Warning, /acts/docs/tracking.md is written in an unsupported language. File is not indexed.
0001 # Tracking in a nutshell
0002
0003 Track reconstruction is the process to recover the properties of a charged
0004 particle from a set of measurements caused by the interaction with some form of
0005 sensitive detector. The goal is to find which measurements are likely to have
0006 been caused by which particle, to group them accordingly, and to estimate the
0007 associated trajectory. Such charged particle trajectories form the basic input
0008 to the majority of higher-level reconstruction procedures in many cases.
0009
0010 :::{figure} /figures/tracking/tracking.svg
0011 :width: 400px
0012 :align: center
0013 Illustration of a track reconstruction chain starting from spacepoints to fully
0014 formed tracks.
0015 :::
0016
0017 This section provides a high-level view of a track reconstruction chain, and is
0018 largely based on [^phd:gessinger:2021]. It gives an overview of the basic
0019 building blocks conceptually, and also connects these building blocks with the
0020 concrete implementations in the core ACTS library, where available.
0021
0022 ## Charged particle detection
0023
0024 The first step in the chain to reconstruct charged particles is their
0025 detection using sensitive elements. Charged particle detection can be
0026 achieved in a variety of ways with very different technologies. It can be
0027 achieved by measuring the interaction of charged particles with matter. When
0028 this occurs, the interacting particle typically ionizes the surrounding
0029 material. Particle detectors make use of this fact by converting the
0030 resulting charge into a measurable signal in various ways.
0031
0032 (segmentation)=
0033 :::{figure} /figures/tracking/segmentation.svg
0034 :align: center
0035 :width: 400px
0036
0037 Illustration of a one-dimensional (a) and a two-dimensional segmentation (b) of a silicon sensor.
0038 :::
0039
0040 A very common electronic detection approach is the use of semiconducting
0041 particle detectors, often made of silicon. When a charged particle traverses
0042 such a sensor, it ionizes the material in the depletion zone, caused by the
0043 interface of two different semiconducting materials. The result are pairs of
0044 opposite charges. These charge pairs are separated by an electric field and
0045 drift toward the electrodes. At this point, an electric signal is created
0046 which can be amplified and read out. By means of segmentation, the measured
0047 signal can be associated with a location on the sensor. Silicon sensors are
0048 usually segmented in one dimension (*strips*) or in two dimensions
0049 (*pixels*) (see {numref}`segmentation`).
0050
0051 (track_parametrization)=
0052 ## Track parametrization
0053
0054 To express the properties of a particle's trajectory, a choice of parameters
0055 has to be made. The parameters need to express all the relevant quantities of
0056 interest. In the presence of a magnetic field, which affects charged
0057 trajectories, the global position and momentum, as well as the charge are
0058 needed to fully specify the particle properties. In addition, a time parameter
0059 can be included. Apart from the global reference frame, track quantities often
0060 need to be represented with respect to a surface. This can be achieved with a
0061 parametrization like
0062
0063 \begin{equation*}
0064 \vec x = \left(l_0, l_1, \phi, \theta, q/p, t\right)^T
0065 \end{equation*}
0066
0067 although other parameter conventions exist as well.
0068 {numref}`parameters` illustrates this choice of parameters. $l_0$, $l_1$ are
0069 the local coordinates of the corresponding surface, $\phi \in [-\pi,\pi)$ and
0070 $\theta \in [0,\pi]$ are the angles in the transverse and longitudinal
0071 direction of the global frame, expressed with respect to the current location
0072 along the trajectory, as indicated in {numref}`parameters` (b). $\theta$ is
0073 the polar angle relative to the positive $z$-axis, and $\phi$ is the azimuth
0074 angle in the transverse plane. Finally, $q/p$ combines the charge of the
0075 particle with the inverse momentum. In {numref}`parameters` (a), the global
0076 momentum vector $\vec p$ is shown, which can be recovered from the parameters
0077 $\vec x$ using $\phi$, $\theta$ and $q/p$.
0078
0079 (parameters)=
0080 :::{figure} /figures/tracking/parameters.svg
0081 :width: 500px
0082 :align: center
0083 Illustration of the parametrization of a particle track with respect to a
0084 two-dimensional surface. (a) shows the local position, global momentum and
0085 their corresponding uncertainties. (b) displays the angles $\phi$ and $\theta$
0086 in the transverse and longitudinal planes.
0087 :::
0088
0089 (perigee)=
0090 :::{figure} /figures/tracking/perigee.svg
0091 :width: 200px
0092 :align: center
0093 Illustration of the perigee parametrization which uses the point of closest
0094 approach relative to a reference point. The impact parameter $d_0$, the
0095 position $l$ and the momentum vector $\vec p$ are shown.
0096 :::
0097
0098 Aside from the nominal quantities captured in $\vec x$, the related
0099 uncertainties and correlations need to be taken into account as well. They
0100 can be expressed as a $6\times 6$ covariance matrix like
0101
0102 \begin{equation*}
0103 C =
0104 \begin{bmatrix}
0105 \sigma^2(l_0)& \text{cov}(l_0,l_1) & \text{cov}(l_0, \phi) & \text{cov}(l_0, \theta) & \text{cov}(l_0, q/p) & \text{cov}(l_0, t) \\
0106 . & \sigma^2(l_1) & \text{cov}(l_1, \phi) & \text{cov}(l_1, \theta) & \text{cov}(l_1, q/p) & \text{cov}(l_1, t) \\
0107 . & . & \sigma^2(\phi) & \text{cov}(\phi,\theta) & \text{cov}(\phi, q/p) & \text{cov}(\phi, t) \\
0108 . & . & . & \sigma^2(\theta) & \text{cov}(\theta, q/p) & \text{cov}(\theta, t) \\
0109 . & . & . & . & \sigma^2(q/p) & \text{cov}(q/p, t) \\
0110 . & . & . & . & . & \sigma^2(t)
0111 \end{bmatrix}
0112 \end{equation*}
0113
0114 Here, $\text{cov}(X,Y)$ is the covariance of variables $X$ and $Y$, while
0115 $\sigma^2(X)$ are the regular variances. As the covariance matrix $C$ is
0116 symmetric, only the upper right half is shown in the matrix above. The
0117 uncertainties associated with the local position, as well as the momentum
0118 direction are indicated in {numref}`parameters` (a) as an ellipse and a cone
0119 around the momentum vector $\vec p$, respectively.
0120
0121 (particle_propagation)=
0122 ## Particle propagation
0123
0124 :::{tip}
0125 A dedicated description of the ACTS implementation of particle propagation can be found [here](propagation_impl).
0126 :::
0127
0128 A central part of track reconstruction is the ability to calculate the trajectory
0129 of a charged particle, given its properties at a given point. This process,
0130 called *particle propagation* or *extrapolation*, is used to predict a
0131 particle's properties after it has travelled a certain distance. In many cases,
0132 the projected intersection with various types of surfaces is desired. The
0133 trajectory of a charged particle is governed by the [magnetic
0134 field](#magnetic-field-core) through which it travels, as well as any
0135 [material effects](#material-core). In case of a homogeneous magnetic field,
0136 and in the absence of material interaction, the particle follows a helical
0137 trajectory. Such a helix can be calculated purely analytically, although
0138 intersections require numerical methods nevertheless.
0139
0140 Often, magnetic fields are not homogeneous, however. In the presence of such
0141 changing fields, the corresponding differential equations of motions need to be
0142 solved using numerical integration techniques.
0143
0144 ### Numerical integration
0145
0146 In ACTS, numerical integration is done using the *Runge-Kutta-Nyström* (RKN) method.
0147 Commonly used in the variant at fourth order, the RKN method describes how to calculate a
0148 solution to an initial value problem that can be formulated generically like
0149
0150 $$
0151 \frac{dy}{dt} = f(t,y), \qquad y(t_0) = y_0,
0152 $$
0153
0154 where $y_0$ refers to the initial value of $y$ at $t_0$, and
0155 $f(t,y)$ is the functional form describing the dynamics. The method then
0156 successively approximates the analytical solution $y(t)$ in a stepwise
0157 fashion. At each step $(t_n, y_n)$, the goal is effectively to approximate
0158 the next step $y(t_{n+1})$. Using a step size $h$, the algorithm evaluates
0159 the function $f$ at four points $k_{1-4}$:
0160
0161 $$
0162 \begin{aligned}
0163 k_1 &= f(t_n, y_n) \\
0164 k_2 &= f\left( t_n + \frac h 2, y_n + h \frac{k_1} 2 \right) \\
0165 k_3 &= f\left( t_n + \frac h 2, y_n + h \frac{k_2} 2 \right)\\
0166 k_4 &= f\left( t_n + h, y_n + hk_3 \right).
0167 \end{aligned}
0168 $$
0169
0170
0171 (rk)=
0172 :::{figure} /figures/tracking/rk.svg
0173 :width: 400px
0174 :align: center
0175 Illustration of the RKN method approximating a first order
0176 differential equation. Shown is the calculation of an estimate $y_{n+1}$ at
0177 $t_{n+1} = t_n + h$, based on the current step $(t_n,y_n)$. Shown are the four
0178 distinct points at which function $y(t)$ is evaluated, and which are blended to
0179 form the estimate.
0180 :::
0181
0182
0183 Looking at {numref}`rk`, the meaning of these four points in relation
0184 to the step size $h$ can be understood. $k_1$ is the derivative at the
0185 current location, $k_{2,3}$ use $k_1$ and $k_2$ respectively to calculate two
0186 envelope derivatives at $h/2$ and $k_4$ uses $k_3$ to make an estimate of the
0187 derivative at $h$. Combining $k_{1-4}$, $(t_{n+1},y_{n+1})$ can be calculated
0188 as the approximation of $y(t_{n+1})$ like
0189
0190 $$
0191 \begin{aligned}
0192 y_{n+1} &= y_n + \frac 1 6 h ( k_1 + 2 k_2 + 2 k_2 + k_4)\\
0193 t_{n+1} &= t_n + h
0194 \end{aligned}
0195 $$
0196
0197 by effectively averaging the four derivatives. It is apparent that
0198 the step size crucially influences the accuracy of the approximation. A large
0199 step size weakens the approximation, especially if the magnetic field changes
0200 strongly. On the other hand, a too small step size will negatively affect the
0201 execution time of the algorithm.
0202
0203 The Runge-Kutta-Nyström method from above can be adapted to handle second order
0204 differential equations, as is needed for the equations of motion in question,
0205
0206 $$
0207 \frac{d^2 \vec r}{ds^2} = \frac q p \left( \frac{d\vec r}{ds} \times \vec B (\vec r) \right) = f(s, \vec r, \vec T), \qquad \vec T \equiv \frac{d \vec r}{ds},
0208 $$
0209
0210 with the global position $\vec r$, the path element $s$, the
0211 normalized tangent vector $\vec T$ and the magnetic field $\vec B(\vec r)$ at
0212 the global position. A slight modification of $k_{1-4}$ is also required,
0213 incorporating the first derivative of $f(s, \vec r, \vec r')$, finally
0214 leading to
0215
0216 $$
0217 \begin{aligned}
0218 \vec T_{n+1} &= \vec T_n + \frac h 6 (k_1 + 2k_2 + 2k_3 + k_4) \\
0219 \vec r_{n+1} &= \vec r_n + h \vec T_n + \frac{h^2}{6} (k_1 + k_2 + k_3).
0220 \end{aligned}
0221 $$
0222
0223 A strategy exists to dynamically adapt the step size according to the magnetic
0224 field strength, with the definition of a target accuracy that the algorithm
0225 tries to achieve. Here, the step size $h$ will successively be decreased and
0226 the approximation recalculated until the accuracy goal is achieved. Even with
0227 these additional calculations, the approach is still preferable over a
0228 consistently low step size.
0229
0230 ### Covariance transport
0231
0232 Aside from the prediction of the track parameters at a given path length, a key
0233 ingredient to many dependent applications are the uncertainties in the form of
0234 the associated covariance matrix $C$. Conversions between covariance matrices
0235 $C^i\to C^f$ can generally be achieved like
0236
0237 $$
0238 C^f = J \cdot C^i \cdot J^T,
0239 $$
0240
0241 using the Jacobian matrix
0242
0243 $$
0244 J = \begin{bmatrix}
0245 \frac{\partial l_0^f}{\partial l_0^i} & \cdots & \frac{\partial l_0^f}{\partial (q/p)^i} \\
0246 \vdots & \ddots & \vdots \\
0247 \frac{\partial (q/p)^f}{\partial l_0^i} & \cdots & \frac{\partial (q/p)^f}{\partial (q/p)^i}
0248 \end{bmatrix},
0249 $$
0250
0251 between initial and final parameters $\vec x^i$ and $\vec x^f$. The
0252 task therefore becomes calculating the necessary Jacobians to achieve correct
0253 transformation.
0254
0255 One part is the transformation between different coordinate systems, but at the
0256 same location along the trajectory. For this purpose, generic Jacobians can be
0257 calculated between each coordinate system type, and a common coordinate system.
0258 The common coordinate system used for this purpose is the curvilinear frame,
0259 which consists of the global direction angles, and a plane surface located at
0260 the track location, with the normal of the plane surface aligned with the track
0261 momentum. By using Jacobians to the curvilinear frame and the corresponding
0262 inverse matrices, conversions between any two coordinate systems can be
0263 performed.
0264
0265 The second part is the calculation of the evolution of the covariance matrix
0266 during the propagation between surfaces. To this end, a semi-analytical
0267 method which calculates the effective derivatives between two consecutive
0268 RKN steps can be used. By accumulating the Jacobian
0269 matrices calculated for each step, the effective Jacobian between the
0270 starting point and the destination can be obtained.
0271
0272 ## Material effects
0273
0274 Charged particles interact with matter as they pass through it. Since
0275 particle detectors inevitably consist of some form or material, this effect
0276 cannot be completely avoided. By building tracking detectors as light as
0277 possible, and arranging passive components, such as services and support
0278 structures carefully, the material a particle encounters before being
0279 measured can be reduced. Charged particles traversing any form of matter
0280 undergo elastic and inelastic interactions with the atomic structure of the
0281 material, depending on the particle properties.
0282
0283 (multiple_scattering)=
0284 :::{figure} /figures/tracking/multiple_scattering.svg
0285 :width: 200px
0286 :align: center
0287 Illustration of the effect of multiple scattering on the
0288 trajectory of a charged particle passing through a block of material.
0289 Entering from the left, it undergoes a series of scattering events,
0290 deflecting the trajectory statistically, before exiting on the right.
0291 :::
0292
0293 In elastic interactions, the particle does not lose a significant amount of
0294 energy, while its trajectory is affected. {numref}`multiple_scattering` shows a
0295 sketch of the way multiple Coulomb scattering affects the direction of a
0296 particle trajectory. In addition, a shift in the transverse plane relative to
0297 the incident direction can occur. As the scattering events occur in
0298 statistically independent directions, the means of both the deflection and
0299 offset tends toward zero as the number of scatters increases. Therefore, in the
0300 numerical particle propagation, this can be accounted for by simply increasing
0301 the uncertainties associated with the direction, depending on the amount of
0302 material encountered.
0303
0304 On the other hand, there are interactions during which the particle loses some
0305 of its energy. Relevant processes here are ionization, as well as
0306 bremsstrahlung for light particles like electrons. For hadronic particles,
0307 hadronic interactions with the nuclei of surrounding material is another
0308 process of interest. In such hadronic interactions, the incoming particle often
0309 disintegrates, and does not propagate further. Since the size of ionization
0310 losses only fluctuates to a small degree for thin layers of material, they can
0311 usually be accounted for by reducing the trajectory energy correspondingly. For
0312 bremsstrahlung, where fluctuations are much larger and the effect cannot be
0313 modelled adequately with a Gaussian distribution, dedicated techniques are
0314 needed (see [](gsf_core)).
0315
0316 Two main approaches are implemented in ACTS. The first approximates the
0317 material interaction by using a description that averages the real material
0318 onto thin surfaces across the detector (more on this in
0319 [](#geometry-and-material-modelling)). When the propagation encounters such a
0320 surface, it retrieves the material properties, and executes parametrized
0321 modifications to the particle properties and uncertainties. In the second
0322 approach, material effects are continuously incorporated during propagation,
0323 rather than at discrete locations. The latter approach is especially suited for
0324 propagation through volumes of dense material, where the discretization of the
0325 material distribution will not work as well.
0326
0327 ## Geometry and material modelling
0328
0329 :::{tip}
0330 A dedicated description of the ACTS implementation of the tracking geometry model can be found [here](geometry_impl).
0331 :::
0332
0333 A detailed model of the geometry of an experiment is required for tracking. In
0334 many cases, external information is needed to associate a sensitive element
0335 with a position and rotation in the laboratory frame. In case of silicon
0336 sensors, the intrinsic information captured by the sensor is restricted to the
0337 measurement plane. Using a transformation matrix, this local measurement can be
0338 turned into a global one.
0339
0340 Full simulation using tools like Geant4 are frequently used in HEP.
0341 It includes its own geometry
0342 description framework. For the precise simulation of particle interactions
0343 with the detector, this geometry modelling
0344 is highly detailed. Even very small details of the
0345 physical hardware can be crucial, and are often included in the geometry
0346 description. An example for this are readout chips on silicon sensors, or cooling
0347 elements. {numref}`geometry_detail` (a) shows a sketch of such a detailed
0348 geometry description. Shown as an example is a *layer* of silicon
0349 sensors in a barrel configuration. The green rectangles represent the actual
0350 sensitive surfaces, while other elements include cooling, readout and other components.
0351
0352 (geometry_detail)=
0353 :::{figure} /figures/tracking/geometry_detail.svg
0354 :align: center
0355
0356 Sketch of the way a fully detailed simulation geometry (a) models passive
0357 elements, in addition to the sensitive elements shown in green. (b) shows a
0358 simplified version, where all non-sensitive elements are approximated.
0359 :::
0360
0361 In the majority of cases in track reconstruction, this detailed
0362 geometry is unnecessary. During track reconstruction, the aforementioned
0363 associated information needs to be accessible for measurements, so all
0364 sensitive elements need to be included in some form. Passive elements, on the
0365 other hand, are only required to factor in material interaction effects (see
0366 [](#particle-propagation)). Moreover, the fully detailed geometry comes
0367 at the disadvantage of introducing significant overhead during navigation. In
0368 this process, an algorithm attempts to figure out which elements the particle
0369 propagation needs to target, as the trajectory is likely to intersect them.
0370 With a geometry description this precise, the navigation process becomes a
0371 significant performance bottleneck.
0372
0373 (layer_barrel)=
0374 :::{figure} /figures/tracking/layer_barrel.svg
0375 :width: 300px
0376 :align: center
0377 Sketch of the way sensitive elements are grouped into layers.
0378 Shown is an $xy$-view of a number of sensors, arranged as in e.g. the ATLAS
0379 silicon detector barrels. The grouping is based on their mounting radius.
0380 The layers are indicated in different colors.
0381 :::
0382
0383 As a compromise between modelling accuracy and performance, ACTS
0384 uses a simplified geometry model. It
0385 focusses on the sensitive elements, which are strictly needed, while passive
0386 elements are discarded from the explicit description and approximated.
0387 {numref}`geometry_detail` (b) shows such a simplified geometry. Here, the
0388 sensitive elements are still shown in green, and other elements are greyed
0389 out, indicating that they are discarded. The sensitive elements are then
0390 grouped into layers, as sketched in {numref}`layer_barrel`. How exactly the
0391 grouping occurs depends on the concrete experiment geometry. In some cases, the layers have the shape
0392 of cylinder surfaces with increasing radii. This example is shown in the
0393 figure in the transverse plane at radii $r_{1,2,3}$. In the endcaps, where
0394 modules are often arranged on disks, these are used as the layer shape. An
0395 illustration of endcap disk layers can be found in {numref}`layer_ec`, where
0396 six disks are located at six distinct positions in $\pm z_{1,2,3}$, and
0397 shown in different colors.
0398
0399 (layer_ec)=
0400 :::{figure} /figures/tracking/layer_ec.svg
0401 :width: 400px
0402 :align: center
0403 Sketch of the way sensitive elements are grouped into layers.
0404 Shown is a view of a number of sensors, arranged as in e.g. the ATLAS
0405 silicon detector endcaps. They are grouped into disks based on their
0406 mounting position in $z$. The layers are indicated in different colors.
0407 :::
0408
0409 During particle propagation, the navigation makes use of this layer
0410 system. Each layer contains a binned structure, which maps a bin to a set
0411 of sensitive surfaces that overlap with the bin area. This is illustrated in
0412 {numref}`geo_binning`, where the left picture shows the sensitive surface
0413 structure of an exemplary endcap disk. The picture on the right overlays the
0414 binning structure that can be used to enable fast retrieval of compatible
0415 sensitive surfaces. By performing a simple bin lookup, the navigation can
0416 ascertain which sensors it needs to attempt propagation to.
0417
0418
0419 (geo_binning)=
0420 :::{figure} /figures/tracking/surface_array.svg
0421 :width: 400px
0422 :align: center
0423 Illustration of the binning structure that is used to subdivide
0424 layer surfaces. (a) shows two sensor rings of
0425 different radii grouped into one disk layer. (b)
0426 overlays the binning structure that the navigation queries for compatible
0427 surfaces.
0428 :::
0429
0430 Furthermore, layers are grouped into volumes. Each volume loosely corresponds
0431 to a region of the detector.
0432 Volumes are set up such that their boundary surfaces always touch another
0433 volume. An exception to this is the outermost volume. Each volume's boundary
0434 surfaces store which volume is located on their other side, essentially
0435 forming portals between the volumes. This glueing enables the geometry
0436 navigation between volumes. When the propagation has finished processing a
0437 set of layers, it attempts to target the boundary surfaces. Once a boundary
0438 surface is reached, the active volume is switched, and the next set of layers
0439 is processed.
0440
0441 Care has to be taken to correctly model the passive material, that is
0442 initially discarded with non-sensitive elements. For the material effects to
0443 be correctly taken into account during particle propagation, the material is
0444 projected onto dedicated material surfaces. These material surfaces are
0445 spread across the detector geometry. Each layer is created with two
0446 *approach surfaces* on either side. Their distance can be interpreted as
0447 the thickness of the layer in question. Examples of these approach surfaces
0448 can be found in {numref}`geometry_detail`, at the inner and outer radius.
0449 Approach surfaces, and the boundary surfaces between volumes mentioned before,
0450 are candidates to receive a projection of the surrounding material.
0451 Additional artificial material layers can also be inserted to receive
0452 projected material.
0453
0454 The projection procedure (see [](#material-core) and [](#material_mapping_howto_core)) works
0455 by extrapolating test particles using the fully detailed simulation geometry.
0456 During the extrapolation, the material properties of the geometry are sampled
0457 in small intervals. Subsequently, the same test particle is extrapolated
0458 through the tracking geometry. All material samples are then assigned and
0459 projected onto the closest material surface. Finally, the projection is
0460 averaged. The exact number and placement of the material surfaces has to be
0461 optimized to yield a sufficiently accurate representation of the inactive
0462 material in the detector.
0463
0464 The numerical integration uses these projected material surfaces. Whenever
0465 such a surface is encountered in the propagation, the material properties are
0466 retrieved, and the corresponding modifications to the trajectory are
0467 executed. In case material is supposed to be integrated in a continuous way
0468 (as mentioned in [](#particle-propagation)), volumes can also store an
0469 effective volumetric material composition, which is queried by the numerical
0470 integration when needed. As the actual physical location of the detection
0471 hardware can vary over time, possible misalignment of the sensors needs to be
0472 handled correctly.
0473
0474 ## Clustering
0475
0476 :::{tip}
0477 See [](clustering_core) for information of the implementation of clustering in
0478 the core library.
0479 :::
0480
0481 The actual track reconstruction procedure itself starts with the conversion of
0482 raw inputs that have been read out from the detector. In case of silicon
0483 detectors, the readout can either be performed in a binary way, only recording
0484 which segments fired, or the amount of charges measured in the segment can be
0485 recorded, e.g. via *time-over-threshold* readout. In all cases, the readout is
0486 attached to an identifier uniquely locating the segment on the corresponding
0487 sensor.
0488
0489 As a next step, these raw readouts need to be *clustered*, in order to
0490 extract an estimate of where particles intersect with the sensor. The general
0491 strategy of clustering algorithms follows the Connected Component Analysis (CCA)
0492 approach, where subsets of segments are successively grouped into clusters.
0493 In case of the Pixel detector, this clustering occurs in two dimensions,
0494 corresponding to the segmentation of its sensors. Here, the CCA can
0495 either consider all eight surrounding pixels as neighboring a central one, or
0496 only consider the four non-diagonal ones, as shown in
0497 {numref}`clustering_cca`. The figure only shows the simplest possible
0498 cluster starting from the central pixel. In reality, the CCA will iteratively
0499 continue from the pixels on the cluster edges.
0500
0501 (clustering_cca)=
0502 :::{figure} /figures/tracking/cca.svg
0503 :align: center
0504 :width: 400px
0505 Illustration of both eight and four cell connectivity.
0506 :::
0507
0508 Subsequently, the effective cluster position needs to be estimated. Multiple
0509 factors play a role here. First of all, the average position of the cluster
0510 can be calculated either using only the geometry position of the segments,
0511
0512 $$
0513 \vec r = \frac{1}{N} \sum_{i=1}^N \vec l_i,
0514 $$
0515
0516 or be weighted by the charge collected in each segment:
0517
0518 $$
0519 \vec r = \frac{1}{\sum_{i=1}^N q_i} \sum_{i=1}^N q_i \vec l_i.
0520 $$
0521
0522 Here, $\vec l_i$ is the local position of the $i$-th segment while
0523 $q_i$ is its charge.
0524
0525 An illustration of the clusterization can be found in {numref}`clustering_image`,
0526 where a pixel sensor is shown to be intersected by a charged particle,
0527 entering on the lower left and exiting on the top right. Three cells shown
0528 with a red frame receive energy from the particle, but the amount is under
0529 the readout threshold. Four other cells receive energy above the threshold
0530 and are read out. The clustering will then group these four cells into a
0531 cluster, and subsequently estimate the cluster position based on the energy
0532 deposited in the cells. In case no charge information is not available
0533 for a given detector, the calculation is purely geometric.
0534
0535
0536 (clustering_image)=
0537 :::{figure} /figures/tracking/clustering.svg
0538 :align: center
0539 :width: 400px
0540 Illustration of the clustering of multiple pixels into a cluster,
0541 in a three-dimensional view on the left and a projection onto the
0542 $xy$-plane on the right. A particle enters the center in the lower left,
0543 crosses several segments before exiting the sensor on the top right. The
0544 cell colors indicate how far along the trajectory they are encountered.
0545 :::
0546
0547 Another factor that needs to be accounted for is the drift direction of the
0548 created charges. In addition to the collection field of the sensor itself,
0549 the surrounding magnetic field modifies the drift direction by the
0550 *Lorentz-angle* $\theta_\text{L}$. Depending on the field strength, this
0551 additional angle can cause segments to be activated that would otherwise not
0552 be geometrically within reach of the charges. Other effects, such as the fact
0553 that the modules are not perfectly flat, as the geometry description assumes,
0554 or cross-talk between readout channels, also play a role at this stage.
0555
0556 In the presence of high event activity, particle intersections on single
0557 sensors can be close enough to one another to result in clusters that are not
0558 clearly separated from each other. This circumstance can be somewhat
0559 mitigated by allowing tracks to share clusters with other particles, which
0560 comes at the price of allowing duplicated tracks to some extent.
0561 Additionally, merged clusters typically feature worse position resolution,
0562 which manifests itself since it negatively affects the final fit of the
0563 track.
0564
0565 (tracking_sp_formation)=
0566 ## Spacepoint formation
0567
0568 The basic input to most forms of pattern recognition algorithms for tracking
0569 are space points, which need to be assembled from the raw measurements. To this
0570 end, the raw measurements are combined with information provided by the
0571 geometry description, such as the location and rotation of the sensors. In this
0572 way, the locations, which are restricted to be local to the sensor surfaces
0573 intrinsically, can be converted into three dimensional points in space. See
0574 [](spacepoint_formation_core) for a description of the implementation of
0575 spacepoint formation in the core library.
0576
0577 {numref}`sensor` shows an illustration of the information that is consumed for
0578 a pixel measurement. Shown are three clusters on a sensor, which are caused by
0579 three tracks intersecting it. The corresponding cluster positions are indicated
0580 as well, and can be converted to global positions using the inverse of the
0581 global-to-local transformation matrix, that is provided by the geometry
0582 description.
0583
0584 (sensor)=
0585 :::{figure} /figures/tracking/sp_l2g.svg
0586 :align: center
0587 :width: 400px
0588 Illustration of a pixel sensor and its local coordinate system in
0589 relation to the global laboratory frame. A transformation allows conversion
0590 between the two systems. Shown are three tracks intersecting the sensor,
0591 alongside clusters that they produce.
0592 :::
0593
0594 In strip detectors, on the other hand, only a single
0595 dimension is segmented, and an individual measurement is therefore only
0596 constrained in one direction on the surface. However, usually the
0597 strip modules are mounted in pairs, with a stereo angle rotation
0598 between the pairs. To form global space points, measurements from both
0599 sensors of a pair need to be combined.
0600 Due to the stereo angle, a two dimensional
0601 location on the orthogonal projection plane relative to the two parallel
0602 pairs can be found. Using the global transformations of the pair, the
0603 combined measurement location can be converted to global coordinates. If
0604 multiple measurements are located on a stereo pair of strip sensors, there
0605 exists an ambiguity on how to combine strips to form space points, which has to be resolved.
0606
0607
0608 ## Seeding
0609
0610 The next step after space point formation is pattern recognition, which be
0611 implemented in various ways. Global methods exist which attempt to cluster
0612 space points, such as conformal mapping. In this approach, the space points are
0613 transformed into a feature parameter space that reveals patterns for hits
0614 belonging to the same track. In the specific example of a Hough transform, a
0615 parameter space $\left(\phi, q/p_\mathrm{T}\right)$ is used. As a result, each
0616 space point is effectively transformed into a line, as a series of combinations
0617 of these parameters would lead to the same space point. The lines from a set of
0618 space points of a single track will intersect in one common area. Such an
0619 intersection can be used to identify which space points originate from the same
0620 track. However, this task grows in complexity as detector activity increases
0621 and is susceptible to material effects. See [](seeding_core) for a description
0622 of the seeding implementation in the core library.
0623
0624 Another group of approaches is the one of seeding and track following. These
0625 algorithms differ from the global ones in that they evaluate individual
0626 combinations of space points, and successively explore the events. One
0627 algorithm from this group is the cellular automaton that iteratively forms
0628 chains of space points going from one layer to the next.
0629
0630 The main approach in ACTS is an algorithm that operates on coarse
0631 subdivisions of the detector is used. This seeding algorithm attempts to find
0632 triplets of space points from increasing radii which are likely to belong to
0633 the same track. It achieves this by iterating the combinatorial triplets and
0634 successively filtering them. Filtering is performed based on the momentum and
0635 impact parameters, which the algorithm attempts to estimate for each triplet.
0636
0637 Under the assumption of a homogeneous magnetic field along the $z$-axis,
0638 charged particles should follow helical trajectories. In the transverse plane,
0639 the motion is circular, while it is a straight line in the $rz$-plane. The
0640 transverse impact parameter and momentum can be estimated from the radius of
0641 the circle in the transverse plane like
0642
0643 $$
0644 d_0 = \sqrt{c_x^2 + c_y^2} - \rho,
0645 $$
0646
0647 with the circle center $(c_x, c_y)$ and radius $\rho$. The
0648 transverse momentum can be related to available quantities like
0649
0650 $$
0651 p_\mathrm{T} \propto \cdot q B \rho
0652 $$
0653
0654 with the charge $q$ and the magnetic field $B$. An intersection
0655 between the straight line in the $rz$-plane with the $z$-axis gives an
0656 estimate of the longitudinal impact parameter.
0657 An illustration of seeds in the transverse plane is found in
0658 {numref}`seeding_figure`. Note that seeds can incorporate hits spread across all of
0659 the layers shown, although this can be a configuration parameter.
0660
0661
0662 (seeding_figure)=
0663 :::{figure} /figures/tracking/seeding.svg
0664 :width: 300px
0665 :align: center
0666 Sketch of seeds in the transverse plane for a number of tracks on
0667 four layers. Seeds can combine hits on any three of these layers. The shown
0668 seeds appear compatible with having originated in the center of the
0669 detector, which is also drawn.
0670 :::
0671
0672 ## Track finding and track fitting
0673
0674 In the track seeding and following approach, track candidates are built from
0675 the initial seeds. One method implemented in ACTS, the [Combinatorial Kalman
0676 Filter](#combinatorial-kalman-filter) (CKF), uses the *Kalman formalism*.
0677 Originally developed for monitoring and steering mechanical systems, it can
0678 also be used to iteratively calculate a track estimate. After a set of track
0679 candidates has been assembled and filtered (see [](#ambiguity-resolution)), an
0680 additional track fit is usually performed to extract the best estimate of the
0681 track. The Kalman formalism can also be used for this, with the addition of a
0682 smoothing step that has certain benefits. Other fit strategies exist, such as
0683 a global $\chi^2$ fit that minimizes the distances between track-sensor
0684 intersections and measurements on all sensors at the same time. One drawback of
0685 this method is the necessity to invert very large matrices, which is
0686 computationally expensive.
0687
0688 In a track fit, the Kalman formalism can be shown to yield optimal estimates
0689 for Gaussian uncertainties. This assumption breaks down when effects like
0690 bremsstrahlung come into play. An extension of the Kalman Filter (KF) exists
0691 that relies on the individual propagation of a set of trajectories, instead of
0692 a single one, to model these biased uncertainties by a sum of Gaussian
0693 components. This [Gaussian Sum Filter](gsf_core) (GSF) achieves better results when
0694 fitting particles such as electrons, likely to undergo bremsstrahlung, and is
0695 deployed in e.g. the ATLAS tracking chain.
0696
0697
0698 ### Kalman formalism and Kalman track fitter
0699
0700 :::{tip}
0701 See [](kf_core) for a description of the implementation of the Kalman Filter in
0702 the core library.
0703 :::
0704
0705 The basis of the Kalman formalism is a state vector, that can be identified
0706 with the set of track parameters $\vec x$. Note that the concrete
0707 parametrization plays a subordinate role in this context. Rather than building
0708 an estimate of the state of a system in real time, a Kalman track fit can be
0709 understood as estimating the parameters iteratively in steps. In the track
0710 fitting application, each step is defined by a measurement to be included.
0711 The evolution of the state vector is described by
0712
0713 $$
0714 \vec x_k = \mathbf F_{k-1} \vec x_{k-1} + \vec w_{k-1},
0715 $$
0716
0717 where the linear function $\mathbf F_{k-1}$ transports the state vector at
0718 step $k-1$ to step $k$. $\vec w_{k-1}$ is additional so-called process noise
0719 that affects the transport additively. Each step has an associated
0720 measurement, with the fixed relationship between the measurement and the state vector
0721
0722 $$
0723 \vec m_k = \mathbf H_k \vec x_k + \epsilon_k.
0724 $$
0725
0726 Here, $\mathbf H_k$ is the *measurement mapping function*, which
0727 transforms the state vector into the measurement space. In the ideal case,
0728 this purpose can be achieved by a simple projection matrix, which extracts a
0729 subspace of the state vector. Additionally, $\epsilon_k$ represents the
0730 measurement uncertainty.
0731
0732 The Kalman fit process is divided into different phases:
0733
0734 1. **Prediction** of the state vector at the next step $k+1$ based on the information at the current step $k$.
0735 2. **Filtering** of the prediction by incorporating the measurement associated to the step
0736 3. **Smoothing** of the state vector by walking back the steps and using information for the subsequent step $k+1$ to improve the estimate at the current step $k$.
0737
0738 An illustration of these concepts is found in {numref}`kalman_filter`. Here,
0739 a series of three sensors is shown with measurements on them. The KF
0740 then predicts the track parameters at an intersection, shown in blue.
0741 Subsequently, a filtered set of parameters is calculated as a mixture between
0742 the measurement and the prediction. Not shown in this picture is the
0743 smoothing step.
0744
0745 (kalman_filter)=
0746 :::{figure} /figures/tracking/kalman.svg
0747 :width: 400px
0748 :align: center
0749 Illustration of the KF. Two of the three stages, the prediction and the
0750 filtering are shown. The filtering updates the prediction with information from
0751 the measurement.
0752 :::
0753
0754
0755 In many cases, the first two phases run in tandem, with prediction and
0756 filtering happening alternatingly at each step. The smoothing phase is
0757 launched once the last measurement has been encountered.
0758 Starting from a state $k$, first, a prediction of the state vector at the
0759 next measurement location is obtained via
0760
0761 (kf_pred)=
0762 $$
0763 \vec x_k^{k-1} = \mathbf F_{k-1} \vec x_{k-1},
0764 $$
0765
0766 with the linear transport function from above. $\vec x_k^{k-1}$ is
0767 the prediction of the state vector at step $k$ based on step $k-1$. The next
0768 stage is the filtering. Here, the state vector is refined by taking into
0769 account the measurement at the current step. Following one of two variants of
0770 filtering from [^Fruhwirth:1987fm], the gain matrix formalism, the state
0771 vector is updated like
0772
0773 $$
0774 \vec x_k = \vec x_k^{k-1} + \mathbf K_k \left( \vec m_k - \mathbf H_k \vec x_k^{k-1} \right),
0775 $$
0776
0777 with the *Kalman gain matrix*
0778
0779 $$
0780 \mathbf K_k = \mathbf C_k^{k-1} \mathbf H_k^\mathrm{T}
0781 \left(
0782 \mathbf V_k + \mathbf H_k \mathbf C_k^{k-1} \mathbf H_k^\mathrm{T}
0783 \right)^{-1}
0784 .
0785 $$
0786
0787 Note that $\vec x_k$ is the filtered state vector at step $k$,
0788 based on information from previous steps and step $k$ itself. This is in
0789 contrast to $\vec x_k^{k-1}$, which is the prediction of the state vector at
0790 step $k$ based on $k-1$, and is used to calculate the filtered state vector.
0791 One input to these equations is the covariance matrix prediction $\mathbf
0792 C_k^{k-1}$ at step $k$ based on step $k-1$, which can be written as
0793
0794 $$
0795 \mathbf C_k^{k-1} = \mathbf F_{k-1} \mathbf C_{k-1} \mathbf F_{k-1}^\mathrm{T} + \mathbf Q_{k-1}
0796 $$
0797
0798 in the linear version from [^Fruhwirth:1987fm], with the
0799 covariance $\mathbf C_{k-1}$ at step $k-1$, and the covariance $\mathbf
0800 Q_{k-1}$ associated with $\vec w_{k-1}$ from above. Also needed is $\mathbf
0801 V_k$, which is the covariance associated with $\epsilon_k$, effectively
0802 representing the measurement uncertainty.
0803
0804 Similar to the state vector itself, the corresponding covariance matrix is
0805 also filtered using
0806
0807 (kf_cov_pred)=
0808 $$
0809 \mathbf C_k = \left( \mathbb 1 - \mathbf K_k \mathbf H_k \right) \mathbf C_k^{k-1}.
0810 $$
0811
0812 In the smoothing phase, the state vector at step $k$ is improved using the
0813 information from the subsequent step $k+1$ using
0814
0815 $$
0816 \vec x_k^n = \vec x_k + \mathbf A_k \left( \vec x_{k+1}^n - \vec x_{k+1}^k \right).
0817 $$
0818
0819 Here, $\vec x_{k+1}^n$ is the smoothed state vector and $\vec
0820 x_{k+1}^k$ the predicted state vector at the subsequent step $k+1$. Also
0821 needed is the *smoother gain matrix*
0822
0823 $$
0824 \mathbf A_k = \mathbf C_k \mathbf F_k^\mathrm{T} \left( \mathbf C^k_{k+1} \right)^{-1},
0825 $$
0826
0827 with the predicted covariance at step $k+1$, $\mathbf C_k^{k+1}$.
0828 Finally, the covariance at the current step $k$ can also be smoothed with
0829
0830 $$
0831 \mathbf C_k^n = \mathbf C_k + \mathbf A_k \left(\mathbf C_{k+1}^n - \mathbf C_{k+1}^k \right) \mathbf A_k^\mathrm{T}.
0832 $$
0833
0834 After smoothing, the parameter estimate at the first step contains information
0835 from all other measurement states. As mentioned above, in case the
0836 uncertainties entering the Kalman fit are Gaussian distributions without
0837 biases, the KF can be shown to be the optimal solution minimizing mean
0838 square estimation error. However, certain caveats exist. The KF assumes
0839 that a linear transport function $\mathbf F$ exists that can propagate the
0840 state vector. In the presence of inhomogeneous magnetic fields this is not
0841 the case. Instead of explicitly applying $\mathbf F$ to the state vector for
0842 the prediction, the ACTS KF turns to the numerical integration,
0843 discussed in [](#numerical-integration). With it, the prediction from
0844 [this equation](kf_pred) is simply the intersection of the extrapolated trajectory
0845 with the next sensitive surface. Aside from this, $\mathbf F$ is also used to
0846 transport the covariance between steps (see [here](kf_cov_pred)). Here, the
0847 semi-analytical method for covariance transport in the numerical integration
0848 can be used. $\mathbf F$ can then be identified with the transport
0849 Jacobian accumulated between surfaces.
0850
0851 For smoothing, two possibilities exist to obtain the needed covariances from
0852 subsequent measurement steps. Either, the inverse transport Jacobian is used
0853 and applied, in a way similar to [this equation](kf_cov_pred), or the numerical
0854 integration is executed again in an inverse fashion, propagating from the
0855 subsequent step to the current one.
0856
0857 ### Combinatorial Kalman Filter
0858
0859 :::{tip}
0860 See [](ckf_core) for information on the CKF implementation found in the core
0861 library.
0862 :::
0863
0864 As mentioned above, the Kalman formalism can be used for track finding. In this
0865 case, the smoothing step can be dropped, as the resulting track candidates are
0866 likely to be refit regardless, therefore saving some time. The CKF explores the
0867 event starting from an initial track seed. It does this by considering not only
0868 a single sequence of measurements, but allowing the branching of the fit at
0869 each sensitive surface that is encountered. To this end, all or a subset of
0870 measurements that are found on each surface are considered. Measurements are
0871 selected based on their compatibility with the current state estimate, by using
0872 their residuals. A predicted residual
0873
0874 $$
0875 \vec r_k^{k-1} = \vec m_k - \mathbf H_k \vec x_k^{k-1},
0876 $$
0877
0878 and a filtered residual
0879
0880 $$
0881 \vec r_k = \vec m_k - \mathbf H_k \vec x_k
0882 ,
0883 $$
0884
0885 can be defined, depending on which state estimate is compared with
0886 the measurement $\vec m_k$. Using the filtered residual, an effective
0887 $\chi^2$ increment
0888
0889 $$
0890 \chi^2_+ = \vec r_k^\mathrm{T}
0891 \left[ \left( \mathbb 1 - \mathbf H_k \mathbf K_k \right) \mathbf V_k \right]^{-1}
0892 \vec r_k
0893 $$
0894
0895 can be calculated. The global $\chi^2$ of the track candidate can
0896 be calculated as the sum of all $\chi^2_+$ across the steps. Measurements
0897 with a large $\chi^2_+$ are considered as outliers, which have low
0898 compatibility with the trajectory. By branching out for measurements below a
0899 certain $\chi^2_+$, and following the branches, a tree-like structure of
0900 compatible track candidates originating from a track seed is assembled. This
0901 feature is shown in {numref}`tracking_ckf`, which displays a circular
0902 trajectory, and a set of iteratively assembled track candidates. Basic
0903 quality criteria can be applied at this stage, to remove bad candidates. A
0904 dedicated [](#ambiguity-resolution).
0905 selects the candidates most likely to belong to real particle tracks.
0906
0907 (tracking_ckf)=
0908 :::{figure} /figures/tracking/finding.svg
0909 :width: 300px
0910 :align: center
0911 Illustration of the way the CKF iteratively explores
0912 measurements from a seed outwards. Measurements are added successively, and
0913 can be shared between the resulting track candidates. Shown in green is a
0914 circular *real* trajectory.
0915 :::
0916
0917 ## Ambiguity resolution
0918
0919 Due to the combinatorial nature of track finding, and to achieve high
0920 efficiencies, this set of candidates is often large, and contains a
0921 non-negligible fraction of *fake* candidates. These fake candidates are either
0922 completely combinatorial, or arise from real particle measurements with
0923 combinatorial additions. Track candidates coming from a single seed necessarily
0924 share a common stem of measurements. Measurements can potentially also be
0925 shared between candidates from different seeds.
0926
0927 One possibility to resolve this (as is done in e.g. ATLAS tracking) is an
0928 ambiguity resolution algorithm, that attempts to filter out as many undesirable
0929 tracks as possible. This is implemented by means of a scoring function, that
0930 combines properties of the track parameters. Higher scores are correlated with
0931 a larger probability to be a desirable track candidate. A larger number of hits
0932 results in an increase in the score, as longer compatible hit chains are less
0933 likely to be random combinations. On the other hand, missing hits in sensors
0934 where a hit was expected negatively impact the score. Experiment specific
0935 scoring of hits from different subsystems is also implemented. The overall
0936 $\chi^2$ value computed for the track candidate also plays a role. Candidates
0937 that share hits with other candidates are penalized. Another quantity is the
0938 measured particle $p_\mathrm{Y}$, which enters the score, to give preference to
0939 tracks with large momenta. For tracks containing measurements with a
0940 substantial local $\chi^2_+$ at the start or end of the trajectory, the
0941 ambiguity resolution step can also attempt to remove these hits, and determine
0942 whether a refit without them yields a more favorable global $\chi^2$.
0943
0944 Finally, the output of the ambiguity resolution step is a set of track candidates
0945 that contain an enhanced fraction of tracks from actual particles, while fake
0946 tracks are suppressed. They are passed into the final precision fit outlined
0947 in [](#track-finding-and-track-fitting), to extract the parameter estimate, and used
0948 for further aspects of reconstruction.
0949
0950 ## Vertex reconstruction
0951
0952 :::{tip}
0953 See [](vertexing_core) for a dedicated description of the vertexing as
0954 implemented in ACTS.
0955 :::
0956
0957 A vertex is a point within the detector, where an interaction or a
0958 decay occurred. We distinguish between primary vertices (from
0959 collisions/interactions) and secondary vertices (from subsequent particle
0960 decays), see {numref}`vertexing_illust`. Primary vertices are further divided
0961 into hard-scatter and pile-up vertices. While primary vertices are located in
0962 the luminous region, secondary vertices are slightly displaced due to the finite
0963 life time of the decaying particle.
0964
0965 (vertexing_illust)=
0966 :::{figure} /figures/tracking/vertexing.svg
0967 :width: 400px
0968 :align: center
0969 Illustration of a set of three vertices in a proton-proton
0970 collision. We distinguish between primary hard-scatter, primary pile-up, and
0971 secondary vertices.
0972 :::
0973
0974 Vertices play an important role in higher-level reconstruction algorithms. For
0975 example, secondary vertices can help with the identification of particles:
0976 During *$b$-tagging*, a displaced vertex located inside a jet is a sign for the
0977 decay of a $b$-hadron.
0978
0979 In analogy to track reconstruction, vertex reconstruction can be divided into
0980 two stages: vertex finding and vertex fitting. As a first step of vertex
0981 finding, we compute a rough estimate of the vertex position from a set of
0982 tracks. This first estimate can be calculated in many different ways, and is
0983 referred to as "vertex seed". Seeding algorithms differ for primary and
0984 secondary vertexing. For primary vertex seeding, one option is to use a
0985 histogram approach to cluster tracks on the $z$-axis[^phd:piacquadio:2010].
0986 This is based on the assumption that primary vertices will be close to the
0987 beamline. Other approaches model tracks as multivariate Gaussian distributions
0988 and identify regions of high track density as vertex seeds[^phd:schlag:2022].
0989 For secondary vertexing, seeds are formed from pairs of reconstructed tracks as
0990 the constraint to the beamline does not apply.
0991
0992 Once a vertex seed is determined, tracks that are compatible with it are
0993 selected as part of the vertex finding.
0994
0995 Before the vertex fit, we linearize tracks in the vicinity of the vertex seed
0996 under assuming that they follow a helical (for constant magnetic field) or
0997 straight (for no magnetic field) trajectory[^phd:piacquadio:2010]. The vertex
0998 fitter then uses this linearization to improve the position of the vertex seed.
0999 Furthermore, the track momenta are refitted under the assumption that the tracks
1000 originate at the vertex[^Fruhwirth:1987fm] [^billoirfitting:1992] .
1001
1002 One issue with an approach like this is that the assignment of tracks to
1003 vertices is ambiguous. As an improvement, one can perform a multi-vertex fit,
1004 where vertices compete for tracks. This means that one track can be assigned to
1005 several vertices. Their contribution to each vertex fit is determined by a
1006 weight factor, which, in turn, depends on the tracks' compatibility with respect
1007 to all vertices[^fruwirth:amvfitting:2004].
1008
1009 A flowchart of a multi-vertex reconstruction chain is shown in
1010 {numref}`vertexing_flowchart`.
1011
1012 (vertexing_flowchart)=
1013 :::{figure} /figures/tracking/vertexing_flowchart.svg
1014 :width: 600px
1015 :align: center
1016 Simplified flowchart of multi-vertex reconstruction. From a set of seed tracks,
1017 we first compute a rough estimate of the vertex position, i.e., the vertex seed.
1018 Then, we evaluate the compatibility of all tracks with the the latter. If a
1019 track is deemed compatible, it is assigned a weight and attached to the vertex
1020 seed. Next, the vertex seed and all previously found vertices that share tracks
1021 with it are (re-)fitted. Finally, after convergence of the fit, we check whether
1022 the vertex candidate is merged with other vertices and discard it if that is the
1023 case. For the next iteration, all tracks that were assigned to the vertex seed
1024 and that have a weight above a certain threshold are removed from the seed
1025 tracks.
1026 :::
1027
1028 [^phd:gessinger:2021]: Gessinger-Befurt, Paul, 30.04.2021. Development and improvement of track reconstruction software and search for disappearing tracks with the ATLAS experiment. [10.25358/openscience-5901](https://doi.org/10.25358/openscience-5901)
1029 [^Fruhwirth:1987fm]: R. Frühwirth, 1987, Application of Kalman filtering to track and vertex fitting, , [11.1016/0168-9002(87)90887-4](https://doi.org/10.1016/0168-9002(87)90887-4)
1030 [^phd:piacquadio:2010]: G. Piacquadio, 2010, Identification of b-jets and investigation of the discovery potential of a Higgs boson in the $W H \rightarrow l \nu \bar{b} b$ channel with the ATLAS experiment.
1031 [^phd:schlag:2022]: B. Schlag, 2022, Advanced Algorithms and Software for Primary Vertex Reconstruction and Search for Flavor-Violating Supersymmetry with the ATLAS Experiment.
1032 [^billoirfitting:1992]: P. Billoir et al., 2022, Fast vertex fitting with a local parametrization of tracks.
1033 [^fruwirth:amvfitting:2004]: R. Frühwirth et al., 2004, Adaptive Multi-Vertex fitting.