Abstract
Magnetic fields generated by normal or superconducting electromagnets are used to guide and focus particle beams in storage rings, synchrotron light sources, mass spectrometers, and beamlines for radiotherapy. The accurate determination of the magnetic field by measurement is critical for the prediction of the particle beam trajectory and hence the design of the accelerator complex. In this context, state-of-the-art numerical field computation makes use of boundary-element methods (BEM) to express the magnetic field. This enables the accurate computation of higher-order partial derivatives and local expansions of magnetic potentials used in efficient numerical codes for particle tracking. In this paper, we present an approach to infer the boundary data of an indirect BEM formulation from magnetic field measurements by ensemble Kálmán filtering. In this way, measurement uncertainties can be propagated to the boundary data, magnetic field and potentials, and to the beam related quantities derived from particle tracking. We provide results obtained from real measurement data of a curved dipole magnet using a Hall probe mapper system.
1 Introduction
Static magnetic fields generated by electromagnets are used in particle accelerators to guide and focus particle beams. Studying the dynamics of the beam is indispensable in the design and operational phase of an accelerator complex and requires the accurate mathematical modeling of the magnetic fields. Preferably, the field model is an exact magneto-static solution obeying Maxwell’s equations.
Hybrid models combine first principles modeling and data-driven modeling to improve the predictive accuracy while preserving the underlying physical relationships. In this way, the differences between the magnet model and the real physical object can be incorporated in the numerical field simulation and particle tracking while preserving the underlying physical nature.
Classically, the integrated field in an accelerator magnet is expressed by eigenfunctions of the Laplace equation in polar coordinates, so called field harmonics. It is common practice to measure the field harmonics by rotating induction-coil systems and to use the measurement results in particle tracking simulations. This approach combines first principles with measurement data and may therefore be regarded as a classical example of hybrid modeling.
The advantages of this method are far-reaching. One can exploit scaling laws derived from field harmonics to reconstruct the field based on measurements on a single radius while still maintaining an exact magneto-static solution. Moreover, sensor systems can be designed particularly for the measurement of certain multipole components and specialized measurement equipment allows for the accurate determination of multipole errors even under the influence of mechanical vibration and shaft deformation [43].
In high energy physics, the particle beams are often contained in straight cylindrical vacuum chambers. In such cases the theory of field harmonics can be extended to three-dimensional cylindrical coordinates by generalized-gradients (see [36, Chapter 6], [37] and [47, Section 1.3.2]).
The motion of charged particles in the accelerator is computed numerically in particle tracking codes. The treatment of two-dimensional field harmonics is well established [47, Chapters 2 and 3]. The treatment of fringe fields and field inhomogeneities using generalized gradients was studied in [46], [34] and [5]. A more general and efficient approach based on a finite-element formulation was developed in [41].
In energy storage rings, beamlines for radiotherapy, mass spectrometers but also detector magnets, the fringe fields in the magnet ends give a relevant contribution to the equations of motion and particles might occupy more complex geometric domains. Hence, more general expressions for the magnetic field in the vacuum chamber are required. Boundary-element methods have been proposed as a field model to provide local expansions of the magnetic vector potential in the vicinity of a reference trajectory; [10, 2]. The local field description may be based on a Taylor approximation or an expansion into solid harmonics. An approach for the efficient treatment of such field descriptions in particle tracking is found in [48].
The quantitative characterization of uncertainties in model parameters due to random input data is the main objective of uncertainty quantification. Deterministic approaches for the uncertainty quantification for boundary-element methods with random right-hand sides are found in [8] and the references therein. Moreover, in [6] an approach for the shape inversion of a random scatterer, based on noisy measurement data is presented.
This article follows up on the advances of [21], but extends the analysis in four ways:
A more general field representation by means of an indirect boundary-element method is adopted in order to cope with more general beamlines in particle accelerators and detectors.
The efficiency of the Bayesian inference is enhanced by an ensemble Kálmán filter, where the statistics of the boundary data is expressed by means of an ensemble of state vectors. This approach is well known in the discipline of data assimilation, where the propagation of the full covariance matrix becomes infeasible due to the high dimension of the state space (see [4]).
As rotating induction-coil measurements are tailored to cylindrical domains, a more flexible Hall probe mapper system is used, which allows us to sample the field locally within a larger class of physical domains. We derive a realistic noise model for this particular measurement system, which includes the effect of position errors and vibrations.
An iterative learning algorithm is proposed that uses the spatial distribution of uncertainty in the predicted field to explore the spatial domain. This makes the inference of boundary data more effective, since measurements in regions with low uncertainty are reduced.
Section 2 of this paper covers the indirect boundary element method, employing a double layer potential that yields a ”lightweight” physical representation for the magnetic vector potential. Section 3 then turns to the evaluation of measurement data, which is suffering from uncertainties. In Section 4 we give details about a real world application to which the partial results of Section 5 will be applied. Particularly, we consider the use-case of the field quality measurement of a strongly-curved dipole magnet. The main result of Section 5 is an ensemble Kálmán filter for the inference of boundary data. It bypasses the need for the propagation of the full posterior covariance matrix and thus enables the efficient uncertainty quantification.
2 The Indirect Double Layer Potential Formulation
Consider an open and simply connected Lipschitz domain Ω in the air gap, or vacuum chamber of an accelerator magnet, with boundary
for the magnetic flux density
holds for the magnetic scalar potential
The vector
Remark 1.
The boundary
One benefit of the field representation by ν is that a magnetic vector potential
This follows from the definition of the magnetic vector potential and the equivalence of the double layer and eddy ring [28].
Remark 2.
Although the boundary data ν in the indirect representation formula is fictitious, evaluating (2.3) always yields an exact magneto-static solution, i.e., (2.1) holds in Ω for
2.1 Discretization
In boundary-element methods (BEM), the boundary data is approximated by locally supported basis functions on a boundary mesh. The algorithms presented in this paper have been implemented in C++ using the Boundary Element Based Engineering Library (BEMBEL) [7]. This enables a discretization of
The normal component of the magnetic flux density
The component
where
The k-th tensor product basis-splines is denoted by
where
is the vectorial surface curl and
The regularization term
Remark 3.
In many applications, a magnet simulation based on the finite element method (FEM) provides a low order approximation of
Remark 4.
Given a low order FEM solution, we can evaluate
We can derive an equation for the evaluation of
where
The Aubin–Nitsche formalism allows to derive point-wise error estimates for the evaluation of the discretized boundary integral equations (see [45, Theorem 12.8], [38, Example 4.2.15]). As it will be relevant for the upcoming discussions, we provide the result for the indirect Neumann problem in Theorem 1.
Theorem 1.
For the solution of the Neumann problem, using the indirect double-layer potential, and assuming a sufficiently smooth boundary, there holds the point-wise estimate [9, Theorem 7.1]
where
The constant
2.2 A Local Expansion for Particle Tracking
In particle beam dynamics the motion of a relativistic particle is usually expressed in so called phase-space coordinates. The momentum variables in phase-space coordinates are more abstract and differ from the ones used in classical mechanics, but this has the advantage that certain invariants of motion are conserved for a passage of the particle through a static magnetic field (see [47, Chapter 2]).
From the Euler–Lagrange formalism, the equations of motion result in Hamilton’s equations
where
where c denotes the velocity of light, q the particle charge and m its mass. Hamilton’s equations (2.11) may be formulated as an ordinary differential equation system of the type
Its solution by numerical integration for a given initial condition is the main objective of single particle tracking.[1]
Due to the partial derivatives
In principle one could apply the representation formula (2.4) directly in the numerical integration scheme. The spatial derivatives can be computed by differentiating the Green’s function
More efficient approaches have been proposed, using local expansions of the magnetic vector potential around a reference trajectory by either Taylor approximations [11] or the fitting of
Using Taylor expansions makes it necessary to impose (2.1) to recover a magneto-static solution. This is implied by using the solid harmonic functions. We therefore follow the approach developed in [2], but in contrast to fitting
Theorem 2 ([29, Section 3.4]).
Consider a point
where
The functions
By Theorem 2, the magnetic vector potential may be expanded at the position
This expansion is valid inside the largest sphere with center
The evaluation position
where
The moments
Remark 5.
The approach based on Theorem 2 can also be applied to other field representations based on integral formulations, such as BEM-FEM coupling or volume integral methods.
Theorem 3.
The error in the truncation
with
where
Proof.
The truncation error is given by the coefficients
Moreover, with
with
for
The truncation error for the evaluation of
Figure 2 shows the theoretical bounds for the truncation error as a function of the ratio
3 The Hall Probe Mapper System
The Hall probe mapper system is shown in Figure 3 (left). It uses the stages of a coordinate measuring machine (CMM) to manipulate the position of a three-axes Hall sensor in the magnetic field. The system allows for a distance triggered acquisition of the Hall voltages during the stage movement. This improves the required measurement duration compared to measurements in start-stop mode, but also causes the mapper arm to vibrate due to perturbations of the stage motion in the
A mathematical model for three axes Hall probe measurements is now derived. This mathematical model is coupled with the boundary integral equation for the magnetic flux density
3.1 The Observation Operator
The voltages measured by Hall sensors can have complicated dependencies on the three components of the magnetic flux density. These stem from planar Hall effects and geometric imperfections, but also from non-linearity errors of the sensor transfer function. Calibration techniques allow for the characterization of planar and axial Hall effects by rotating the sensor in a reference field. This gives rise to the sensitivity function
with the complex calibration coefficients
Remark 6.
In case of a linear Hall sensor, the sensitivity function
To establish the observation operator, we define the three Hall voltages by
and sort the measurement vector according to the scheme
for
Remark 7.
One benefit of the indirect double layer potential formulation is that the observation operator imposes a magneto-static solution without the necessity to solve a linear equation system. To see this, we consider, as an alternative, a field model based on a finite element approximation. The coefficients
3.2 The Noise Model
For the uncertainty quantification, a noise model is needed which represents the real measurement data. We therefore follow the guide to the expression of uncertainty in measurement (GUM) [49]. It proposes to introduce input quantities to the measurement equation (3.3), in order to model the relevant physical phenomena, which are perturbing the measurements. Then, the uncertainty in the measurands is quantified by means of a first order Taylor approximation of the measurement equation with respect to the input quantities. This procedure is justified as long as the values of the input quantities are small deviations from zero.
In case of the Hall probe mapper system, the measurements are mostly affected by perturbations of the sensor position and orientation, due to the vibration of the mapper arm. We therefore include five degrees of freedom for each measurement position as input quantities. Three displacements, denoted as
The mechanical perturbations are caused by random perturbations of the stage motion and are therefore not deterministic. However, the vibrations of the mapper arm lead to correlations between the measuring positions. These correlations have been determined by optical measurements in the machine setup. This yields a statistical model for the mechanical perturbations, which is of the type
As proposed in the GUM, the impact on the measurement vector
The matrices
Proposition 1.
With the Gaussian model for the mechanical perturbations
Proof.
The electronic noise is assumed to be additive, such that
Remark 8.
Both covariance matrices for mechanical and electronic noise,
4 Magnetic Measurements and the Prior Field Solution
It is possible to obtain the boundary data from a numerical field simulation. This procedure was described in detail in Section 2. The following section aims at deriving a hybrid model, that combines the first principle modeling (see Section 2) with data driven modeling, based on measurements [25]. In this context, it is important to consider the measurement uncertainties described in the previous section.
The inference of boundary data from measurements leads to the concept of Bayesian inference. Before going into detail, we will give an example for the practical application of our approach. The theoretical considerations of the next section are applied to this practical example.
We consider the problem of determining the field quality in a curved dipole magnet, shown in Figure 4 (left). It is a spare magnet for in the Extra Low ENergy Antiproton ring (ELENA) [32]. The design of these magnets was associated with several challenges due to the low working range in terms of magnetic flux density in the iron yoke, also the tough field quality requirements in combination with a small bending radius of only
We consider a curved domain of interest Ω, which is covering the magnet’s air gap and fringe field region. This domain is shown in Figure 4 (right) together with a boundary mesh used for the numerical approximation. Some parameters related to the domain of interest can be found in Table 1.
Magnet Parameters | ||||
Min/Max Gap Height | Pole Width | Mean Radius of Curvature | Bending Angle | Good Field Region (
|
|
|
|
|
|
Boundary Parameters | ||||
Height | Width | Length (z) | Number of Elements | Number of DoFs |
|
|
|
8576 | 13402 |
A numerical field simulation of the magnet is available by means of a lowest order finite element solution using the software Opera 3D and it is used to construct a prior BEM approximation. We will see in Section 5 how this field solution can be incorporated as a prior for the hybrid model.
We make use of the approach discussed in Remark 4 and perform
Remark 9.
It should be noted that the role of the FEM solution is to provide an initial guess for the Neumann data and its smoothness in the fringe field region. As it is the case for the magnet under consideration, standard tools for magnet design use low-order finite element methods, with a considerably lower convergence compared to a higher-order BEM formulation. Approximation errors in the process of fitting the Neumann data are negligible with respect to the engineering bias due to manufacturing tolerances, as well as uncertainties in geometric and material properties (see Figure 8 (left)). This engineering bias must be corrected by magnetic measurements.
The Hall probe mapper system is configured to sample the field along the domain boundary. Due to the thickness of the supporting structure and also the sensor housing, it is not possible to sample the
At this point we need to be careful not to distribute measurements too close to the boundary
The absolute values for the approximation error depend on the smoothness of the solution and cannot be determined a priori. A-posteriori error estimates for boundary element methods are found in [26]. As we are generating a boundary mesh based on a given FEM approximation, the approximation errors can be estimated by computing the difference between the BEM and FEM solutions in the vicinity of the boundary. In the example presented in this work, approximation errors are found to fall below 2 units in 10 000, within a distance of
We can exploit the physical relations imposed by the field model and place measurements only at the boundary of the measurement domain
The Hall probe mapper is equipped with a calibrated three axes Hall sensor, hosting an electronic board for signal conditioning. We are therefore working with a linear model for the sensitivity function s, with
5 Measurement Postprocessing
In order to update our prior assumptions on the boundary data, which were obtained from the first principle modeling discussed in Section 2, we follow the Bayesian viewpoint where the state vector
In the following, we denote by
Central to Bayesian inference is Bayes’ rule of probability
which states that the posterior pdf
The likelihood function is derived by coupling the statistical model for the measurement noise with the numerical model for the measurement operation by means of the observation operator
The impact of mechanical vibrations and positioning errors depends on the magnetic field, and therefore the state vector
In this work, we will make use of the prior state vector estimated from the numerical magnet simulation. The inference from measurement data aims to correct small differences in
5.1 The Kálmán Filter
The update step of a linear Kálmán filter is derived from the posterior under the assumption of a linear observation operator
and the expected value
A well-established alternative representation for
where
This follows from the Sherman–Morrison–Woodbury formula [40].
Both formulations require a matrix inversion to compute the posterior covariance matrix
Theorem 4.
The solution
is a sample from the Gaussian distribution
Proof.
Again we make use of the linearity of the expected value function, to show that
and
Remark 10.
The noise covariance matrix
Approximating the posterior covariance matrix by an ensemble of K state vectors with
In Figure 6, we illustrate the posterior covariance matrix approximated for different ensemble sizes K and also the case of a direct inversion of
Remark 11.
The convergence of the posterior covariance matrix for
5.2 The Ensemble Kálmán Filter
The equations for the Kálmán filter require the inverse of the prior covariance matrix
In the case of the ensemble Kálmán filter, the prior and posterior covariance matrices are represented by ensembles of state vectors. In this way a significant reduction of computational complexity can be achieved.
Similarly to (5.7) the ensemble Kálmán filter is based on a randomized linear equation system. The proof for the following statement is completely analogous to the one provided for Theorem 4 and it is found in [13].
Proposition 2.
The state vector
where
The following equations summarize the analysis step of the ensemble Kálmán filter as they are well known in the field of data assimilation (see for instance [4], [13] or [14]). The ensemble Kálmán filter follows from Proposition 2 after estimating the mean and covariance of the prior
The update step can then be performed directly on the ensemble
The matrix
In the case of a non-linear observation operator
where
Remark 12.
Working with a nonlinear observation operator in ensemble Kálmán filtering comes with approximation errors, as the true nonlinear update is fitted with a Gaussian model. In complex high-dimensional problems of data assimilation, the ensemble Kálmán filter is still popular, as it is often the only way to perform approximate inference, and alternative techniques can only be applied to highly simplified versions of the original problem [23].
In this work the ensemble Kálmán filter is used to update the prior ensemble
5.3 The Prior
Due to the smoothing property of the boundary integral equation, the inverse problem is ill-conditioned. To ensure a numerically stable solution we derive a regularization term which is based on a prior field solution. It is now shown how mean
From (2.5) we derive the randomized indirect Neumann problem:
The process noise
due to the symmetry of
Equation (5.15) provides a Gaussian prior
The regularization parameter δ controls the impact of the prior. Since
On the other hand, the prior state vector
The parameter δ has the same role as the ridge parameter in the Tikhonov regularization. Different approaches have been developed for its selection (see for instance [1, Chapter 1]). In the following we make use of a set of validation measurements, providing us with the vector
where
In Figure 8 (right), the validation error
Scenario | Moves | M | Measurement Resolution
|
Measurement Duration | Regularization Parameter δ |
1 | 281 | 4093 |
|
|
|
2 | 562 | 16324 |
|
|
|
5.4 Design of Experiment
Design of experiment is about determining the minimum amount of measurements required to meet the accuracy requirements. As we have seen in the previous section, the number of measurements, their spatial distribution and also the prior
Using the procedure developed in the previous sections, we are capable to propagate the measurement uncertainty down to beam related quantities by particle tracking. In the following, we consider the trajectory of a proton with momentum
This trajectory is illustrated in Figure 9, which shows the solution of (2.13) for a reference particle using an explicit, fourth-order Runge–Kutta scheme for the numerical integration (see [20, Chapter 2]).
Remark 13.
The explicit Runge–Kutta scheme is not symplectic, meaning that the Hamiltonian H, which is an invariant of motion, is not preserved by the numerical integration scheme. With the expression for the magnetic vector potential by means of (2.4), it is possible to use symplectic numerical integrators for multi-turn particle simulations, which are preserving the Hamiltonian by design (see [20, Chapter 6]). A comprehensive study of different numerical integration schemes is found in [42]). The use of spherical harmonics in symplectic particle tracking has been presented in [2]. The chosen procedure serves to illustrate the potential of our method. More detailed analyses of the numerical integration are planned for future publications.
We have calculated the trajectory for
In contrast to the posterior covariance matrix, the validation error cannot be determined a-priori. It mainly origins from locations in the fringe fields, where the spatial resolution of
The initial ensemble is updated move-by-move, using the equations of the ensemble Kálmán filter with
Table 3 gives the resulting validation errors. By means of the ensemble Kálmán updates, the validation error was reduced by a factor 2.5. As a side effect, more measurement data increases the precision in the prediction of the particle trajectory. This is shown in Figure 9, where we have added the trajectories evaluated from the posterior ensembles
Prior
|
Initialization
|
Updates Left
|
Updates Right
|
|
ϵ in
|
37.01 | 5.04 | 3.86 | 2.01 |
Equivalent B in
|
7.40 | 1.00 | 0.77 | 0.401 |
6 Conclusion
We have developed an approach that improves the performance of Hall probe field mapping in several aspects: The measurement duration to obtain a three-dimensional field map could be reduced from
To achieve these goals we have used an iso-geometric boundary-element method based on an indirect double-layer potential formulation. In terms of applicability, we emphasized the scalability of the algorithms and presented a method for efficient evaluation of the integral equation along the reference trajectory.
For the inference of boundary data from magnetic measurements we have made use of ensemble Kálmán filter, which provides us with an efficient compression technique of the posterior covariance matrix. This yields a significant reduction in computational complexity.
Funding statement: The work of Melvin Liebsch is supported by the Graduate School CE within the Centre for Computational Engineering at Technische Universität Darmstadt.
Acknowledgements
The authors thank Thomas Zickler for providing the OPERA model of the ELENA bending dipole magnet and Dimitrios Loukrezis and Lucio Fiscarelli for reviewing and discussing this article.
References
[1] J. M. Bardsley, Computational Uncertainty Quantification for Inverse Problems, Comput. Sci. Eng. 19, Society for Industrial and Applied Mathematics, Philadelphia, 2018. 10.1137/1.9781611975383Search in Google Scholar
[2] L. Bojtár, Efficient evaluation of arbitrary static electromagnetic fields with applications for symplectic particle tracking, Nucl. Instrum. Methods Phys. Res. A 948 (2019), Article ID 162841. 10.1016/j.nima.2019.162841Search in Google Scholar
[3] A. Buffa, J. Dölz, S. Kurz, S. Schöps, R. Vázquez and F. Wolf, Multipatch approximation of the de Rham sequence and its traces in isogeometric analysis, Numer. Math. 144 (2020), no. 1, 201–236. 10.1007/s00211-019-01079-xSearch in Google Scholar
[4] A. Carrassi, M. Bocquet, L. Bertino and G. Evensen, Data assimilation in the geosciences - an overview on methods, issues and perspectives, WIREs Climate Change 9 (2017), 10.1002/wcc.535. 10.1002/wcc.535Search in Google Scholar
[5] B. Dalena, O. Gabouev, J. Payet, Antoine Chance, D. R. Brett, R. B. Appleby, R. DeMaria and M. Giovannozzi, Fringe fields modeling for the high luminosity LHC large aperture quadrupoles, Proceedings of the 5th International Particle Accelerator Conference (IPAC 2014), JACoW Publishing, Geneva (2014), 993–996. Search in Google Scholar
[6] J. Dölz, H. Harbrecht, C. Jerez-Hanckes and M. Multerer, Isogeometric multilevel quadrature for forward and inverse random acoustic scattering, Comput. Methods Appl. Mech. Engrg. 388 (2022), Paper No. 114242. 10.1016/j.cma.2021.114242Search in Google Scholar
[7] J. Dölz, H. Harbrecht, S. Kurz, M. Multerer, S. Schöps and F. Wolf, Bembel: Boundary element method based engineering library, 2022. Search in Google Scholar
[8]
J. Dölz, H. Harbrecht and M. Peters,
[9] J. Dölz, H. Harbrecht and M. Peters, An interpolation-based fast multipole method for higher-order boundary elements on parametric surfaces, Internat. J. Numer. Methods Engrg. 108 (2016), no. 13, 1705–1728. 10.1002/nme.5274Search in Google Scholar
[10] A. Dragt, T. J. Stasevich and P. Walstrom, Computation of charged-particle transfer maps for general fields and geometries using electromagnetic boundary-value data, Proceedings of the 2001 Particle Accelerator Conference (PACS2001), IEEE Press, Piscataway (2001), 1776–1777. Search in Google Scholar
[11] A. J. Dragt, F. Neri, G. Rangarajan, D. R. Douglas, L. M. Healy and R. D. Ryne, Lie algebraic treatment of linear and nonlinear beam dynamics, Ann. Rev. Nuclear Particle Sci. 38 (1988), no. 1, 455–496. 10.1146/annurev.ns.38.120188.002323Search in Google Scholar
[12] A. Ern and J.-L. Guermond, Theory and Practice of Finite Elements, Appl. Math. Sci. 159, Springer, New York, 2013. Search in Google Scholar
[13] G. Evensen, The ensemble Kalman filter: Theoretical formulation and practical implementation, Ocean Dynam. 53 (2003), no. 4, 343–367. 10.1007/s10236-003-0036-9Search in Google Scholar
[14] G. Evensen, Spurious correlations, localization, and inflation, Data Assimilation, Springer, Berlin (2009), 237–253. 10.1007/978-3-642-03711-5_15Search in Google Scholar
[15] P. Förster, S. Schöps, J. Enders, M. Herbert and A. Simona, Freeform shape optimization of a compact dc photoelectron gun using isogeometric analysis, Phys. Rev. Accel. Beams 25 (2022), Article ID 034601. 10.1103/PhysRevAccelBeams.25.034601Search in Google Scholar
[16] R. Furrer and T. Bengtsson, Estimation of high-dimensional prior and posterior covariance matrices in Kalman filter variants, J. Multivariate Anal. 98 (2007), no. 2, 227–255. 10.1016/j.jmva.2006.08.003Search in Google Scholar
[17] L. Greengard and V. Rokhlin, A new version of the fast multipole method for the Laplace equation in three dimensions, Acta Numer. 6 (1997), 229–269. 10.1017/S0962492900002725Search in Google Scholar
[18] G. Guennebaud, B. Jacob, Eigen v3, http://eigen.tuxfamily.org, 2010. Search in Google Scholar
[19] E. Hairer, Geometric Numerical Integration: Structure-Preserving Algorithms for Ordinary Differential Equations, Springer, Berlin, 2006. Search in Google Scholar
[20] E. Hairer, C. Lubich and G. Wanner, Geometric Numerical Integration: Structure-Preserving Algorithms for Ordinary Differential Equations, 2nd ed., Springer Ser. Comput. Math. 31, Springer, Berlin, 2006. Search in Google Scholar
[21] I. G. Ion, M. Liebsch, A. Simona, D. Loukrezis, C. Petrone, S. Russenschuck, H. De Gersem and S. Schöps, Local field reconstruction from rotating coil measurements in particle accelerator magnets, Nucl. Instrum. Methods Phys. Res. A 1011 (2021), Article ID 165580. 10.1016/j.nima.2021.165580Search in Google Scholar
[22] I. G. Ion, C. Wildner, D. Loukrezis, H. Koeppl and H. De Gersem, Tensor-train approximation of the chemical master equation and its application for parameter inference, J. Chem. Phys. 155 (2021), no. 3, Article ID 034102. 10.1063/5.0045521Search in Google Scholar PubMed
[23] M. Katzfuss, J. R. Stroud and C. K. Wikle, Understanding the ensemble Kalman filter, Amer. Statist. 70 (2016), no. 4, 350–357. 10.1080/00031305.2016.1141709Search in Google Scholar
[24] H. R. Künsch, State space and hidden Markov models, Complex Stochastic Systems (Eindhoven 1999), Monogr. Statist. Appl. Probab. 87, Chapman & Hall/CRC, Boca Raton (2001), 109–173. Search in Google Scholar
[25] S. Kurz, H. De Gersem, A. Galetzka, A. Klaedtke, M. Liebsch, D. Loukrezis, S. Russenschuck and M. Schmidt, Hybrid modeling: Towards the next level of scientific computing in engineering, J. Math. Indust. 12 (2022), Article ID 8. 10.1186/s13362-022-00123-0Search in Google Scholar
[26] S. Kurz, D. Pauly, D. Praetorius, S. Repin and D. Sebastian, Functional a posteriori error estimates for boundary element methods, Numer. Math. 147 (2021), no. 4, 937–966. 10.1007/s00211-021-01188-6Search in Google Scholar
[27] F. Le Gland, V. Monbet and V.-D. Tran, Large sample asymptotics for the ensemble Kalman filter, Research Report RR-7014, INRIA, 2009. Search in Google Scholar
[28] G. Lehner and S. Kurz, Electromagnetic Field Theory for Engineers and Physicists, Springer, Berlin, 2010. 10.1007/978-3-540-76306-2Search in Google Scholar
[29] Y. Liu, Fast Multipole Boundary Element Method: Theory and Applications in Engineering, Cambridge University, Cambridge, 2009. 10.1017/CBO9780511605345Search in Google Scholar
[30] J. Mandel, Efficient implementation of the ensemble Kalman filter, UCDHSC/CCM Report no. 231, University of Colorado at Denver and Health Sciences Center, 2006. Search in Google Scholar
[31] R. Nertens, U. Pahner, K. Hameyer, R. Belmans and R. De Weerdt, Force calculation based on a local solution of laplace’s equation, IEEE Trans. Magnetics 33 (1997), no. 2, 1216–1218. 10.1109/20.582472Search in Google Scholar
[32] W. Oelert, The ELENA project at CERN, Acta Phys. Polonica B 46 (2015), Article ID 181. 10.5506/APhysPolB.46.181Search in Google Scholar
[33] E. Ott, B. R. Hunt, I. Szunyogh, A. V. Zimin, E. J. Kostelich, M. Corazza, E. Kalnay, D. J. Patil and J. A. Yorke, A local ensemble Kálmán filter for atmospheric data assimilation, Tellus A Dyn. Meteorol. Oceanography 56 (2004), no. 5, 415–428. 10.1111/j.1600-0870.2004.00076.xSearch in Google Scholar
[34] Y. Papaphilippou, J. Wei and R. Talman, Deflections in magnet fringe fields, Phys. Rev. E 67 (2003), Article ID 046502. 10.1103/PhysRevE.67.046502Search in Google Scholar PubMed
[35] S. Rjasanow and O. Steinbach, The Fast Solution of Boundary Integral Equations, Springer, New York, 2007. Search in Google Scholar
[36] S. Russenschuck, Field Computation for Accelerator Magnets, Wiley-VCH, Weinheim, 2010. 10.1002/9783527635467Search in Google Scholar
[37] S. Russenschuck, Rotating- and translating-coil magnetometers for extracting pseudo-multipoles in accelerator magnets, COMPEL 36 (2017), no. 5, 1552–1567. 10.1108/COMPEL-02-2017-0059Search in Google Scholar
[38] S. A. Sauter and C. Schwab, Boundary Element Methods, Springer Ser. Comput. Math. 39, Springer, Berlin, 2011. 10.1007/978-3-540-68093-2Search in Google Scholar
[39] D. Schoerling, Design study: ELENA bending magnet prototype, Report CERN-ACC-2013-0261, CERN, Geneva, 2013. Search in Google Scholar
[40] J. Sherman and W. J. Morrison, Adjustment of an inverse matrix corresponding to a change in one element of a given matrix, Ann. Math. Statistics 21 (1950), 124–127. 10.1214/aoms/1177729893Search in Google Scholar
[41] A. Simona, Numerical methods for the simulation of particle motion in electromagnetic fields, PhD thesis, Politecnico di Milano, Milano, 2020. Search in Google Scholar
[42] A. Simona, L. Bonaventura, T. Pugnat and B. Dalena, High order time integrators for the simulation of charged particle motion in magnetic quadrupoles, Comput. Phys. Commun. 239 (2019), 33–52. 10.1016/j.cpc.2019.01.018Search in Google Scholar
[43] S. Sorti, C. Petrone, S. Russenschuck and F. Braghin, A magneto-mechanical model for rotating-coil magnetometers, Nucl. Instrum. Methods Phys. Res. A 984 (2020), Article ID 164599. 10.1016/j.nima.2020.164599Search in Google Scholar
[44] M. Spink, D. Claxton, C. de Falco and R. Vázquez, The NURBS toolbox, http://octave.sourceforge.net/nurbs/index.html. Search in Google Scholar
[45] O. Steinbach, Numerical Approximation Methods for Elliptic Boundary Value Problems, Springer, New York, 2008. 10.1007/978-0-387-68805-3Search in Google Scholar
[46] M. Venturini, D. Abell and A. Dragt, Map computation from magnetic field data and application to the LHC high-gradient quadrupoles, 1999. Search in Google Scholar
[47] A. Wolski, Beam Dynamics in High Energy Particle Accelerators, Imperial College, London, 2014. 10.1142/p899Search in Google Scholar
[48] Y. K. Wu, E. Forest and D. Robin, Explicit symplectic integrator of s-dependent static magnetic field, Phys. Rev. E 68 (2003), Article ID 046502. 10.1103/PhysRevE.68.046502Search in Google Scholar PubMed
[49] International Organization for Standardization, Guide to the expression of uncertainty in measurement (GUM)-Supplement 1: Numerical methods for the propagation of distributions, volume ISO draft guide DGUIDE99998, International Organization for Standardization, Geneva, 2004. Search in Google Scholar
[50] NIST Digital Library of Mathematical Functions, http://dlmf.nist.gov/, Release 1.1.3 of 2021-09-15, F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, B. V. Saunders, H. S. Cohl, and M. A. McClain. Search in Google Scholar
© 2023 Walter de Gruyter GmbH, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.