Back to home page

sPhenix code displayed by LXR

 
 

    


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 $5\times 5$ 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) \\
0106    . & \sigma^2(l_1) & \text{cov}(l_1, \phi) & \text{cov}(l_1, \theta) & \text{cov}(l_1, q/p) \\
0107    . & . &  \sigma^2(\phi) & \text{cov}(\phi,\theta) & \text{cov}(\phi, q/p) \\
0108    . & . & . & \sigma^2(\theta) & \text{cov}(\theta, q/p) \\
0109    . & . & . & . & \sigma^2(q/p)
0110   \end{bmatrix}
0111 \end{equation*}
0112 
0113 Here, $\text{cov}(X,Y)$ is the covariance of variables $X$ and $Y$, while
0114 $\sigma^2(X)$ are the regular variances. As the covariance matrix $C$ is
0115 symmetric, only the upper right half is shown in the matrix above. The
0116 uncertainties associated with the local position, as well as the momentum
0117 direction are indicated in {numref}`parameters` (a) as an ellipse and a cone
0118 around the momentum vector $\vec p$, respectively.
0119 
0120 (particle_propagation)=
0121 ## Particle propagation
0122 
0123 :::{tip}
0124 A dedicated description of the ACTS implementation of particle propagation can be found [here](propagation_impl).
0125 :::
0126 
0127 A central part of track reconstruction is the ability to calculate the trajectory
0128 of a charged particle, given its properties at a given point. This process,
0129 called *particle propagation* or *extrapolation*, is used to predict a
0130 particle's properties after it has travelled a certain distance. In many cases,
0131 the projected intersection with various types of surfaces is desired. The
0132 trajectory of a charged particle is governed by the [magnetic
0133 field](#magnetic-field-core) through which it travels, as well as any
0134 [material effects](#material-core). In case of a homogeneous magnetic field,
0135 and in the absence of material interaction, the particle follows a helical
0136 trajectory. Such a helix can be calculated purely analytically, although
0137 intersections require numerical methods nevertheless.
0138 
0139 Often, magnetic fields are not homogeneous, however. In the presence of such
0140 changing fields, the corresponding differential equations of motions need to be
0141 solved using numerical integration techniques.
0142 
0143 ### Numerical integration
0144 
0145 In ACTS, numerical integration is done using the *Runge-Kutta-Nyström* (RKN) method.
0146 Commonly used in the variant at fourth order, the RKN method describes how to calculate a
0147 solution to an initial value problem that can be formulated generically like
0148 
0149 $$
0150   \frac{dy}{dt} = f(t,y), \qquad y(t_0) = y_0,
0151 $$
0152 
0153  where $y_0$ refers to the initial value of $y$ at $t_0$, and
0154 $f(t,y)$ is the functional form describing the dynamics. The method then
0155 successively approximates the analytical solution $y(t)$ in a stepwise
0156 fashion. At each step $(t_n, y_n)$, the goal is effectively to approximate
0157 the next step $y(t_{n+1})$. Using a step size $h$, the algorithm evaluates
0158 the function $f$ at four points $k_{1-4}$:
0159 
0160 $$
0161   \begin{aligned}
0162     k_1 &= f(t_n, y_n) \\
0163     k_2 &= f\left( t_n + \frac h 2, y_n + h \frac{k_1} 2 \right) \\
0164     k_3 &= f\left( t_n + \frac h 2, y_n + h \frac{k_2} 2 \right)\\
0165     k_4 &= f\left( t_n + h, y_n + hk_3 \right).
0166   \end{aligned}
0167 $$
0168 
0169 
0170 (rk)=
0171 :::{figure} /figures/tracking/rk.svg
0172 :width: 400px
0173 :align: center
0174 Illustration of the RKN method approximating a first order
0175 differential equation. Shown is the calculation of an estimate $y_{n+1}$ at
0176 $t_{n+1} = t_n + h$, based on the current step $(t_n,y_n)$. Shown are the four
0177 distinct points at which function $y(t)$ is evaluated, and which are blended to
0178 form the estimate.
0179 :::
0180 
0181 
0182 Looking at {numref}`rk`, the meaning of these four points in relation
0183 to the step size $h$ can be understood. $k_1$ is the derivative at the
0184 current location, $k_{2,3}$ use $k_1$ and $k_2$ respectively to calculate two
0185 envelope derivatives at $h/2$ and $k_4$ uses $k_3$ to make an estimate of the
0186 derivative at $h$. Combining $k_{1-4}$, $(t_{n+1},y_{n+1})$ can be calculated
0187 as the approximation of $y(t_{n+1})$ like
0188 
0189 $$
0190   \begin{aligned}
0191     y_{n+1} &= y_n + \frac 1 6 h ( k_1 + 2 k_2 + 2 k_2 + k_4)\\
0192     t_{n+1} &= t_n + h
0193   \end{aligned}
0194 $$
0195 
0196 by effectively averaging the four derivatives. It is apparent that
0197 the step size crucially influences the accuracy of the approximation. A large
0198 step size weakens the approximation, especially if the magnetic field changes
0199 strongly. On the other hand, a too small step size will negatively affect the
0200 execution time of the algorithm.
0201 
0202 The Runge-Kutta-Nyström method from above can be adapted to handle second order
0203 differential equations, as is needed for the equations of motion in question,
0204 
0205 $$
0206   \frac{d^2 \vec r}{ds^2} = \frac q p \left( \frac{d\vec r}{ds} \times \vec B (\vec r) \right) = f(s, \vec r, \vec T), \qquad \vec T \equiv \frac{d \vec r}{ds},
0207 $$
0208 
0209 with the global position $\vec r$, the path element $s$, the
0210 normalized tangent vector $\vec T$ and the magnetic field $\vec B(\vec r)$ at
0211 the global position. A slight modification of $k_{1-4}$ is also required,
0212 incorporating the first derivative of $f(s, \vec r, \vec r')$, finally
0213 leading to
0214 
0215 $$
0216   \begin{aligned}
0217     \vec T_{n+1} &= \vec T_n + \frac h 6 (k_1 + 2k_2 + 2k_3 + k_4) \\
0218     \vec r_{n+1} &= \vec r_n + h \vec T_n + \frac{h^2}{6} (k_1 + k_2 + k_3).
0219   \end{aligned}
0220 $$
0221 
0222 A strategy exists to dynamically adapt the step size according to the magnetic
0223 field strength, with the definition of a target accuracy that the algorithm
0224 tries to achieve. Here, the step size $h$ will successively be decreased and
0225 the approximation recalculated until the accuracy goal is achieved. Even with
0226 these additional calculations, the approach is still preferable over a
0227 consistently low step size.
0228 
0229 ### Covariance transport
0230 
0231 Aside from the prediction of the track parameters at a given path length, a key
0232 ingredient to many dependent applications are the uncertainties in the form of
0233 the associated covariance matrix $C$. Conversions between covariance matrices
0234 $C^i\to C^f$ can generally be achieved like
0235 
0236 $$
0237   C^f = J \cdot C^i \cdot J^T,
0238 $$
0239 
0240 using the Jacobian matrix
0241 
0242 $$
0243   J = \begin{bmatrix}
0244     \frac{\partial l_0^f}{\partial l_0^i} & \cdots & \frac{\partial l_0^f}{\partial (q/p)^i} \\
0245     \vdots & \ddots & \vdots \\
0246     \frac{\partial (q/p)^f}{\partial l_0^i} & \cdots & \frac{\partial (q/p)^f}{\partial (q/p)^i}
0247   \end{bmatrix},
0248 $$
0249 
0250 between initial and final parameters $\vec x^i$ and $\vec x^f$. The
0251 task therefore becomes calculating the necessary Jacobians to achieve correct
0252 transformation.
0253 
0254 One part is the transformation between different coordinate systems, but at the
0255 same location along the trajectory. For this purpose, generic Jacobians can be
0256 calculated between each coordinate system type, and a common coordinate system.
0257 The common coordinate system used for this purpose is the curvilinear frame,
0258 which consists of the global direction angles, and a plane surface located at
0259 the track location, with the normal of the plane surface aligned with the track
0260 momentum. By using Jacobians to the curvilinear frame and the corresponding
0261 inverse matrices, conversions between any two coordinate systems can be
0262 performed.
0263 
0264 The second part is the calculation of the evolution of the covariance matrix
0265 during the propagation between surfaces. To this end, a semi-analytical
0266 method which calculates the effective derivatives between two consecutive
0267 RKN steps can be used. By accumulating the Jacobian
0268 matrices calculated for each step, the effective Jacobian between the
0269 starting point and the destination can be obtained.
0270 
0271 ## Material effects
0272 
0273 Charged particles interact with matter as they pass through it. Since
0274 particle detectors inevitably consist of some form or material, this effect
0275 cannot be completely avoided. By building tracking detectors as light as
0276 possible, and arranging passive components, such as services and support
0277 structures carefully, the material a particle encounters before being
0278 measured can be reduced. Charged particles traversing any form of matter
0279 undergo elastic and inelastic interactions with the atomic structure of the
0280 material, depending on the particle properties.
0281 
0282 (multiple_scattering)=
0283 :::{figure} /figures/tracking/multiple_scattering.svg
0284 :width: 200px
0285 :align: center
0286 Illustration of the effect of multiple scattering on the
0287 trajectory of a charged particle passing through a block of material.
0288 Entering from the left, it undergoes a series of scattering events,
0289 deflecting the trajectory statistically, before exiting on the right.
0290 :::
0291 
0292 In elastic interactions, the particle does not lose a significant amount of
0293 energy, while its trajectory is affected. {numref}`multiple_scattering` shows a
0294 sketch of the way multiple Coulomb scattering affects the direction of a
0295 particle trajectory. In addition, a shift in the transverse plane relative to
0296 the incident direction can occur. As the scattering events occur in
0297 statistically independent directions, the means of both the deflection and
0298 offset tends toward zero as the number of scatters increases. Therefore, in the
0299 numerical particle propagation, this can be accounted for by simply increasing
0300 the uncertainties associated with the direction, depending on the amount of
0301 material encountered.
0302 
0303 On the other hand, there are interactions during which the particle loses some
0304 of its energy. Relevant processes here are ionization, as well as
0305 bremsstrahlung for light particles like electrons. For hadronic particles,
0306 hadronic interactions with the nuclei of surrounding material is another
0307 process of interest. In such hadronic interactions, the incoming particle often
0308 disintegrates, and does not propagate further. Since the size of ionization
0309 losses only fluctuates to a small degree for thin layers of material, they can
0310 usually be accounted for by reducing the trajectory energy correspondingly. For
0311 bremsstrahlung, where fluctuations are much larger and the effect cannot be
0312 modelled adequately with a Gaussian distribution, dedicated techniques are
0313 needed (see [](gsf_core)).
0314 
0315 Two main approaches are implemented in ACTS. The first approximates the
0316 material interaction by using a description that averages the real material
0317 onto thin surfaces across the detector (more on this in
0318 [](#geometry-and-material-modelling)). When the propagation encounters such a
0319 surface, it retrieves the material properties, and executes parametrized
0320 modifications to the particle properties and uncertainties. In the second
0321 approach, material effects are continuously incorporated during propagation,
0322 rather than at discrete locations. The latter approach is especially suited for
0323 propagation through volumes of dense material, where the discretization of the
0324 material distribution will not work as well.
0325 
0326 ## Geometry and material modelling
0327 
0328 :::{tip}
0329 A dedicated description of the ACTS implementation of the tracking geometry model can be found [here](geometry_impl).
0330 :::
0331 
0332 A detailed model of the geometry of an experiment is required for tracking. In
0333 many cases, external information is needed to associate a sensitive element
0334 with a position and rotation in the laboratory frame. In case of silicon
0335 sensors, the intrinsic information captured by the sensor is restricted to the
0336 measurement plane. Using a transformation matrix, this local measurement can be
0337 turned into a global one.
0338 
0339 Full simulation using tools like Geant4 are frequently used in HEP.
0340 It includes its own geometry
0341 description framework. For the precise simulation of particle interactions
0342 with the detector, this geometry modelling
0343 is highly detailed. Even very small details of the
0344 physical hardware can be crucial, and are often included in the geometry
0345 description. An example for this are readout chips on silicon sensors, or cooling
0346 elements. {numref}`geometry_detail` (a) shows a sketch of such a detailed
0347 geometry description. Shown as an example is a *layer* of silicon
0348 sensors in a barrel configuration. The green rectangles represent the actual
0349 sensitive surfaces, while other elements include cooling, readout and other components.
0350 
0351 (geometry_detail)=
0352 :::{figure} /figures/tracking/geometry_detail.svg
0353 :align: center
0354 
0355 Sketch of the way a fully detailed simulation geometry (a) models passive
0356 elements, in addition to the sensitive elements shown in green. (b) shows a
0357 simplified version, where all non-sensitive elements are approximated.
0358 :::
0359 
0360 In the majority of cases in track reconstruction, this detailed
0361 geometry is unnecessary. During track reconstruction, the aforementioned
0362 associated information needs to be accessible for measurements, so all
0363 sensitive elements need to be included in some form. Passive elements, on the
0364 other hand, are only required to factor in material interaction effects (see
0365 [](#particle-propagation)). Moreover, the fully detailed geometry comes
0366 at the disadvantage of introducing significant overhead during navigation. In
0367 this process, an algorithm attempts to figure out which elements the particle
0368 propagation needs to target, as the trajectory is likely to intersect them.
0369 With a geometry description this precise, the navigation process becomes a
0370 significant performance bottleneck.
0371 
0372 (layer_barrel)=
0373 :::{figure} /figures/tracking/layer_barrel.svg
0374 :width: 300px
0375 :align: center
0376 Sketch of the way sensitive elements are grouped into layers.
0377 Shown is an $xy$-view of a number of sensors, arranged as in e.g. the ATLAS
0378 silicon detector barrels. The grouping is based on their mounting radius.
0379 The layers are indicated in different colors.
0380 :::
0381 
0382 As a compromise between modelling accuracy and performance, ACTS
0383 uses a simplified geometry model. It
0384 focusses on the sensitive elements, which are strictly needed, while passive
0385 elements are discarded from the explicit description and approximated.
0386 {numref}`geometry_detail` (b) shows such a simplified geometry. Here, the
0387 sensitive elements are still shown in green, and other elements are greyed
0388 out, indicating that they are discarded. The sensitive elements are then
0389 grouped into layers, as sketched in {numref}`layer_barrel`. How exactly the
0390 grouping occurs depends on the concrete experiment geometry. In some cases, the layers have the shape
0391 of cylinder surfaces with increasing radii. This example is shown in the
0392 figure in the transverse plane at radii $r_{1,2,3}$. In the endcaps, where
0393 modules are often arranged on disks, these are used as the layer shape. An
0394 illustration of endcap disk layers can be found in {numref}`layer_ec`, where
0395 six disks are located at six distinct positions in $\pm z_{1,2,3}$, and
0396 shown in different colors.
0397 
0398 (layer_ec)=
0399 :::{figure} /figures/tracking/layer_ec.svg
0400 :width: 400px
0401 :align: center
0402 Sketch of the way sensitive elements are grouped into layers.
0403 Shown is a view of a number of sensors, arranged as in e.g. the ATLAS
0404 silicon detector endcaps. They are grouped into disks based on their
0405 mounting position in $z$. The layers are indicated in different colors.
0406 :::
0407 
0408 During particle propagation, the navigation makes use of this layer
0409 system. Each layer contains a binned structure, which maps a bin to a set
0410 of sensitive surfaces that overlap with the bin area. This is illustrated in
0411 {numref}`geo_binning`, where the left picture shows the sensitive surface
0412 structure of an exemplary endcap disk. The picture on the right overlays the
0413 binning structure that can be used to enable fast retrieval of compatible
0414 sensitive surfaces. By performing a simple bin lookup, the navigation can
0415 ascertain which sensors it needs to attempt propagation to.
0416 
0417 
0418 (geo_binning)=
0419 :::{figure} /figures/tracking/surface_array.svg
0420 :width: 400px
0421 :align: center
0422 Illustration of the binning structure that is used to subdivide
0423 layer surfaces. (a) shows two sensor rings of
0424 different radii grouped into one disk layer. (b)
0425 overlays the binning structure that the navigation queries for compatible
0426 surfaces.
0427 :::
0428 
0429 Furthermore, layers are grouped into volumes. Each volume loosely corresponds
0430 to a region of the detector.
0431 Volumes are set up such that their boundary surfaces always touch another
0432 volume. An exception to this is the outermost volume. Each volume's boundary
0433 surfaces store which volume is located on their other side, essentially
0434 forming portals between the volumes. This glueing enables the geometry
0435 navigation between volumes. When the propagation has finished processing a
0436 set of layers, it attempts to target the boundary surfaces. Once a boundary
0437 surface is reached, the active volume is switched, and the next set of layers
0438 is processed.
0439 
0440 Care has to be taken to correctly model the passive material, that is
0441 initially discarded with non-sensitive elements. For the material effects to
0442 be correctly taken into account during particle propagation, the material is
0443 projected onto dedicated material surfaces. These material surfaces are
0444 spread across the detector geometry. Each layer is created with two
0445 *approach surfaces* on either side. Their distance can be interpreted as
0446 the thickness of the layer in question. Examples of these approach surfaces
0447 can be found in {numref}`geometry_detail`, at the inner and outer radius.
0448 Approach surfaces, and the boundary surfaces between volumes mentioned before,
0449 are candidates to receive a projection of the surrounding material.
0450 Additional artificial material layers can also be inserted to receive
0451 projected material.
0452 
0453 The projection procedure (see [](#material-core) and [](#material_mapping_howto_core)) works
0454 by extrapolating test particles using the fully detailed simulation geometry.
0455 During the extrapolation, the material properties of the geometry are sampled
0456 in small intervals. Subsequently, the same test particle is extrapolated
0457 through the tracking geometry. All material samples are then assigned and
0458 projected onto the closest material surface. Finally, the projection is
0459 averaged. The exact number and placement of the material surfaces has to be
0460 optimized to yield a sufficiently accurate representation of the inactive
0461 material in the detector.
0462 
0463 The numerical integration uses these projected material surfaces. Whenever
0464 such a surface is encountered in the propagation, the material properties are
0465 retrieved, and the corresponding modifications to the trajectory are
0466 executed. In case material is supposed to be integrated in a continuous way
0467 (as mentioned in [](#particle-propagation)), volumes can also store an
0468 effective volumetric material composition, which is queried by the numerical
0469 integration when needed. As the actual physical location of the detection
0470 hardware can vary over time, possible misalignment of the sensors needs to be
0471 handled correctly.
0472 
0473 ## Clustering
0474 
0475 :::{tip}
0476 See [](clustering_core) for information of the implementation of clustering in
0477 the core library.
0478 :::
0479 
0480 The actual track reconstruction procedure itself starts with the conversion of
0481 raw inputs that have been read out from the detector. In case of silicon
0482 detectors, the readout can either be performed in a binary way, only recording
0483 which segments fired, or the amount of charges measured in the segment can be
0484 recorded, e.g. via *time-over-threshold* readout. In all cases, the readout is
0485 attached to an identifier uniquely locating the segment on the corresponding
0486 sensor.
0487 
0488 As a next step, these raw readouts need to be *clustered*, in order to
0489 extract an estimate of where particles intersect with the sensor. The general
0490 strategy of clustering algorithms follows the Connected Component Analysis (CCA)
0491 approach, where subsets of segments are successively grouped into clusters.
0492 In case of the Pixel detector, this clustering occurs in two dimensions,
0493 corresponding to the segmentation of its sensors. Here, the CCA can
0494 either consider all eight surrounding pixels as neighboring a central one, or
0495 only consider the four non-diagonal ones, as shown in
0496 {numref}`clustering_cca`. The figure only shows the simplest possible
0497 cluster starting from the central pixel. In reality, the CCA will iteratively
0498 continue from the pixels on the cluster edges.
0499 
0500 (clustering_cca)=
0501 :::{figure} /figures/tracking/cca.svg
0502 :align: center
0503 :width: 400px
0504 Illustration of both eight and four cell connectivity.
0505 :::
0506 
0507 Subsequently, the effective cluster position needs to be estimated. Multiple
0508 factors play a role here. First of all, the average position of the cluster
0509 can be calculated either using only the geometry position of the segments,
0510 
0511 $$
0512   \vec r = \frac{1}{N} \sum_{i=1}^N \vec l_i,
0513 $$
0514 
0515 or be weighted by the charge collected in each segment:
0516 
0517 $$
0518   \vec r = \frac{1}{\sum_{i=1}^N q_i} \sum_{i=1}^N q_i \vec l_i.
0519 $$
0520 
0521 Here, $\vec l_i$ is the local position of the $i$-th segment while
0522 $q_i$ is its charge.
0523 
0524 An illustration of the clusterization can be found in {numref}`clustering_image`,
0525 where a pixel sensor is shown to be intersected by a charged particle,
0526 entering on the lower left and exiting on the top right. Three cells shown
0527 with a red frame receive energy from the particle, but the amount is under
0528 the readout threshold. Four other cells receive energy above the threshold
0529 and are read out. The clustering will then group these four cells into a
0530 cluster, and subsequently estimate the cluster position based on the energy
0531 deposited in the cells. In case no charge information is not available
0532 for a given detector, the calculation is purely geometric.
0533 
0534 
0535 (clustering_image)=
0536 :::{figure} /figures/tracking/clustering.svg
0537 :align: center
0538 :width: 400px
0539 Illustration of the clustering of multiple pixels into a cluster,
0540 in a three-dimensional view on the left and a projection onto the
0541 $xy$-plane on the right. A particle enters the center in the lower left,
0542 crosses several segments before exiting the sensor on the top right. The
0543 cell colors indicate how far along the trajectory they are encountered.
0544 :::
0545 
0546 Another factor that needs to be accounted for is the drift direction of the
0547 created charges. In addition to the collection field of the sensor itself,
0548 the surrounding magnetic field modifies the drift direction by the
0549 *Lorentz-angle* $\theta_\text{L}$. Depending on the field strength, this
0550 additional angle can cause segments to be activated that would otherwise not
0551 be geometrically within reach of the charges. Other effects, such as the fact
0552 that the modules are not perfectly flat, as the geometry description assumes,
0553 or cross-talk between readout channels, also play a role at this stage.
0554 
0555 In the presence of high event activity, particle intersections on single
0556 sensors can be close enough to one another to result in clusters that are not
0557 clearly separated from each other. This circumstance can be somewhat
0558 mitigated by allowing tracks to share clusters with other particles, which
0559 comes at the price of allowing duplicated tracks to some extent.
0560 Additionally, merged clusters typically feature worse position resolution,
0561 which manifests itself since it negatively affects the final fit of the
0562 track.
0563 
0564 (tracking_sp_formation)=
0565 ## Spacepoint formation
0566 
0567 The basic input to most forms of pattern recognition algorithms for tracking
0568 are space points, which need to be assembled from the raw measurements. To this
0569 end, the raw measurements are combined with information provided by the
0570 geometry description, such as the location and rotation of the sensors. In this
0571 way, the locations, which are restricted to be local to the sensor surfaces
0572 intrinsically, can be converted into three dimensional points in space.  See
0573 [](spacepoint_formation_core) for a description of the implementation of
0574 spacepoint formation in the core library.
0575 
0576 {numref}`sensor` shows an illustration of the information that is consumed for
0577 a pixel measurement. Shown are three clusters on a sensor, which are caused by
0578 three tracks intersecting it. The corresponding cluster positions are indicated
0579 as well, and can be converted to global positions using the inverse of the
0580 global-to-local transformation matrix, that is provided by the geometry
0581 description.
0582 
0583 (sensor)=
0584 :::{figure} /figures/tracking/sp_l2g.svg
0585 :align: center
0586 :width: 400px
0587 Illustration of a pixel sensor and its local coordinate system in
0588 relation to the global laboratory frame. A transformation allows conversion
0589 between the two systems. Shown are three tracks intersecting the sensor,
0590 alongside clusters that they produce.
0591 :::
0592 
0593 In strip detectors, on the other hand, only a single
0594 dimension is segmented, and an individual measurement is therefore only
0595 constrained in one direction on the surface. However, usually the
0596 strip modules are mounted in pairs, with a stereo angle rotation
0597 between the pairs. To form global space points, measurements from both
0598 sensors of a pair need to be combined.
0599 Due to the stereo angle, a two dimensional
0600 location on the orthogonal projection plane relative to the two parallel
0601 pairs can be found. Using the global transformations of the pair, the
0602 combined measurement location can be converted to global coordinates. If
0603 multiple measurements are located on a stereo pair of strip sensors, there
0604 exists an ambiguity on how to combine strips to form space points, which has to be resolved.
0605 
0606 
0607 ## Seeding
0608 
0609 The next step after space point formation is pattern recognition, which be
0610 implemented in various ways.  Global methods exist which attempt to cluster
0611 space points, such as conformal mapping. In this approach, the space points are
0612 transformed into a feature parameter space that reveals patterns for hits
0613 belonging to the same track. In the specific example of a Hough transform, a
0614 parameter space $\left(\phi, q/p_\mathrm{T}\right)$ is used. As a result, each
0615 space point is effectively transformed into a line, as a series of combinations
0616 of these parameters would lead to the same space point. The lines from a set of
0617 space points of a single track will intersect in one common area. Such an
0618 intersection can be used to identify which space points originate from the same
0619 track. However, this task grows in complexity as detector activity increases
0620 and is susceptible to material effects. See [](seeding_core) for a description
0621 of the seeding implementation in the core library.
0622 
0623 Another group of approaches is the one of seeding and track following. These
0624 algorithms differ from the global ones in that they evaluate individual
0625 combinations of space points, and successively explore the events. One
0626 algorithm from this group is the cellular automaton that iteratively forms
0627 chains of space points going from one layer to the next.
0628 
0629 The main approach in ACTS is an algorithm that operates on coarse
0630 subdivisions of the detector is used. This seeding algorithm attempts to find
0631 triplets of space points from increasing radii which are likely to belong to
0632 the same track. It achieves this by iterating the combinatorial triplets and
0633 successively filtering them. Filtering is performed based on the momentum and
0634 impact parameters, which the algorithm attempts to estimate for each triplet.
0635 
0636 Under the assumption of a homogeneous magnetic field along the $z$-axis,
0637 charged particles should follow helical trajectories. In the transverse plane,
0638 the motion is circular, while it is a straight line in the $rz$-plane.  The
0639 transverse impact parameter and momentum can be estimated from the radius of
0640 the circle in the transverse plane like
0641 
0642 $$
0643   d_0 = \sqrt{c_x^2 + c_y^2} - \rho,
0644 $$
0645 
0646 with the circle center $(c_x, c_y)$ and radius $\rho$. The
0647 transverse momentum can be related to available quantities like
0648 
0649 $$
0650   p_\mathrm{T} \propto \cdot q B \rho
0651 $$
0652 
0653 with the charge $q$ and the magnetic field $B$. An intersection
0654 between the straight line in the $rz$-plane with the $z$-axis gives an
0655 estimate of the longitudinal impact parameter.
0656 An illustration of seeds in the transverse plane is found in
0657 {numref}`seeding_figure`. Note that seeds can incorporate hits spread across all of
0658 the layers shown, although this can be a configuration parameter.
0659 
0660 
0661 (seeding_figure)=
0662 :::{figure} /figures/tracking/seeding.svg
0663 :width: 300px
0664 :align: center
0665 Sketch of seeds in the transverse plane for a number of tracks on
0666 four layers. Seeds can combine hits on any three of these layers. The shown
0667 seeds appear compatible with having originated in the center of the
0668 detector, which is also drawn.
0669 :::
0670 
0671 ## Track finding and track fitting
0672 
0673 In the track seeding and following approach, track candidates are built from
0674 the initial seeds. One method implemented in ACTS, the [Combinatorial Kalman
0675 Filter](#combinatorial-kalman-filter) (CKF), uses the *Kalman formalism*.
0676 Originally developed for monitoring and steering mechanical systems, it can
0677 also be used to iteratively calculate a track estimate. After a set of track
0678 candidates has been assembled and filtered (see [](#ambiguity-resolution)), an
0679 additional track fit is usually performed to extract the best estimate of the
0680 track. The Kalman formalism can also be used for this, with the addition of a
0681 smoothing step that has certain benefits.  Other fit strategies exist, such as
0682 a global $\chi^2$ fit that minimizes the distances between track-sensor
0683 intersections and measurements on all sensors at the same time. One drawback of
0684 this method is the necessity to invert very large matrices, which is
0685 computationally expensive.
0686 
0687 In a track fit, the Kalman formalism can be shown to yield optimal estimates
0688 for Gaussian uncertainties. This assumption breaks down when effects like
0689 bremsstrahlung come into play. An extension of the Kalman Filter (KF) exists
0690 that relies on the individual propagation of a set of trajectories, instead of
0691 a single one, to model these biased uncertainties by a sum of Gaussian
0692 components. This [Gaussian Sum Filter](gsf_core) (GSF) achieves better results when
0693 fitting particles such as electrons, likely to undergo bremsstrahlung, and is
0694 deployed in e.g. the ATLAS tracking chain.
0695 
0696 
0697 ### Kalman formalism and Kalman track fitter
0698 
0699 :::{tip}
0700 See [](kf_core) for a description of the implementation of the Kalman Filter in
0701 the core library.
0702 :::
0703 
0704 The basis of the Kalman formalism is a state vector, that can be identified
0705 with the set of track parameters $\vec x$. Note that the concrete
0706 parametrization plays a subordinate role in this context. Rather than building
0707 an estimate of the state of a system in real time, a Kalman track fit can be
0708 understood as estimating the parameters iteratively in steps. In the track
0709 fitting application, each step is defined by a measurement to be included.
0710 The evolution of the state vector is described by
0711 
0712 $$
0713   \vec x_k = \mathbf F_{k-1} \vec x_{k-1} + \vec w_{k-1},
0714 $$
0715 
0716 where the linear function $\mathbf F_{k-1}$ transports the state vector at
0717 step $k-1$ to step $k$. $\vec w_{k-1}$ is additional so-called process noise
0718 that affects the transport additively. Each step has an associated
0719 measurement, with the fixed relationship between the measurement and the state vector
0720 
0721 $$
0722   \vec m_k = \mathbf H_k \vec x_k + \epsilon_k.
0723 $$
0724 
0725 Here, $\mathbf H_k$ is the *measurement mapping function*, which
0726 transforms the state vector into the measurement space. In the ideal case,
0727 this purpose can be achieved by a simple projection matrix, which extracts a
0728 subspace of the state vector. Additionally, $\epsilon_k$ represents the
0729 measurement uncertainty.
0730 
0731 The Kalman fit process is divided into different phases:
0732 
0733 1. **Prediction** of the state vector at the next step $k+1$ based on the information at the current step $k$.
0734 2. **Filtering** of the prediction by incorporating the measurement associated to the step
0735 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 $k$.
0736 
0737 An illustration of these concepts is found in {numref}`kalman_filter`. Here,
0738 a series of three sensors is shown with measurements on them. The KF
0739 then predicts the track parameters at an intersection, shown in blue.
0740 Subsequently, a filtered set of parameters is calculated as a mixture between
0741 the measurement and the prediction. Not shown in this picture is the
0742 smoothing step.
0743 
0744 (kalman_filter)=
0745 :::{figure} /figures/tracking/kalman.svg
0746 :width: 400px
0747 :align: center
0748 Illustration of the KF. Two of the three stages, the prediction and the
0749 filtering are shown. The filtering updates the prediction with information from
0750 the measurement.
0751 :::
0752 
0753 
0754 In many cases, the first two phases run in tandem, with prediction and
0755 filtering happening alternatingly at each step. The smoothing phase is
0756 launched once the last measurement has been encountered.
0757 Starting from a state $k$, first, a prediction of the state vector at the
0758 next measurement location is obtained via
0759 
0760 (kf_pred)=
0761 $$
0762   \vec x_k^{k-1} = \mathbf F_{k-1} \vec x_{k-1},
0763 $$
0764 
0765 with the linear transport function from above. $\vec x_k^{k-1}$ is
0766 the prediction of the state vector at step $k$ based on step $k-1$. The next
0767 stage is the filtering. Here, the state vector is refined by taking into
0768 account the measurement at the current step. Following one of two variants of
0769 filtering from [^Fruhwirth:1987fm], the gain matrix formalism, the state
0770 vector is updated like
0771 
0772 $$
0773   \vec x_k = \vec x_k^{k-1} + \mathbf K_k \left( \vec m_k - \mathbf H_k \vec x_k^{k-1} \right),
0774 $$
0775 
0776 with the *Kalman gain matrix*
0777 
0778 $$
0779   \mathbf K_k = \mathbf C_k^{k-1} \mathbf H_k^\mathrm{T}
0780     \left(
0781       \mathbf V_k + \mathbf H_k \mathbf C_k^{k-1} \mathbf H_k^\mathrm{T}
0782     \right)^{-1}
0783     .
0784 $$
0785 
0786 Note that $\vec x_k$ is the filtered state vector at step $k$,
0787 based on information from previous steps and step $k$ itself. This is in
0788 contrast to $\vec x_k^{k-1}$, which is the prediction of the state vector at
0789 step $k$ based on $k-1$, and is used to calculate the filtered state vector.
0790 One input to these equations is the covariance matrix prediction $\mathbf
0791 C_k^{k-1}$ at step $k$ based on step $k-1$, which can be written as
0792 
0793 $$
0794   \mathbf C_k^{k-1}  = \mathbf F_{k-1} \mathbf C_{k-1} \mathbf F_{k-1}^\mathrm{T} + \mathbf Q_{k-1}
0795 $$
0796 
0797 in the linear version from [^Fruhwirth:1987fm], with the
0798 covariance $\mathbf C_{k-1}$ at step $k-1$, and the covariance $\mathbf
0799 Q_{k-1}$ associated with $\vec w_{k-1}$ from above. Also needed is $\mathbf
0800 V_k$, which is the covariance associated with $\epsilon_k$, effectively
0801 representing the measurement uncertainty.
0802 
0803 Similar to the state vector itself, the corresponding covariance matrix is
0804 also filtered using
0805 
0806 (kf_cov_pred)=
0807 $$
0808   \mathbf C_k = \left( \mathbb 1 - \mathbf K_k \mathbf H_k \right) \mathbf C_k^{k-1}.
0809 $$
0810 
0811 In the smoothing phase, the state vector at step $k$ is improved using the
0812 information from the subsequent step $k+1$ using
0813 
0814 $$
0815   \vec x_k^n = \vec x_k + \mathbf A_k \left( \vec x_{k+1}^n - \vec x_{k+1}^k \right).
0816 $$
0817 
0818 Here, $\vec x_{k+1}^n$ is the smoothed state vector and $\vec
0819 x_{k+1}^k$ the predicted state vector at the subsequent step $k+1$. Also
0820 needed is the *smoother gain matrix*
0821 
0822 $$
0823   \mathbf A_k = \mathbf C_k \mathbf F_k^\mathrm{T} \left( \mathbf C^k_{k+1} \right)^{-1},
0824 $$
0825 
0826 with the predicted covariance at step $k+1$, $\mathbf C_k^{k+1}$.
0827 Finally, the covariance at the current step $k$ can also be smoothed with
0828 
0829 $$
0830   \mathbf C_k^n = \mathbf C_k + \mathbf A_k \left(\mathbf C_{k+1}^n - \mathbf C_{k+1}^k \right) \mathbf A_k^\mathrm{T}.
0831 $$
0832 
0833 After smoothing, the parameter estimate at the first step contains information
0834 from all other measurement states. As mentioned above, in case the
0835 uncertainties entering the Kalman fit are Gaussian distributions without
0836 biases, the KF can be shown to be the optimal solution minimizing mean
0837 square estimation error. However, certain caveats exist. The KF assumes
0838 that a linear transport function $\mathbf F$ exists that can propagate the
0839 state vector. In the presence of inhomogeneous magnetic fields this is not
0840 the case. Instead of explicitly applying $\mathbf F$ to the state vector for
0841 the prediction, the ACTS KF turns to the numerical integration,
0842 discussed in [](#numerical-integration). With it, the prediction from
0843 [this equation](kf_pred) is simply the intersection of the extrapolated trajectory
0844 with the next sensitive surface. Aside from this, $\mathbf F$ is also used to
0845 transport the covariance between steps (see [here](kf_cov_pred)). Here, the
0846 semi-analytical method for covariance transport in the numerical integration
0847 can be used. $\mathbf F$ can then be identified with the transport
0848 Jacobian accumulated between surfaces.
0849 
0850 For smoothing, two possibilities exist to obtain the needed covariances from
0851 subsequent measurement steps. Either, the inverse transport Jacobian is used
0852 and applied, in a way similar to [this equation](kf_cov_pred), or the numerical
0853 integration is executed again in an inverse fashion, propagating from the
0854 subsequent step to the current one.
0855 
0856 ### Combinatorial Kalman Filter
0857 
0858 :::{tip}
0859 See [](ckf_core) for information on the CKF implementation found in the core
0860 library.
0861 :::
0862 
0863 As mentioned above, the Kalman formalism can be used for track finding. In this
0864 case, the smoothing step can be dropped, as the resulting track candidates are
0865 likely to be refit regardless, therefore saving some time. The CKF explores the
0866 event starting from an initial track seed. It does this by considering not only
0867 a single sequence of measurements, but allowing the branching of the fit at
0868 each sensitive surface that is encountered. To this end, all or a subset of
0869 measurements that are found on each surface are considered. Measurements are
0870 selected based on their compatibility with the current state estimate, by using
0871 their residuals. A predicted residual
0872 
0873 $$
0874   \vec r_k^{k-1} = \vec m_k - \mathbf H_k \vec x_k^{k-1},
0875 $$
0876 
0877 and a filtered residual
0878 
0879 $$
0880   \vec r_k = \vec m_k - \mathbf H_k \vec x_k
0881   ,
0882 $$
0883 
0884 can be defined, depending on which state estimate is compared with
0885 the measurement $\vec m_k$. Using the filtered residual, an effective
0886 $\chi^2$ increment
0887 
0888 $$
0889   \chi^2_+ = \vec r_k^\mathrm{T}
0890   \left[ \left( \mathbb 1 - \mathbf H_k  \mathbf K_k \right)  \mathbf V_k \right]^{-1}
0891   \vec r_k
0892 $$
0893 
0894 can be calculated. The global $\chi^2$ of the track candidate can
0895 be calculated as the sum of all $\chi^2_+$ across the steps. Measurements
0896 with a large $\chi^2_+$ are considered as outliers, which have low
0897 compatibility with the trajectory. By branching out for measurements below a
0898 certain $\chi^2_+$, and following the branches, a tree-like structure of
0899 compatible track candidates originating from a track seed is assembled. This
0900 feature is shown in {numref}`tracking_ckf`, which displays a circular
0901 trajectory, and a set of iteratively assembled track candidates. Basic
0902 quality criteria can be applied at this stage, to remove bad candidates. A
0903 dedicated [](#ambiguity-resolution).
0904 selects the candidates most likely to belong to real particle tracks.
0905 
0906 (tracking_ckf)=
0907 :::{figure} /figures/tracking/finding.svg
0908 :width: 300px
0909 :align: center
0910 Illustration of the way the CKF iteratively explores
0911 measurements from a seed outwards. Measurements are added successively, and
0912 can be shared between the resulting track candidates. Shown in green is a
0913 circular *real* trajectory.
0914 :::
0915 
0916 ## Ambiguity resolution
0917 
0918 Due to the combinatorial nature of track finding, and to achieve high
0919 efficiencies, this set of candidates is often large, and contains a
0920 non-negligible fraction of *fake* candidates. These fake candidates are either
0921 completely combinatorial, or arise from real particle measurements with
0922 combinatorial additions. Track candidates coming from a single seed necessarily
0923 share a common stem of measurements. Measurements can potentially also be
0924 shared between candidates from different seeds.
0925 
0926 One possibility to resolve this (as is done in e.g. ATLAS tracking) is an
0927 ambiguity resolution algorithm, that attempts to filter out as many undesirable
0928 tracks as possible. This is implemented by means of a scoring function, that
0929 combines properties of the track parameters. Higher scores are correlated with
0930 a larger probability to be a desirable track candidate. A larger number of hits
0931 results in an increase in the score, as longer compatible hit chains are less
0932 likely to be random combinations. On the other hand, missing hits in sensors
0933 where a hit was expected negatively impact the score.  Experiment specific
0934 scoring of hits from different subsystems is also implemented. The overall
0935 $\chi^2$ value computed for the track candidate also plays a role. Candidates
0936 that share hits with other candidates are penalized. Another quantity is the
0937 measured particle $p_\mathrm{Y}$, which enters the score, to give preference to
0938 tracks with large momenta. For tracks containing measurements with a
0939 substantial local $\chi^2_+$ at the start or end of the trajectory, the
0940 ambiguity resolution step can also attempt to remove these hits, and determine
0941 whether a refit without them yields a more favorable global $\chi^2$.
0942 
0943 Finally, the output of the ambiguity resolution step is a set of track candidates
0944 that contain an enhanced fraction of tracks from actual particles, while fake
0945 tracks are suppressed. They are passed into the final precision fit outlined
0946 in [](#track-finding-and-track-fitting), to extract the parameter estimate, and used
0947 for further aspects of reconstruction.
0948 
0949 ## Vertex reconstruction
0950 
0951 :::{tip}
0952 See [](vertexing_core) for a dedicated description of the vertexing as
0953 implemented in ACTS.
0954 :::
0955 
0956 A vertex is a point within the detector, where an interaction or a 
0957 decay occurred. We distinguish between primary vertices (from 
0958 collisions/interactions) and secondary vertices (from subsequent particle 
0959 decays), see {numref}`vertexing_illust`. Primary vertices are further divided 
0960 into hard-scatter and pile-up vertices. While primary vertices are located in 
0961 the luminous region, secondary vertices are slightly displaced due to the finite
0962  life time of the decaying particle. 
0963 
0964 (vertexing_illust)=
0965 :::{figure} /figures/tracking/vertexing.svg
0966 :width: 400px
0967 :align: center
0968 Illustration of a set of three vertices in a proton-proton
0969 collision. We distinguish between primary hard-scatter, primary pile-up, and 
0970 secondary vertices.
0971 :::
0972 
0973 Vertices play an important role in higher-level reconstruction algorithms. For 
0974 example, secondary vertices can help with the identification of particles: 
0975 During *$b$-tagging*, a displaced vertex located inside a jet is a sign for the 
0976 decay of a $b$-hadron.
0977 
0978 In analogy to track reconstruction, vertex reconstruction can be divided into 
0979 two stages: vertex finding and vertex fitting. As a first step of vertex 
0980 finding, we compute a rough estimate of the vertex position from a set of 
0981 tracks. This first estimate can be calculated in many different ways, and is
0982 referred to as "vertex seed". Seeding algorithms differ for primary and 
0983 secondary vertexing. For primary vertex seeding, one option is to use a 
0984 histogram approach to cluster tracks on the $z$-axis[^phd:piacquadio:2010]. 
0985 This is based on the assumption that primary vertices will be close to the 
0986 beamline. Other approaches model tracks as multivariate Gaussian distributions 
0987 and identify regions of high track density as vertex seeds[^phd:schlag:2022]. 
0988 For secondary vertexing, seeds are formed from pairs of reconstructed tracks as 
0989 the constraint to the beamline does not apply.
0990 
0991 Once a vertex seed is determined, tracks that are compatible with it are 
0992 selected as part of the vertex finding.
0993 
0994 Before the vertex fit, we linearize tracks in the vicinity of the vertex seed
0995 under assuming that they follow a helical (for constant magnetic field) or
0996 straight (for no magnetic field) trajectory[^phd:piacquadio:2010]. The vertex 
0997 fitter then uses this linearization to improve the position of the vertex seed. 
0998 Furthermore, the track momenta are refitted under the assumption that the tracks 
0999 originate at the vertex[^Fruhwirth:1987fm] [^billoirfitting:1992] . 
1000 
1001 One issue with an approach like this is that the assignment of tracks to 
1002 vertices is ambiguous. As an improvement, one can perform a multi-vertex fit,
1003 where vertices compete for tracks. This means that one track can be assigned to 
1004 several vertices. Their contribution to each vertex fit is determined by a
1005 weight factor, which, in turn, depends on the tracks' compatibility with respect
1006 to all vertices[^fruwirth:amvfitting:2004].
1007 
1008 A flowchart of a multi-vertex reconstruction chain is shown in 
1009 {numref}`vertexing_flowchart`.
1010 
1011 (vertexing_flowchart)=
1012 :::{figure} /figures/tracking/vertexing_flowchart.svg
1013 :width: 600px
1014 :align: center
1015 Simplified flowchart of multi-vertex reconstruction. From a set of seed tracks, 
1016 we first compute a rough estimate of the vertex position, i.e., the vertex seed. 
1017 Then, we evaluate the compatibility of all tracks with the the latter. If a 
1018 track is deemed compatible, it is assigned a weight and attached to the vertex 
1019 seed. Next, the vertex seed and all previously found vertices that share tracks 
1020 with it are (re-)fitted. Finally, after convergence of the fit, we check whether
1021 the vertex candidate is merged with other vertices and discard it if that is the 
1022 case. For the next iteration, all tracks that were assigned to the vertex seed 
1023 and that have a weight above a certain threshold are removed from the seed 
1024 tracks.
1025 :::
1026 
1027 [^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)
1028 [^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)
1029 [^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.
1030 [^phd:schlag:2022]: B. Schlag, 2022, Advanced Algorithms and Software for Primary Vertex Reconstruction and Search for Flavor-Violating Supersymmetry with the ATLAS Experiment.
1031 [^billoirfitting:1992]: P. Billoir et al., 2022, Fast vertex fitting with a local parametrization of tracks.
1032 [^fruwirth:amvfitting:2004]: R. Frühwirth et al., 2004, Adaptive Multi-Vertex fitting.