is a two-element vector, and Σ[m] Here µ[m] are mean and covariance of the Gaus- sian representing the k-th landmark θk, attached to the m-th particle. In the planar robot navigation scenario, each mean µ[m] k The posterior over the k-th landmark pose θk is easily ob- tained. Its computation depends on whether or not nt = k, that is, whether or not θk was observed at time t. For nt = k, we obtain is a 2 by 2 matrix. p(θk j st, zt, ut, nt) Bayes/ p(zt j θk, st, zt−1, ut, nt) p(θk j st, zt−1, ut, nt) Markov= p(zt j θk, st, nt) p(θk j st−1, zt−1, ut−1, nt−1) For nt 6= k, we simply leave the Gaussian unchanged: p(θk j st, zt, ut, nt) = p(θk j st−1, zt−1, ut−1, nt−1) (10) The FastSLAM algorithm implements the update equation (9) using the extended Kalman filter (EKF). As in existing EKF approaches to SLAM, this filter uses a linearized ver- j st, θ, nt) [2]. Thus, sion of the perceptual model p(zt FastSLAM’s EKF is similar to the traditional EKF for SLAM [16] in that it approximates the measurement model using a linear Gaussian function. We note that, with a lin- ear Gaussian observation model, the resulting distribution p(θk j st, zt, ut, nt) is exactly a Gaussian, even if the mo- tion model is not linear. This is a consequence of the use of sampling to approximate the distribution over the robot’s pose. One significant difference between the FastSLAM algo- rithm’s use of Kalman filters and that of the traditional SLAM algorithm is that the updates in the FastSLAM algo- rithm involve only a Gaussian of dimension two (for the two landmark location parameters), whereas in the EKF-based SLAM approach a Gaussian of size 2K+3 has to be updated (with K landmarks and 3 robot pose parameters). This cal- culation can be done in constant time in FastSLAM, whereas it requires time quadratic in K in standard SLAM. Calculating the Importance Weights Let us now return to the problem of calculating the impor- tance weights w[m] needed for particle filter resampling, as defined in (7): p(st,[m] j zt, ut, nt) t w[m] t / p(st,[m] j zt−1, ut, nt−1) p(zt, nt j st,[m], zt−1, ut, nt−1) p(zt, nt j zt−1, ut, nt−1) Bayes= AAAI-02 595
k £ 4 ? k £ 4 ? F F T T k £ 2 ? k £ 2 ? k £ 6 ? k £ 6 ? T T F F T T F F k £ 1 ? k £ 1 ? k £ 3 ? k £ 3 ? k £ 5 ? k £ 5 ? k £ 7 ? k £ 7 ? T T F F T T [m] [m] 1,S 1,S [m] [m] 1 1 [m] [m] 2,S 2,S [m] [m] 2 2 [m] [m] 3,S 3,S [m] [m] 3 3 F F [m] [m] 4,S 4,S T T F F T T F F [m] [m] 4 4 [m] [m] 5,S 5,S [m] [m] 5 5 [m] [m] 6,S 6,S [m] [m] 6 6 [m] [m] 7,S 7,S [m] [m] 7 7 [m] [m] 8,S 8,S [m] [m] 8 8 Figure 2: A tree representing K = 8 landmark estimates within a single particle. p(st,[m] j zt−1, ut, nt) p(st,[m] j zt−1, ut, nt) p(zt, nt j zt−1, ut, nt−1) p(zt, nt j st,[m], zt−1, ut, nt−1) p(zt, nt j st,[m], zt−1, ut, nt−1) p(zt, nt j θ, st,[m], zt−1, ut, nt−1) p(θ j st,[m], zt−1, ut, nt) dθ t ) , nt) p(nt j θ, s[m] p(zt, nt j θ, s[m] p(θ j st−1,[m], zt−1, ut−1, nt−1) dθ p(zt j θ, s[m] p(θ j st−1,[m], zt−1, ut−1, nt−1) dθ p(zt j θ, s[m] p(θ j st−1,[m], zt−1, ut−1, nt−1) dθ p(zt j θ[m] , nt) p(θ[m] nt ) dθnt , s[m] , nt) nt ) t t t t (11) = / = Markov= = / EKF j θ, s[m] Here we assume that the distribution p(nt ) is uniform—a common assumption in SLAM. In the last line, “EKF” makes explicit the use of a linearized model as an ap- proximation to the observation model p(zt j θ[m] ), and the resulting Gaussian posterior p(θ[m] nt ). The final integra- tion is easily calculated in closed form for a linear Gaussian. nt , s[m] t t Efficient Implementation The FastSLAM algorithm, as described thus far, may require time linear in the number of landmarks K for each update iteration if implemented naively. This is because of the re- sampling step; every time a particle is added to St, its has to be copied. Since each particle contains K landmark esti- mates, this copying procedure requires O(M K) time. How- ever, most of this copying can be avoided. Our approach makes it possible to execute a FastSLAM iteration in O(M log K) time. The basic idea is that the set of Gaussians in each particle is represented by a balanced bi- nary tree. Figure 2 shows such a tree for a single particle, in the case of 8 landmarks. The Gaussian parameters µ[m] and Σ[m] are located at the leaves of the tree. Clearly, accessing k k 596 AAAI-02 T T k £ 4 ? k £ 4 ? F F k £ 2 ? k £ 2 ? F F T T k £ 3 ? k £ 3 ? F F T T 3,S 3,S [m] [m] [m] [m] 3 3 new particle new particle k £ 4 ? k £ 4 ? k £ 4 ? k £ 4 ? k £ 4 ? k £ 4 ? k £ 4 ? F F F F F F F T T T T T T T old particle old particle k £ 2 ? k £ 2 ? k £ 2 ? k £ 2 ? k £ 2 ? k £ 2 ? k £ 2 ? T T T T T T T F F F F F F F k £ 6 ? k £ 6 ? k £ 6 ? k £ 6 ? k £ 6 ? k £ 6 ? k £ 6 ? T T T T T T T F F F F F F F k £ 1 ? k £ 1 ? k £ 1 ? k £ 1 ? k £ 1 ? k £ 1 ? k £ 1 ? k £ 3 ? k £ 3 ? k £ 3 ? k £ 3 ? k £ 3 ? k £ 3 ? k £ 3 ? k £ 1 ? k £ 1 ? k £ 1 ? k £ 1 ? k £ 1 ? k £ 1 ? k £ 1 ? k £ 3 ? k £ 3 ? k £ 3 ? k £ 3 ? k £ 3 ? k £ 3 ? k £ 3 ? T T T T T T T F F F F F F F T T T T T T T F F F F F F F T T T T T T T F F F F F F F T T T T T T T F F F F F F F [m] [m] [m] [m] [m] [m] [m] 1,S 1,S 1,S 1,S 1,S 1,S 1,S [m] [m] [m] [m] [m] [m] [m] 1 1 1 1 1 1 1 [m] [m] [m] [m] [m] [m] [m] 2,S 2,S 2,S 2,S 2,S 2,S 2,S [m] [m] [m] [m] [m] [m] [m] 2 2 2 2 2 2 2 [m] [m] [m] [m] [m] [m] [m] 3,S 3,S 3,S 3,S 3,S 3,S 3,S [m] [m] [m] [m] [m] [m] [m] 3 3 3 3 3 3 3 [m] [m] [m] [m] [m] [m] [m] 4,S 4,S 4,S 4,S 4,S 4,S 4,S [m] [m] [m] [m] [m] [m] [m] 4 4 4 4 4 4 4 [m] [m] [m] [m] [m] [m] [m] 5,S 5,S 5,S 5,S 5,S 5,S 5,S [m] [m] [m] [m] [m] [m] [m] 5 5 5 5 5 5 5 [m] [m] [m] [m] [m] [m] [m] 6,S 6,S 6,S 6,S 6,S 6,S 6,S [m] [m] [m] [m] [m] [m] [m] 6 6 6 6 6 6 6 [m] [m] [m] [m] [m] [m] [m] 7,S 7,S 7,S 7,S 7,S 7,S 7,S [m] [m] [m] [m] [m] [m] [m] 7 7 7 7 7 7 7 [m] [m] [m] [m] [m] [m] [m] 8,S 8,S 8,S 8,S 8,S 8,S 8,S [m] [m] [m] [m] [m] [m] [m] 8 8 8 8 8 8 8 Figure 3: Generating a new particle from an old one, while modi- fying only a single Gaussian. The new particle receives only a par- tial tree, consisting of a path to the modified Gaussian. All other pointers are copied from the generating tree. each Gaussian requires time logarithmic in K. Suppose FastSLAM incorporates a new control ut and a new measurement zt. Each new particle in St will differ from the corresponding one in St−1 in two ways: First, it will possess a different path estimate obtained via (6), and second, the Gaussian with index nt will be different in ac- cordance with (9). All other Gaussians will be equivalent to the generating particle. 3 3 and Σ[m] When copying the particle, thus, only a single path has to be modified in the tree representing all Gaussians. An example is shown in Figure 3: Here we assume nt = 3, that is, only the Gaussian parameters µ[m] are updated. Instead of generating an entirely new tree, only a single path is created, leading to the Gaussian nt = 3. This path is an incomplete tree. To complete the tree, for all branches that leave this path the corresponding pointers are copied from the tree of the generating particle. Thus, branches that leave the path will point to the same (unmodified) subtree as that of the generating tree. Clearly, generating such an incomplete tree takes only time logarithmic in K. Moreover, accessing a Gaussian also takes time logarithmic in K, since the number of steps required to navigate to a leaf of the tree is equivalent to the length of the path (which is by definition logarithmic). Thus, both generating and accessing a partial tree can be done in time O(log K). Since in each updating step M new particles are created, an entire update requires time in O(M log K). Data Association In many real-world problems, landmarks are not identifi- able, and the total number of landmarks K cannot be ob- tained trivially—as was the case above. In such situations, the robot has to solve a data association problem between momentary landmarks sightings zt and the set of landmarks in the map θ. It also has to determine if a measurement cor- responds to a new, previously unseen landmark, in which case the map should be augmented accordingly. m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m m
(a) (b) (c) Figure 4: (a) Physical robot mapping rocks, in a testbed developed for Mars Rover research. (b) Raw range and path data. (c) Map generated using FastSLAM (dots), and locations of rocks determined manually (circles). In most existing SLAM solutions based on EKFs, these problems are solved via maximum likelihood. More specif- ically, the probability of a data association nt is given by = PF p(nt j zt, ut) Markov= m m Bayes/ p(nt j st, zt, ut) p(st j zt, ut) dst p(nt j st,[m], zt, ut) p(nt j s[m] p(zt j s[m] , nt) , zt) t t (12) m The step labeled “PF” uses the particle filter approxima- tion to the posterior p(st j zt, ut). The final step assumes a uniform prior p(nt j st), which is commonly used [2]. The maximum likelihood data association is simply the in- dex nt that maximizes (12). If the maximum value of p(nt j zt, ut)—with careful consideration of all constants in (12)—is below a threshold α, the landmark is considered previously unseen and the map is augmented accordingly. t t p(ztjs[m] t = argmaxnt In FastSLAM, the data association is estimated on a per- particle basis: n[m] , nt). As a result, different particles may rely on different values of n[m] . They might even possess different numbers of landmarks in their respective maps. This constitutes a primary difference to EKF approaches, which determine the data association only once for each sensor measurement. It has been observed frequently that false data association will make the conven- tional EKF approach fail catastrophically [2]. FastSLAM is more likely to recover, thanks to its ability to pursue multi- ple data associations simultaneously. Particles with wrong data association are (in expectation) more likely to disap- pear in the resampling process than those that guess the data association correctly. We believe that, under mild assumptions (e.g., minimum spacing between landmarks and bounded sensor error), the data association search can be implemented in time loga- rithmic in N. One possibility is the use of kd-trees as an indexing scheme in the tree structures above, instead of the landmark number, as proposed in [11]. Experimental Results The FastSLAM algorithm was tested extensively under vari- ous conditions. Real-world experiments were complimented by systematic simulation experiments, to investigate the scaling abilities of the approach. Overall, the results indicate favorably scaling to large number of landmarks and small particle sets. A fixed number of particles (e.g., M = 100) appears to work well across a large number of situations. Figure 4a shows the physical robot testbed, which consists of a small arena set up under NASA funding for Mars Rover research. A Pioneer robot equipped with a SICK laser range finder was driven along an approximate straight line, gener- ating the raw data shown in Figure 4b. The resulting map generated with M = 10 samples is depicted in Figure 4c, with manually determined landmark locations marked by circles. The robot’s estimates are indicated by x’s, illustrat- ing the high accuracy of the resulting maps. FastSLAM re- sulted in an average residual map error of 8.3 centimeters, when compared to the manually generated map. Unfortunately, the physical testbed does not allow for sys- tematic experiments regarding the scaling properties of the approach. In extensive simulations, the number of land- marks was increased up to a total of 50,000, which Fast- SLAM successfully mapped with as few as 100 particles. Here, the number of parameters in FastSLAM is approx- imately 0.3% of that in the conventional EKF. Maps with 50,000 landmarks are out of range for conventional SLAM techniques, due to their enormous computational complex- ity. Figure 5 shows example maps with smaller numbers of landmarks, for different maximum sensor ranges as indi- cated. The ellipses in Figure 5 visualize the residual uncer- tainty when integrated over all particles and Gaussians. In a set of experiments specifically aimed to elucidate the scaling properties of the approach, we evaluated the map and robot pose errors as a function of the number of landmarks K, and the number of particles M, respectively. The results are graphically depicted in Figure 6. Figure 6a illustrates that an increase in the number of landmarks K mildly re- duces the error in the map and the robot pose. This is be- cause the larger the number of landmarks, the smaller the robot pose error at any point in time. Increasing the number of particles M also bears a positive effect on the map and pose errors, as illustrated in Figure 6b. In both diagrams, the bars correspond to 95% confidence intervals. AAAI-02 597
(a) (b) Figure 5: Maps and estimated robot path, generated using sensors with (a) large and (b) small perceptual fields. The correct landmark locations are shown as dots, and the estimates as ellipses, whose sizes correspond to the residual uncertainty. Conclusion We presented the FastSLAM algorithm, an efficient new so- lution to the concurrent mapping and localization problem. This algorithm utilizes a Rao-Blackwellized representation of the posterior, integrating particle filter and Kalman filter representations. Similar to Murphy’s work [13], FastSLAM is based on an inherent conditional independence property of the SLAM problem, using Rao-Blackwellized particle fil- ters in the estimation. However, Murphy’s approach main- tains grid maps with discrete values similar to occupancy grid maps [12], hence does not address the common SLAM problem of estimating continuous landmark locations. In FastSLAM, landmark estimates are efficiently repre- sented using tree structures. Updating the posterior requires O(M log K) time, where M is the number of particles and K the number of landmarks. This is in contrast to the O(K 2) complexity of the common Kalman-filter based ap- proach to SLAM. Experimental results illustrate that Fast- SLAM can build maps with orders of magnitude more land- marks than previous methods. They also demonstrate that under certain conditions, a small number of particles works well regardless of the number of landmarks. Acknowledgments We thank Kevin Murphy and Nando de Freitas for insightful discussions on this topic. This research was sponsored by DARPA’s MARS Program (Contract number N66001-01-C-6018) and the National Science Foundation (CA- REER grant number IIS-9876136 and regular grant number IIS- 9877033). We thank the Hertz Foundation for their support of Michael Montemerlo’s graduate research. Daphne Koller was supported by the Office of Naval Research, Young Investigator (PECASE) grant N00014-99-1-0464. This work was done while Sebastian Thrun was visiting Stanford University. References localization for mobile robots. ICRA-99. [2] G. Dissanayake, P. Newman, S. Clark, H.F. Durrant-Whyte, and M. Csorba. An experimental and theoretical investigation into simultaneous localisation and map building (SLAM). Lecture Notes in Control and Information Sciences: Exper- imental Robotics VI, Springer, 2000. [3] G. Dissanayake, P. Newman, S. Clark, H.F. Durrant-Whyte, and M. Csorba. A solution to the simultaneous localisation and map building (SLAM) problem. IEEE Transactions of Robotics and Automation, 2001. [4] A. Doucet, J.F.G. de Freitas, and N.J. Gordon, editors. Se- quential Monte Carlo Methods In Practice. Springer, 2001. [5] A Doucet, N. de Freitas, K. Murphy, and S. Russell. Rao- Blackwellised particle filtering for dynamic Bayesian net- works. UAI-2000. Accuracy of FastSLAM as K Increases Robot Position Error Landmark Position Error 10 50 100 500 1000 5000 Number of landmarks K Accuracy of FastSLAM as Number of Particles Increases Robot Position Error Landmark Position Error (a) (b) 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 2.5 2 1.5 1 0.5 i ) n o i t a v e d d r a d n a t s ( r o r r E n o i t i s o P ) n o i t i a v e d d r a d n a t s ( r o r r E n o i t i s o P 0 0 50 100 150 Number of Particles 200 250 Figure 6: Accuracy of the FastSLAM algorithm as a function of (a) the number of landmarks N, and (b) the number of particles M. Large number of landmarks reduce the robot localization error, with little effect on the map error. Good results can be achieved with as few as 100 particles. [6] J. Guivant and E. Nebot. Optimization of the simultaneous localization and map building algorithm for real time imple- mentation. IEEE Transaction of Robotic and Automation, May 2001. [7] D. Kortenkamp, R.P. Bonasso, and R. Murphy, editors. AI- based Mobile Robots: Case studies of successful robot sys- tems, MIT Press, 1998. [8] J.J. Leonard and H.J.S. Feder. A computationally efficient method for large-scale concurrent mapping and localization. ISRR-99. [9] F. Lu and E. Milios. Globally consistent range scan alignment for environment mapping. Autonomous Robots, 4, 1997. [10] N. Metropolis, A.W. Rosenbluth, M.N. Rosenbluth, A.H. Teller, and E. Teller. Equations of state calculations by fast computing machine. Journal of Chemical Physics, 21, 1953. [11] A.W. Moore. Very fast EM-based mixture model clustering using multiresolution kd-trees. NIPS-98. [12] H. P. Moravec. Sensor fusion in certainty grids for mobile robots. AI Magazine, 9(2):61–74, 1988. NIPS-99. [14] K. Murphy and S. Russell. Rao-blackwellized particle fil- tering for dynamic bayesian networks. In Sequential Monte Carlo Methods in Practice, Springer, 2001. [15] P. Newman. On the Structure and Solution of the Simulta- neous Localisation and Map Building Problem. PhD thesis, Univ. of Sydney, 2000. [16] R. Smith, M. Self, and P. Cheeseman. Estimating uncertain spatial relationships in robotics. In Autonomous Robot Vehni- cles, Springer, 1990. [17] C. Thorpe and H. Durrant-Whyte. Field robots. ISRR-2001. [18] S. Thrun, D. Fox, and W. Burgard. A probabilistic approach to concurrent mapping and localization for mobile robots. Machine Learning, 31, 1998. [1] F. Dellaert, D. Fox, W. Burgard, and S. Thrun. Monte Carlo [13] K. Murphy. Bayesian map learning in dynamic environments. 598 AAAI−02
分享到:
收藏