Systematic approach to measure the performance of microchannel-plate photomultipliers

A. Lehmann albert.lehmann@fau.de M. Böhm M. Götz K. Gumbert S. Krauss D. Miehling M. Pfaffinger Friedrich Alexander-University of Erlangen-Nuremberg, Erlangen, Germany
Abstract

In this paper, we present our approach to systematically measure numerous performance parameters of microchannel plate photomultiplier tubes. The experimental setups, the analyses and selected results are discussed. Although the techniques used may be different in other locations, the document is intended as a guide for comparable measurements with other types of microchannel plate photomultiplier tubes. Measurements are shown for the following performance parameters: spectral and spatial quantum efficiency, collection efficiency, gain as a function of voltage, position and magnetic field, time resolution, rate capability and lifetime. By using a dedicated 3-axis stepper and an FPGA-based data acquisition system, also inner PMT parameters are measured as a function of the active area, such as relative detection efficiency, dark count rate, time resolution, recoil electron and afterpulse distributions, as well as charge sharing and electronic crosstalk. In addition, some of the parameters are investigated inside a strong magnetic field. For many of these measurements, the change of most setup parameters and the subsequent analysis can be controlled semi-automatically by software scripts.

keywords:
PANDA , multianode microchannel-plate PMTs , performance parameters , quantum efficiency , collection efficiency , gain , time resolution , rate capability , lifetime , dark count rate , recoil electrons , afterpulses , crosstalk , magnetic field
journal: Journal of  Templates
\geometry

margin=1.4cm

1 Introduction

Recently, microchannel-plate photomultiplier tubes (MCP-PMTs) have become very attractive sensor candidates for applications where ultrafast single photon detection is required in high magnetic fields, at very high photon rates and in harsh radiation environments [1, 2, 3, 4]. This is due to the fact that their lifetime has increased considerably in the last greater-than-or-equivalent-to\gtrsim10 years [5]. Typical sizes today range from 1×\times×1 inch2 and 2×\times×2 inch2 active area offered by commercial suppliers such as Hamamatsu Photonics K.K. (HPK) [6], Photek Ltd [7] and PHOTONIS [8], to 8×\times×8 inch2 Large Area Picosecond Photodetectors (LAPPD) [9] from Incom Inc [10]. These PMTs can be manufactured in various configurations depending on the intended application. A schematic representation of a typical MCP-PMT with a connected voltage divider is shown in Fig. 1. For detailed qualification tests of such tubes, reliable setups and measurement methods are required to measure and compare the performance parameters.

Refer to caption
Figure 1: Diagram of an MCP-PMT with its high voltage (HV) divider [17]. Many of the terms and designations used in this document are defined here. The capacitances inside the HV divider are usually 4.7 nF.

MCP-PMTs are the photon sensors of choice for the DIRC detectors [11, 12, 13, 14] of the PANDA experiment [15, 16] at the new FAIR accelerator complex at GSI in Darmstadt, Germany. In this context, various specialized measurement setups and analysis tools have been developed over the years for a thorough and comprehensive investigation of the performance characteristics of MCP-PMTs. Recently, the first series production MCP-PMTs for the PANDA Barrel DIRC [11, 12] were tested for their quality by measuring and comparing as many key performance parameters as possible. Here, we present a systematic approach to the detailed measurement of most MCP-PMT performance parameters within a manageable period of a few days. In particular, we focus on the description of the experimental setups and the procedures used to record the different parameters. Since the approaches are quite different, we describe them separately in each section. In addition, the analyses used and some representative results for each parameter are discussed.

In Section 2 the spectral quantum efficiency (QE) between 200 and 850 nm and the spatial QE homogeneity are determined in specific scans. A setup for measuring the collection efficiency (CE) is also presented, which is applicable for almost all PMTs. The gain of the MCP-PMTs (Section 3) is determined as a function of voltage and its homogeneity over the active area, for some of the tubes also within a magnetic field. Both the precise transit time spread (TTS) and a root-mean-square (RMS) time resolution are determined in Section 4, using a fast oscilloscope or a waveform digitizer. The advantages and disadvantages of two methods – ”current mode” and ”pulse mode” – for measuring rate capability are explained in Section 5, followed by the presentation of an approach for measuring the MCP recharging time. The recently observed ”escalation effect” (Section 6) is briefly debated before our illumination setup and the measurement and analysis procedure for the aging studies are discussed in detail in Section 7.

In the last two chapters, we present an efficient method for measuring some internal parameters of an MCP-PMT as a function of the active area. Position scans are performed with a self-built 3-axis stepper and an FPGA-based TRB3 data acquisition (DAQ) system developed at GSI. The setup, the scanning procedure and the analyses are discussed in detail in Section 8, so that it should be possible to assemble a comparable system at other locations. The parameters accessible with this setup are the relative detection efficiency, the dark count rate (DCR), the time resolution (including electronics) and the distribution of recoil electrons and afterpulses. The investigation and quantification of different types of crosstalk is also possible with this system, e.g. charge sharing, electronic and ”coherent” oscillations. Finally, the approach to perform such scans within the magnetic field is presented in Section 9. The latter provides further insight into the behavior of primary and secondary electrons on the PMT performance when electromagnetic focusing is involved.

Although numerous other MCP-PMTs have also been investigated, in this article we focus on the performance measurements of the PHOTONIS Planacon 2-inch MCP-PMTs XP85112-S-BA and XP85132. These have a square shape with 2 MCP stages, 8×\times×8 or 100×\times×3 anode pixels, respectively, and 10 µm pores coated with an ALD (Atomic Layer Deposition) technique [18] to extend the lifetime of the tube.

2 Photon Detection Efficiency

A very important parameter of any photosensor is its photon detection efficiency (PDE), which is a measure of the probability that an incident photon at the photocathode will be converted into a detectable signal at the anode. The PDE is the product of the QE, the CE and the fraction of active area or form factor (FF) of the sensor:

PDE=QECEFF𝑃𝐷𝐸𝑄𝐸𝐶𝐸𝐹𝐹PDE=QE\cdot CE\cdot FFitalic_P italic_D italic_E = italic_Q italic_E ⋅ italic_C italic_E ⋅ italic_F italic_F (1)

2.1 Fraction of Active Area or Form Factor

The form factor of an MCP-PMT is determined by the ratio between the active photocathode area and the total surface area of the PMT exposed to the incident photon flux. Obviously, the FF depends on the type and geometry of the PMT. It ranges from much-less-than\ll50% for small MCP-PMTs to 81% for the 2-inch Planacon tubes used in the PANDA DIRCs. The value of the FF is determined exclusively by the layout and overall dimensions of the respective tube.

2.2 Quantum Efficiency

The QE describes the probability that a photon hitting the photocathode (PC) of the PMT is converted into a photoelectron (PE) that escapes into the vacuum. A typical peak QE of MCP-PMTs ranges from similar-to\sim20% to >>>30%, depending on the PC material. To qualify the detection efficiency of the PMT, we measure the QE as a function of the photon wavelength, but also its homogeneity across the photocathode at a fixed wavelength (in our case 372 nm or 633 nm, depending on the pulsed picosecond laser (PiLas from Advanced Laser Diode Systems [19]) used).

2.2.1 Spectral QE

The spectral QE is measured in a special in-house setup (see Fig. 2), which consists of a Xenon lamp as light source and a monochromator LOT Oriel MSH301 (later renamed Oriel Cornerstone 260 [20]) with different gratings and <<<1 nm wavelength resolution in a range from 200 to 850 nm. After passing through slits, a grating, several mirrors and, at the exit of the monochromator, another slit to eliminate second-order diffraction effects and a filter wheel to adjust the intensity, the wavelength-selected light is directed into a dark box containing the MCP-PMT and a reference diode [Hamamatsu S6337-01, QEref(λ)𝑄subscript𝐸𝑟𝑒𝑓𝜆QE_{ref}(\lambda)italic_Q italic_E start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT ( italic_λ )]. In the dark box the beam is pointed at either the reference sensors or the MCP-PMT by using a mirror. A PC current of the PMT of typically a few nA and/or the diode current are measured with a Keithley [21] 487 picoammeter. The entire setup is controlled by a Python software script on a Windows computer, which communicates with the monochromator, the filter wheel and the picoammeter via a GPIB interface.

Refer to caption
Figure 2: Schematic of the automated setup for measuring the QE as a function of the wavelength of the photons [22].

In practice, the Xenon lamp is warmed up for 2 to 3 hours before the actual measurement program is started to ensure a stable light intensity. The MCP-PMT is placed in the dark box well before the start of data collection in order to reduce the dark count rate. A voltage of 200 V is applied between the PC and MCP-In, and the photocurrent IPC(λ)subscript𝐼𝑃𝐶𝜆I_{PC}(\lambda)italic_I start_POSTSUBSCRIPT italic_P italic_C end_POSTSUBSCRIPT ( italic_λ ) directly at the PC is measured with the picoammeter. The other components (MCPs, anode) of the MCP-PMT are usually shorted. Before and after each wavelength scan, the dark current Idark,PMTsubscript𝐼𝑑𝑎𝑟𝑘𝑃𝑀𝑇I_{dark,PMT}italic_I start_POSTSUBSCRIPT italic_d italic_a italic_r italic_k , italic_P italic_M italic_T end_POSTSUBSCRIPT is measured for a short time (similar-to\sim10 s) and subtracted in the QE analysis. The diode current Iref(λ)subscript𝐼𝑟𝑒𝑓𝜆I_{ref}(\lambda)italic_I start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT ( italic_λ ) for each wavelength is measured before and sometimes after scanning one or more MCP-PMTs and then used to calculate the QE of the PMT at each wavelength (Eq. 2). The scan over the entire wavelength range runs automatically.

QEPC(λ)=QEref(λ)IPC(λ)Idark,PMTIref(λ)𝑄subscript𝐸𝑃𝐶𝜆𝑄subscript𝐸𝑟𝑒𝑓𝜆subscript𝐼𝑃𝐶𝜆subscript𝐼𝑑𝑎𝑟𝑘𝑃𝑀𝑇subscript𝐼𝑟𝑒𝑓𝜆QE_{PC}(\lambda)=QE_{ref}(\lambda)\cdot\frac{I_{PC}(\lambda)-I_{dark,PMT}}{I_{% ref}(\lambda)}italic_Q italic_E start_POSTSUBSCRIPT italic_P italic_C end_POSTSUBSCRIPT ( italic_λ ) = italic_Q italic_E start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT ( italic_λ ) ⋅ divide start_ARG italic_I start_POSTSUBSCRIPT italic_P italic_C end_POSTSUBSCRIPT ( italic_λ ) - italic_I start_POSTSUBSCRIPT italic_d italic_a italic_r italic_k , italic_P italic_M italic_T end_POSTSUBSCRIPT end_ARG start_ARG italic_I start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT ( italic_λ ) end_ARG (2)

Normally, the QE is measured at two (or more) PC positions of the MCP-PMT. The illuminated light spots with a diameter of similar-to\sim5 mm are indicated by the green and magenta circles on the anode pixel map in Fig. 3 (left). As an example, the measured QE wavelength distributions are shown in Fig. 3 (right), which illustrate the slight position dependence of the QE over the PC.

Refer to caption
Refer to caption
Figure 3: Positions where the QE was measured (left) and spectral QE obtained at these two spots for the PHOTONIS XP85112 [SN 9002193] (right).

2.2.2 QE Homogeneity

In our quality control tests of the MCP-PMTs, we also measure the spatial dependence of the QE over the entire photocathode area. This is done inside a light-tight box (referred to as ”dark box” in the following) with an automated 3-axis (xyz) stepper (see Section 8.1) using a PiLas at a wavelength λscansubscript𝜆𝑠𝑐𝑎𝑛\lambda_{scan}italic_λ start_POSTSUBSCRIPT italic_s italic_c italic_a italic_n end_POSTSUBSCRIPT of usually 372 nm (UV/blue) and sometimes also at 633 nm (red) with a timing resolution (σ𝜎\sigmaitalic_σ) of 14 ps or 12 ps (σ𝜎\sigmaitalic_σ), respectively. The light is guided from the laser head to the PC surface using single-mode optical fibers that end in a microfocus lens. The light intensity can be controlled and adjusted using neutral density (ND) filters and the laser rate. The final focus of the laser beam on the PC can be optimized by adjusting the distance between the lens and the MCP-PMT with the z-coordinate of the stepper. However, the light spot should not be too narrow in order to avoid saturation of the PC current. An additional aperture of about 1 mm diameter is located downstream of the lens to avoid stray light. The actual QE position scan is performed in steps of typically 0.5 mm in both the x and y directions. Our stepper can cover a range of similar-to\sim40 cm in both directions. This allows to measure four (and even more) different MCP-PMTs without having to change the setup, and we can perform long-term scans over nights and weekends. Here too, a voltage of 200 V is applied between the PC and MCP-In, and the other PMT electrodes are shorted. To obtain a measurable PC current of 5-10 nA, we usually do not use ND filters in the QE scans and apply a laser rate of similar-to\sim100 kHz. The local PC current IPC(xy)subscript𝐼𝑃𝐶𝑥𝑦I_{PC}(xy)italic_I start_POSTSUBSCRIPT italic_P italic_C end_POSTSUBSCRIPT ( italic_x italic_y ) at each scanning position xy𝑥𝑦xyitalic_x italic_y is measured with a Keithley 6485 or 6487 picoammeter. A block diagram of this setup is shown in Fig. 4.

Refer to caption
Figure 4: Block diagram of the setup for measuring the position-dependent QE.

Before the actual measurement, the MCP-PMT is placed in the dark for some time to reduce the DCR. After typically four scanned x-rows, the laser spot is moved to the position of the reference diode S6337-01 with calibrated QEref(λscan)𝑄subscript𝐸𝑟𝑒𝑓subscript𝜆𝑠𝑐𝑎𝑛QE_{ref}(\lambda_{scan})italic_Q italic_E start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_s italic_c italic_a italic_n end_POSTSUBSCRIPT ) and a reference current Irefsubscript𝐼𝑟𝑒𝑓I_{ref}italic_I start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT and the dark current Idark,PMTsubscript𝐼𝑑𝑎𝑟𝑘𝑃𝑀𝑇I_{dark,PMT}italic_I start_POSTSUBSCRIPT italic_d italic_a italic_r italic_k , italic_P italic_M italic_T end_POSTSUBSCRIPT of the MCP-PMT are measured. Idark,PMTsubscript𝐼𝑑𝑎𝑟𝑘𝑃𝑀𝑇I_{dark,PMT}italic_I start_POSTSUBSCRIPT italic_d italic_a italic_r italic_k , italic_P italic_M italic_T end_POSTSUBSCRIPT is interpolated for all points along the position scan. The local quantum efficiency QE(xy)𝑄𝐸𝑥𝑦QE(xy)italic_Q italic_E ( italic_x italic_y ) at the PC is calculated as follows:

QE(xy)=QEref(λscan)IPC(xy)Idark,PMTIref𝑄𝐸𝑥𝑦𝑄subscript𝐸𝑟𝑒𝑓subscript𝜆𝑠𝑐𝑎𝑛subscript𝐼𝑃𝐶𝑥𝑦subscript𝐼𝑑𝑎𝑟𝑘𝑃𝑀𝑇subscript𝐼𝑟𝑒𝑓QE(xy)=QE_{ref}(\lambda_{scan})\cdot\frac{I_{PC}(xy)-I_{dark,PMT}}{I_{ref}}italic_Q italic_E ( italic_x italic_y ) = italic_Q italic_E start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_s italic_c italic_a italic_n end_POSTSUBSCRIPT ) ⋅ divide start_ARG italic_I start_POSTSUBSCRIPT italic_P italic_C end_POSTSUBSCRIPT ( italic_x italic_y ) - italic_I start_POSTSUBSCRIPT italic_d italic_a italic_r italic_k , italic_P italic_M italic_T end_POSTSUBSCRIPT end_ARG start_ARG italic_I start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT end_ARG (3)

Fig. 5 (left) shows the QE of the PHOTONIS SN 9002228 as a function of the PC area. Typically, the QE is fairly homogeneous, with slightly lower values towards the rims of the tube. The black grid in the graph shows the position of the PMT anode pixels. Using the measured local QE values at different (>>>100) illuminated xy𝑥𝑦xyitalic_x italic_y points, an average QE value can be determined for each pixel. The ratio between the pixel with maximum QE and the QEs of all other pixels (the QE max/min ratio) is calculated from the mean QEs. The resulting diagram is shown in Fig. 5 (right). This approach can be used to quantify how much of the PC area has a low QE, but for typically much-greater-than\gg90% of the area the QE is very homogeneous with a max/min ratio of <<<1.2.

Refer to caption
Figure 5: QE as a function of PC position (left) for the PHOTONIS SN 9002228, with the anode grid superimposed in black, and typical QE max/min ratios for a set of five PHOTONIS MCP-PMTs (right).

2.3 Collection Efficiency

In order to check the CE values advertised by the various manufacturers, we have developed our own measurement method. Although the calculation of the CE appears simple at first glance, tiny PC currents in the picoampere range make the actual measurement considerably more difficult. The CE is given by

CE=Npe@anodeNpe@PC=Npe@anode@LReLRIPC@HRIdiode@HRIdiode@LR𝐶𝐸subscript𝑁𝑝𝑒@𝑎𝑛𝑜𝑑𝑒subscript𝑁𝑝𝑒@𝑃𝐶subscript𝑁𝑝𝑒@𝑎𝑛𝑜𝑑𝑒@𝐿𝑅𝑒𝐿𝑅subscript𝐼𝑃𝐶@𝐻𝑅subscript𝐼𝑑𝑖𝑜𝑑𝑒@𝐻𝑅subscript𝐼𝑑𝑖𝑜𝑑𝑒@𝐿𝑅CE=\frac{N_{pe@anode}}{N_{pe@PC}}=\frac{N_{pe@anode@LR}\cdot e\cdot LR}{I_{PC@% HR}}\cdot\frac{I_{diode@HR}}{I_{diode@LR}}italic_C italic_E = divide start_ARG italic_N start_POSTSUBSCRIPT italic_p italic_e @ italic_a italic_n italic_o italic_d italic_e end_POSTSUBSCRIPT end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_p italic_e @ italic_P italic_C end_POSTSUBSCRIPT end_ARG = divide start_ARG italic_N start_POSTSUBSCRIPT italic_p italic_e @ italic_a italic_n italic_o italic_d italic_e @ italic_L italic_R end_POSTSUBSCRIPT ⋅ italic_e ⋅ italic_L italic_R end_ARG start_ARG italic_I start_POSTSUBSCRIPT italic_P italic_C @ italic_H italic_R end_POSTSUBSCRIPT end_ARG ⋅ divide start_ARG italic_I start_POSTSUBSCRIPT italic_d italic_i italic_o italic_d italic_e @ italic_H italic_R end_POSTSUBSCRIPT end_ARG start_ARG italic_I start_POSTSUBSCRIPT italic_d italic_i italic_o italic_d italic_e @ italic_L italic_R end_POSTSUBSCRIPT end_ARG (4)

where Npe@anodesubscript𝑁𝑝𝑒@𝑎𝑛𝑜𝑑𝑒N_{pe@anode}italic_N start_POSTSUBSCRIPT italic_p italic_e @ italic_a italic_n italic_o italic_d italic_e end_POSTSUBSCRIPT is the number of photoelectrons (Npe) derived directly from the measured pulse height distribution at the anode, and Npe@PCsubscript𝑁𝑝𝑒@𝑃𝐶N_{pe@PC}italic_N start_POSTSUBSCRIPT italic_p italic_e @ italic_P italic_C end_POSTSUBSCRIPT is deduced from the PC current. Npe@anode@LRsubscript𝑁𝑝𝑒@𝑎𝑛𝑜𝑑𝑒@𝐿𝑅N_{pe@anode@LR}italic_N start_POSTSUBSCRIPT italic_p italic_e @ italic_a italic_n italic_o italic_d italic_e @ italic_L italic_R end_POSTSUBSCRIPT is measured at a low photon rate (LR, 𝒪𝒪\mathcal{O}caligraphic_O(kHz)), while IPC@HRsubscript𝐼𝑃𝐶@𝐻𝑅I_{PC@HR}italic_I start_POSTSUBSCRIPT italic_P italic_C @ italic_H italic_R end_POSTSUBSCRIPT for approximately the same Npe is the PC current obtained at a high photon rate (HR, >>>10 MHz). The currents Idiode@HRsubscript𝐼𝑑𝑖𝑜𝑑𝑒@𝐻𝑅I_{diode@HR}italic_I start_POSTSUBSCRIPT italic_d italic_i italic_o italic_d italic_e @ italic_H italic_R end_POSTSUBSCRIPT and Idiode@LRsubscript𝐼𝑑𝑖𝑜𝑑𝑒@𝐿𝑅I_{diode@LR}italic_I start_POSTSUBSCRIPT italic_d italic_i italic_o italic_d italic_e @ italic_L italic_R end_POSTSUBSCRIPT are determined by a reference diode to scale the low and high rate measurements. e𝑒eitalic_e is the elementary charge.

Refer to caption
Figure 6: Block diagram of the setup for measuring collection efficiency.

The CE setup is sketched in Fig. 6. For the measurement of IPCsubscript𝐼𝑃𝐶I_{PC}italic_I start_POSTSUBSCRIPT italic_P italic_C end_POSTSUBSCRIPT the voltage UPCMCPsubscript𝑈𝑃𝐶𝑀𝐶𝑃U_{PC-MCP}italic_U start_POSTSUBSCRIPT italic_P italic_C - italic_M italic_C italic_P end_POSTSUBSCRIPT between PC and MCP-In is the same as in normal operation and ranges from similar-to\sim150 V in an RPCsubscript𝑅𝑃𝐶R_{PC}italic_R start_POSTSUBSCRIPT italic_P italic_C end_POSTSUBSCRIPT:RMCPsubscript𝑅𝑀𝐶𝑃R_{MCP}italic_R start_POSTSUBSCRIPT italic_M italic_C italic_P end_POSTSUBSCRIPT:RAsubscript𝑅𝐴R_{A}italic_R start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = 1:10:1 voltage divider configuration to similar-to\sim750 V in a 4:10:1 configuration. In this case, the electrode MCP-Out and the anode are usually shorted. It is important to note that the CE depends slightly on the configuration used. For the latest PHOTONIS MCP-PMTs with the 4:10:1 voltage divider configuration, a large CE close to 100% was measured. The large CE is probably achieved by the deposition of a secondary emissive layer on top of the input electrode MCP-In [23]. However, the latter increases the tail of the time distribution (see Fig. 17) and requires the use of a higher UPCMCPsubscript𝑈𝑃𝐶𝑀𝐶𝑃U_{PC-MCP}italic_U start_POSTSUBSCRIPT italic_P italic_C - italic_M italic_C italic_P end_POSTSUBSCRIPT.

Before the actual measurements, the MCP-PMT is placed in the dark for several hours to reduce the dark currents. The light from a PiLas laser is passed through a beam splitter, with one part falling on the PC of the MCP-PMT and the other part on the reference diode. The current of the reference diode is measured with a Keithley 6485 picoammeter at both low (Idiode@LRsubscript𝐼𝑑𝑖𝑜𝑑𝑒@𝐿𝑅I_{diode@LR}italic_I start_POSTSUBSCRIPT italic_d italic_i italic_o italic_d italic_e @ italic_L italic_R end_POSTSUBSCRIPT) and high (Idiode@HRsubscript𝐼𝑑𝑖𝑜𝑑𝑒@𝐻𝑅I_{diode@HR}italic_I start_POSTSUBSCRIPT italic_d italic_i italic_o italic_d italic_e @ italic_H italic_R end_POSTSUBSCRIPT) illumination rates. This is necessary because the light intensity of the laser is not proportional to its pulse frequencies LR𝐿𝑅LRitalic_L italic_R and HR𝐻𝑅HRitalic_H italic_R. The number of photoelectrons (Npe@PCsubscript𝑁𝑝𝑒@𝑃𝐶N_{pe@PC}italic_N start_POSTSUBSCRIPT italic_p italic_e @ italic_P italic_C end_POSTSUBSCRIPT) emerging from the PC is determined by measuring its current (IPC@HRsubscript𝐼𝑃𝐶@𝐻𝑅I_{PC@HR}italic_I start_POSTSUBSCRIPT italic_P italic_C @ italic_H italic_R end_POSTSUBSCRIPT) at high laser rates using a second Keithley picoammeter (6487). As an example, the time evolution of the raw PC current IPCsubscript𝐼𝑃𝐶I_{PC}italic_I start_POSTSUBSCRIPT italic_P italic_C end_POSTSUBSCRIPT and the diode current Idiodesubscript𝐼𝑑𝑖𝑜𝑑𝑒I_{diode}italic_I start_POSTSUBSCRIPT italic_d italic_i italic_o italic_d italic_e end_POSTSUBSCRIPT is shown in Fig. 7 for several applied laser rates (peaks) from 10 MHz to 50 MHz. The tiny photon-induced PC current of less-than-or-similar-to\lesssim12 pA with laser switched on indicates the difficulty of the measurement.

To measure Npe@anodesubscript𝑁𝑝𝑒@𝑎𝑛𝑜𝑑𝑒N_{pe@anode}italic_N start_POSTSUBSCRIPT italic_p italic_e @ italic_a italic_n italic_o italic_d italic_e end_POSTSUBSCRIPT, the full voltage settings are applied to the MCP-PMT. The number of photoelectrons (Npe@anodesubscript𝑁𝑝𝑒@𝑎𝑛𝑜𝑑𝑒N_{pe@anode}italic_N start_POSTSUBSCRIPT italic_p italic_e @ italic_a italic_n italic_o italic_d italic_e end_POSTSUBSCRIPT) that are amplified and reach the anode is derived from the pulse height distribution (see section 3) measured with a 3 GHz LeCroy WavePro 7300A oscilloscope with 20 GSa/s for a 2-channel (or 10 GSa/s for a 4-channel) operation at low light intensity (controlled by ND filters) and at low rates (LR𝐿𝑅LRitalic_L italic_R). All these measurements are performed for different rate and thus intensity configurations, and the final CE is determined as the average of the obtained results.

Refer to caption
Figure 7: Comparison of the currents obtained during a CE measurement of a PHOTONIS (SN 9002230) MCP-PMT at the reference diode (left) and at the photocathode (right) with UPCMCPsubscript𝑈𝑃𝐶𝑀𝐶𝑃U_{PC-MCP}italic_U start_POSTSUBSCRIPT italic_P italic_C - italic_M italic_C italic_P end_POSTSUBSCRIPT = 190 V. The peaks are generated by laser on and laser off settings and correspond to a laser rate from 10 to 50 MHz in 5 MHz steps. After the tube was in the dark for similar-to\sim1 hour, the average dark current during this measurement was  0.564 nA.

3 Gain

The gain is a measure of the number of secondary electrons generated in the amplification process within the MCPs for each photoelectron that reaches the MCP-In. This is a very important performance parameter for an MCP-PMT and can range from <<<105 to >>>107. The secondary electrons are generated in a continuous amplification process in the MCP lead or borosilicate glass pores with a diameter of 3 to 25 µm. The amplification is a function of the length-to-diameter aspect ratio of the MCP pores, the voltage (HV) applied across the MCPs and a potential magnetic field (magnitude and orientation). In this work, we show results of MCP-PMTs with 10 µm pores coated by an ALD technique to extend the lifetime of the PMT.

3.1 Gain vs High Voltage

One of the first qualification steps is to measure the gain of each MCP-PMT as function of HV by illuminating the PC with laser light at a very low photon rate and intensity. The latter is controlled by neutral density filters. A primary goal of this step is to determine the voltage setting at which a gain of 106 is reached. This gain ensures efficient single photon detection and is the goal for operation at the PANDA DIRCs.

Refer to caption
Figure 8: Block diagram of the setup for gain measurements at different HVs.

The anode signal of a central pixel (e.g. 44 or 45 in Fig. 3 (left)) is fed into the LeCroy WavePro 7300A oscilloscope [24] or a CAEN DT5742B desktop waveform digitizer [25]. The signal charge is integrated (usually directly in the oscilloscope) and the accumulated pulse height distributions are stored for further analysis. A schematic representation of our typical setup is shown in Fig. 8. The gain and the number of photoelectrons can be derived by fitting the pulse height distribution (Fig. 9 (left)). If required, this plot can also be used to determine the peak-to-valley (P/V) ratio between the noise band (pedestal) and the single photon signal (first Gaussian of the fit). The measured pulse height distribution is deconvoluted using an approach described in [26]:

S(x)=n=0μneμn!((1w)Gn(xQ0)+wα2exp[α(xQnασn2)](erf[|Q0Qnσn2α|σn2]+sign(xQnσn2α)erf[|xQnσn2α|σn2]))𝑆𝑥superscriptsubscript𝑛0superscript𝜇𝑛superscript𝑒𝜇𝑛1𝑤subscript𝐺𝑛𝑥subscript𝑄0𝑤𝛼2𝛼𝑥subscript𝑄𝑛𝛼superscriptsubscript𝜎𝑛2erfdelimited-[]subscript𝑄0subscript𝑄𝑛superscriptsubscript𝜎𝑛2𝛼subscript𝜎𝑛2sign𝑥subscript𝑄𝑛superscriptsubscript𝜎𝑛2𝛼erfdelimited-[]𝑥subscript𝑄𝑛superscriptsubscript𝜎𝑛2𝛼subscript𝜎𝑛2S(x)=\sum_{n=0}^{\infty}\frac{\mu^{n}\cdot e^{-\mu}}{n!}\cdot\Biggl{(}(1-w)G_{% n}(x-Q_{0})+w\cdot\frac{\alpha}{2}\exp\left[-\alpha(x-Q_{n}-\alpha\sigma_{n}^{% 2})\right]\\ \cdot\biggl{(}\text{erf}\left[\frac{|Q_{0}-Q_{n}-\sigma_{n}^{2}\alpha|}{\sigma% _{n}\sqrt{2}}\right]+\text{sign}(x-Q_{n}-\sigma_{n}^{2}\alpha)\cdot\text{erf}% \left[\frac{\left|x-Q_{n}-\sigma_{n}^{2}\alpha\right|}{\sigma_{n}\sqrt{2}}% \right]\biggr{)}\Biggr{)}start_ROW start_CELL italic_S ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_μ start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ⋅ italic_e start_POSTSUPERSCRIPT - italic_μ end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG ⋅ ( ( 1 - italic_w ) italic_G start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x - italic_Q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + italic_w ⋅ divide start_ARG italic_α end_ARG start_ARG 2 end_ARG roman_exp [ - italic_α ( italic_x - italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_α italic_σ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ] end_CELL end_ROW start_ROW start_CELL ⋅ ( erf [ divide start_ARG | italic_Q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_σ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_α | end_ARG start_ARG italic_σ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT square-root start_ARG 2 end_ARG end_ARG ] + sign ( italic_x - italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_σ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_α ) ⋅ erf [ divide start_ARG | italic_x - italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_σ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_α | end_ARG start_ARG italic_σ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT square-root start_ARG 2 end_ARG end_ARG ] ) ) end_CELL end_ROW (5)

with

Gn(xQ0)subscript𝐺𝑛𝑥subscript𝑄0\displaystyle G_{n}(x-Q_{0})italic_G start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x - italic_Q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) =1σn2πexp((xQn)22σn2)absent1subscript𝜎𝑛2𝜋superscript𝑥subscript𝑄𝑛22superscriptsubscript𝜎𝑛2\displaystyle=\frac{1}{\sigma_{n}\sqrt{2\pi}}\cdot\exp\left(-\frac{(x-Q_{n})^{% 2}}{2\sigma_{n}^{2}}\right)= divide start_ARG 1 end_ARG start_ARG italic_σ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT square-root start_ARG 2 italic_π end_ARG end_ARG ⋅ roman_exp ( - divide start_ARG ( italic_x - italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) (6)
Qnsubscript𝑄𝑛\displaystyle Q_{n}italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT =Q0+nQ1absentsubscript𝑄0𝑛subscript𝑄1\displaystyle=Q_{0}+nQ_{1}= italic_Q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_n italic_Q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (7)
σnsubscript𝜎𝑛\displaystyle\sigma_{n}italic_σ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT =σ02+nσ12absentsuperscriptsubscript𝜎02𝑛superscriptsubscript𝜎12\displaystyle=\sqrt{\sigma_{0}^{2}+n\sigma_{1}^{2}}= square-root start_ARG italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_n italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (8)

The response function S(x)𝑆𝑥S(x)italic_S ( italic_x ) has seven free parameters (5): Q0subscript𝑄0Q_{0}italic_Q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and σ0subscript𝜎0\sigma_{0}italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT define the pedestal, Q1subscript𝑄1Q_{1}italic_Q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and σ1subscript𝜎1\sigma_{1}italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT characterize the pulse height distribution of the first photoelectron. μ𝜇\muitalic_μ is the mean number of photoelectrons and n𝑛nitalic_n the photoelectron index, which is usually limited to n24𝑛24n\leq 24italic_n ≤ 24 in our fits. w𝑤witalic_w and α𝛼\alphaitalic_α describe a potential exponential background below the pedestal that extends into the photon peak. The most important parameters for the analyses are: (a) the mean number of photoelectrons μ𝜇\muitalic_μ that emerge from the PC and are amplified in the MCPs, and (b) the mean signal charge for a photoelectron Q1subscript𝑄1Q_{1}italic_Q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. The PMT gain G𝐺Gitalic_G can be calculated directly from Q1subscript𝑄1Q_{1}italic_Q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT.

Refer to caption
Figure 9: Pulse height distribution and response function fit (left) and gain curve (right) for different voltage divider ratios of the PHOTONIS MCP-PMT SN 9002224. In the fit, w𝑤witalic_w and α𝛼\alphaitalic_α of equation (5) were set to zero.

In our analyses, we generally use the simplified fit function SMCP(x)subscript𝑆𝑀𝐶𝑃𝑥S_{MCP}(x)italic_S start_POSTSUBSCRIPT italic_M italic_C italic_P end_POSTSUBSCRIPT ( italic_x ) (9). The background is neglected (w=0𝑤0w=0italic_w = 0 and α=0𝛼0\alpha=0italic_α = 0), as the accumulation of the pulse height distribution is only triggered if a photon pulse was emitted by the laser. This simplifies the response function to

SMCP(x)=n=024{μneμn!1σn2πexp[(xQn)22σn2]}subscript𝑆𝑀𝐶𝑃𝑥superscriptsubscript𝑛024superscript𝜇𝑛superscript𝑒𝜇𝑛1subscript𝜎𝑛2𝜋superscript𝑥subscript𝑄𝑛22superscriptsubscript𝜎𝑛2S_{MCP}(x)=\sum_{n=0}^{24}\left\{\frac{\mu^{n}\cdot e^{-\mu}}{n!}\cdot\frac{1}% {\sigma_{n}\sqrt{2\pi}}\cdot\exp\left[-\frac{(x-Q_{n})^{2}}{2\sigma_{n}^{2}}% \right]\right\}italic_S start_POSTSUBSCRIPT italic_M italic_C italic_P end_POSTSUBSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 24 end_POSTSUPERSCRIPT { divide start_ARG italic_μ start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ⋅ italic_e start_POSTSUPERSCRIPT - italic_μ end_POSTSUPERSCRIPT end_ARG start_ARG italic_n ! end_ARG ⋅ divide start_ARG 1 end_ARG start_ARG italic_σ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT square-root start_ARG 2 italic_π end_ARG end_ARG ⋅ roman_exp [ - divide start_ARG ( italic_x - italic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] } (9)

Typical gain curves are shown in Fig. 9 (right) for different voltage divider configurations. The voltage between the PC and the MCP is varied in the same way as the voltage between the two MCPs. The curves show that the voltage between the PC and MCP also has a significant influence on the gain.

3.2 Gain Homogeneity

Knowing the gain homogeneity across the active sensor area is important, as it affects the overall photon detection efficiency when the same signal threshold is applied to all pixels during PMT operation. The gain homogeneity measurement setup is based on the same principle as the QE position scan. A schematic representation of the measurement setup is shown in Fig. 10.

Refer to caption
Figure 10: Block diagram for the setup of position-dependent gain measurements.

The MCP-PMT is operated with a voltage configuration that provides a gain of approximately 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT and the sensor surface is automatically scanned in xy steps of typically 0.5 mm with an attenuated laser beam. At an intensity of similar-to\sim1 photoelectron, 50-100 kHz laser rate and similar-to\sim106 gain, an anode current of IA(xy)510similar-tosubscript𝐼𝐴𝑥𝑦510I_{A}(xy)\sim 5-10italic_I start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_x italic_y ) ∼ 5 - 10 nA is measured for about 0.2 s at each scan position. In this case, the laser spot is tightly focused, but before the scan, the anode current is checked at different laser rates to avoid charge saturation. In this measurement, all anode pixels are shorted, and we measure the anode current at each xy position. Similarly to the QE scans, after typically four scanned x-rows, the laser spot is placed far outside the active PC area and the dark current Idarksubscript𝐼𝑑𝑎𝑟𝑘I_{dark}italic_I start_POSTSUBSCRIPT italic_d italic_a italic_r italic_k end_POSTSUBSCRIPT of the MCP-PMT is measured. Each raw anode current IA(xy)subscript𝐼𝐴𝑥𝑦I_{A}(xy)italic_I start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_x italic_y ) is corrected for the dark current Idarksubscript𝐼𝑑𝑎𝑟𝑘I_{dark}italic_I start_POSTSUBSCRIPT italic_d italic_a italic_r italic_k end_POSTSUBSCRIPT and scaled with the current IA,refpixsubscript𝐼𝐴𝑟𝑒𝑓𝑝𝑖𝑥I_{A,refpix}italic_I start_POSTSUBSCRIPT italic_A , italic_r italic_e italic_f italic_p italic_i italic_x end_POSTSUBSCRIPT and the gain Grefpixsubscript𝐺𝑟𝑒𝑓𝑝𝑖𝑥G_{refpix}italic_G start_POSTSUBSCRIPT italic_r italic_e italic_f italic_p italic_i italic_x end_POSTSUBSCRIPT at the reference pixel (e.g. 44 or 45) measured in the HV scan (see Section 3.1). After the correction for QE(xy)𝑄𝐸𝑥𝑦QE(xy)italic_Q italic_E ( italic_x italic_y ) at each measured position (see Section 2.2.2), this results in a QE-corrected gain distribution G(xy)𝐺𝑥𝑦G(xy)italic_G ( italic_x italic_y ) over the entire active PC surface.

G(xy)=IA(xy)IdarkIA,refpixIdarkGrefpix/QE(xy)𝐺𝑥𝑦subscript𝐼𝐴𝑥𝑦subscript𝐼𝑑𝑎𝑟𝑘subscript𝐼𝐴𝑟𝑒𝑓𝑝𝑖𝑥subscript𝐼𝑑𝑎𝑟𝑘subscript𝐺𝑟𝑒𝑓𝑝𝑖𝑥𝑄𝐸𝑥𝑦G(xy)=\frac{I_{A}(xy)-I_{dark}}{I_{A,refpix}-I_{dark}}\cdot G_{refpix}~{}/~{}% QE(xy)italic_G ( italic_x italic_y ) = divide start_ARG italic_I start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_x italic_y ) - italic_I start_POSTSUBSCRIPT italic_d italic_a italic_r italic_k end_POSTSUBSCRIPT end_ARG start_ARG italic_I start_POSTSUBSCRIPT italic_A , italic_r italic_e italic_f italic_p italic_i italic_x end_POSTSUBSCRIPT - italic_I start_POSTSUBSCRIPT italic_d italic_a italic_r italic_k end_POSTSUBSCRIPT end_ARG ⋅ italic_G start_POSTSUBSCRIPT italic_r italic_e italic_f italic_p italic_i italic_x end_POSTSUBSCRIPT / italic_Q italic_E ( italic_x italic_y ) (10)

Fig. 11 (left) shows the gain of the PHOTONIS SN 9002228 as a function of the PC position. It is clear that the gain is less homogeneous than the QE. It is generally significantly lower towards the rims of the tube. Here too, the black grid in the plot illustrates the position of the anode pixels of the PMT. Similar to Section 2.2.2, an average gain is calculated for each anode pixel from the various measurement positions and a max/min ratio of the gain is determined from these values. The results are shown in Fig. 11 (right) for five example PMTs. This approach can be used to quantify the gain variation over the active area of the MCP-PMT. Typically, >>>80% of the area has a gain max/min ratio of <<<2. This is also observed in many multianode dynode PMTs, and thus is not uncommon.

Refer to caption
Figure 11: QE-corrected gain as a function of PC position (left) for the PHOTONIS SN 9002228 with the black overlaid anode grid and the ratio of maximum gain to minimum gain (max/min) for a set of five PHOTONIS MCP-PMTs (right).

3.3 Gain inside Magnetic Field

The very compact PANDA detector leads to the requirement that the MCP-PMTs of the DIRCs must be placed inside a magnetic field of similar-to\sim1 Tesla, with the field lines crossing the PMT axis at an inclination angle θ𝜃\thetaitalic_θ of up to similar-to\sim20 degrees. Even under these conditions, the gain must be at least 106 to enable efficient single photon detection. This requires a measurement of the gain as a function of the magnitude and angle of the magnetic field.

The measurements were carried out on a dipole magnet with a field of up to 2 Tesla at the Julich Research Center in Germany. The B-field was monitored with a Hall probe. Unfortunately, the pole shoe gap of this dipole magnet is only 6 cm wide. To place the PMTs in this gap, we designed a special aluminum dark box with blackened side walls (Fig. 12). The total size of the box is 32×\times×19×\times×6 cm3, and it consists of four main parts: (1) a top cover with a small mirror mounted on the bottom at a 45 angle, (2) a light-tight box with a holder for inserting a lens and several cable feed-throughs for the high voltage and the anode signals, (3) four special screws for adjusting the tilt angle θ𝜃\thetaitalic_θ and (4) an interior slide which holds the MCP-PMT in place. The slide can be moved in x and y directions using micrometer screws. The screws can also be moved with stepper motors to enable automatic position scans (Section 9).

Refer to caption
Figure 12: Dark box with the most important parts for B-field measurements.

Fig. 13 shows a schematic representation of the setup for the measurements within the B-field. The MCP-PMT is illuminated with a laser beam that is attenuated to single photon level with ND filters. The light is transported into the box through an optical fiber, which is connected to a microfocus lens at its downstream end. In a special tube, the lens is pushed into the dark box and the photons are deflected at the top cover mirror by 90. By varying the lens position, the light is focused to a point on the PC of the MCP-PMT that corresponds to the center of one of the anode pixels, except for scans. The high voltages for the tube are fed into the box via SHV feed-throughs, while the voltage divider is placed outside the box. The measurements are performed at low laser rates (\leq5 kHz) to avoid gain saturation due to high local intensities. The anode signals from one or more pixels are read out via Samtec LSHM 140 (or 150) 50 ΩΩ\Omegaroman_Ω high-speed microcoaxial cables and amplified with a ×\times×20 amplifier (Phillips Scientific PS775) before being passed to a LeCroy WavePro oscilloscope which accumulates pulse height distributions with typically 50k - 100k events for each selected setting [HV, tilt angle, B-field]. The data sets are stored and analyzed in the same way as described in an earlier subchapter.

Refer to caption
Figure 13: Setup for measuring the gain as a function of the magnetic field.

For an MCP-PMT with 6 µm pores from Photek and one with 10 µm pores from PHOTONIS, the resulting gains for different B-fields and tilt angles are shown in the plots of Fig. 14. It can be clearly seen that the decrease in gain by a factor similar-to\sim3 between 0 and 1 T is similar in both PMTs. Beyond 1 T, the Photek PMT performs significantly better than the PHOTONIS tube, which is to be expected due to the smaller MCP pore diameter.

Refer to caption
Figure 14: Gain as a function of the magnetic field for different tilt angles θ𝜃\thetaitalic_θ for the Photek A1200107 and the PHOTONIS SN 9002192 MCP-PMT with 6 µm and 10 µm pores, respectively.

4 Time Resolution

A unique feature of MCP-PMTs is their excellent time resolution for single photons. Due to the compact dimensions and small path variations of the electrons on their way from the PC to the anode, the transit time spread (σTTSsubscript𝜎𝑇𝑇𝑆\sigma_{TTS}italic_σ start_POSTSUBSCRIPT italic_T italic_T italic_S end_POSTSUBSCRIPT) is typically much-less-than\ll50 ps. However, this is a lower limit as the true time resolution is impaired by some primary electrons bouncing back from the MCP-In electrode, which becomes visible as an additional time component arriving later than the main TTS peak (as shown in Fig. 17). For experimental applications, it is therefore better to also determine the mean time resolution σRMSsubscript𝜎𝑅𝑀𝑆\sigma_{RMS}italic_σ start_POSTSUBSCRIPT italic_R italic_M italic_S end_POSTSUBSCRIPT within a certain time interval, in our case from -0.5 to 2.0 ns around the TTS peak. This is actually the most important time resolution for many applications. Another approach has been chosen in [27] by fitting two Gaussians with σnarrowsubscript𝜎𝑛𝑎𝑟𝑟𝑜𝑤\sigma_{narrow}italic_σ start_POSTSUBSCRIPT italic_n italic_a italic_r italic_r italic_o italic_w end_POSTSUBSCRIPT and σwidesubscript𝜎𝑤𝑖𝑑𝑒\sigma_{wide}italic_σ start_POSTSUBSCRIPT italic_w italic_i italic_d italic_e end_POSTSUBSCRIPT to the time distribution.

The time resolution can be measured by illuminating a single pixel in the center of the MCP-PMT with laser light attenuated to single photon level (Npe@anodesubscript𝑁𝑝𝑒@𝑎𝑛𝑜𝑑𝑒N_{pe@anode}italic_N start_POSTSUBSCRIPT italic_p italic_e @ italic_a italic_n italic_o italic_d italic_e end_POSTSUBSCRIPT \approx 0.5 - 2). The anode signal is amplified 200-fold with an Ortec FTA820 device of 350 MHz bandwidth before being split into two parts with a simple passive 50 ΩΩ\Omegaroman_Ω splitter circuit. One signal branch is fed directly into the LeCroy WavePro 7300A oscilloscope (3 GHz; 20 GSa/s) to measure the integrated anode charge of the pulse, the other signal branch is passed through a leading edge discriminator (LeCroy 821, Phillips Scientific PS705, or PS707) which provides the timing signal at the oscilloscope. Also, the CAEN DT5742B digitizer is sometimes used to measure many waveform samples of the anode signals, which are later analyzed to obtain the time resolution. Due to the limited bandwidth (500 MHz) and sampling frequency (5 GSa/s) of the digitizer, the time resolution obtained is slightly worse than that determined with the oscilloscope. Data acquisition and analysis are automatic and controlled by Python scripts. A schematic representation of the time resolution setup is shown in Fig. 15, while a screenshot in Fig. 16 illustrates the most significant signals at the oscilloscope.

Refer to caption
Figure 15: Block diagram of the setup for the time resolution measurements.
Refer to caption
Figure 16: Oscilloscope screenshot of the signals used to determine the time resolution: amplified anode charge signal (red), laser trigger signal (yellow) and discriminated anode signal (blue). The time resolution is the jitter in the delay ΔtΔ𝑡\Delta troman_Δ italic_t between the yellow and blue signal. The time resolution of the laser trigger signal is <<<3 ps (RMS), and the internal time resolution of an oscilloscope channel was measured to be similar-to\sim5 ps (σ𝜎\sigmaitalic_σ).

By analyzing the jitter of the time delay between the anode time signal and the trigger signal of the laser controller, we can measure both σTTSsubscript𝜎𝑇𝑇𝑆\sigma_{TTS}italic_σ start_POSTSUBSCRIPT italic_T italic_T italic_S end_POSTSUBSCRIPT and σRMSsubscript𝜎𝑅𝑀𝑆\sigma_{RMS}italic_σ start_POSTSUBSCRIPT italic_R italic_M italic_S end_POSTSUBSCRIPT. The charge and delay of each photoelectron are stored in files, with one measurement containing similar-to\sim100k events. In the analysis, these charge and delay values are used for a time-walk correction, resulting in distributions as shown in Fig. 17. The obtained histograms are fitted with two Gaussian curves to roughly unfold the main TTS peak (left) and the recoil electron peak (right). The two distributions shown were measured with different HV divider configurations, where UPCMCPsubscript𝑈𝑃𝐶𝑀𝐶𝑃U_{PC-MCP}italic_U start_POSTSUBSCRIPT italic_P italic_C - italic_M italic_C italic_P end_POSTSUBSCRIPT differs by a factor of 4. This leads to a shift of the recoil electron peak, which shows that a larger voltage between PC and MCP-In reduces σRMSsubscript𝜎𝑅𝑀𝑆\sigma_{RMS}italic_σ start_POSTSUBSCRIPT italic_R italic_M italic_S end_POSTSUBSCRIPT.

Refer to caption
Figure 17: Time distributions (σTTSsubscript𝜎𝑇𝑇𝑆\sigma_{TTS}italic_σ start_POSTSUBSCRIPT italic_T italic_T italic_S end_POSTSUBSCRIPT, σRMSsubscript𝜎𝑅𝑀𝑆\sigma_{RMS}italic_σ start_POSTSUBSCRIPT italic_R italic_M italic_S end_POSTSUBSCRIPT) for a PHOTONIS MCP-PMT with different voltage divider configurations [5]: RPCsubscript𝑅𝑃𝐶R_{PC}italic_R start_POSTSUBSCRIPT italic_P italic_C end_POSTSUBSCRIPT:RMCPsubscript𝑅𝑀𝐶𝑃R_{MCP}italic_R start_POSTSUBSCRIPT italic_M italic_C italic_P end_POSTSUBSCRIPT:RAsubscript𝑅𝐴R_{A}italic_R start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = 1:10:1 (left) and 4:10:1 (right). The right-hand plot shows a clear improvement in σRMSsubscript𝜎𝑅𝑀𝑆\sigma_{RMS}italic_σ start_POSTSUBSCRIPT italic_R italic_M italic_S end_POSTSUBSCRIPT.

5 Rate Capability

The photoelectrons emerging from the PC are accelerated towards the MCP, where secondary electrons are generated during the amplification process in the MCP pores. The missing charge of these electrons must be restored in the MCP walls, which defines a MCP recovery or recharge time τ=RC𝜏𝑅𝐶\tau=RCitalic_τ = italic_R italic_C. The capacitance [𝒪𝒪\mathcal{O}caligraphic_O(100 pC)] depends on the MCP area and the resistance R𝑅Ritalic_R (tens of MΩΩ\Omegaroman_Ω) is rather large. Consequently, the recharge current is limited. If the amount of emitted secondary electrons is large, which is particularly the case in the second MCP, the gain can be reduced at high photon rates. This effect limits the rate capability of MCP-PMTs and depends on the photon intensity and the illuminated area. We measure the rate capability as a function of the anode current per area in two different ways: (1) illuminating the entire active PC area and directly reading the current with all anode pixels shorted (”current mode”); (2) illuminating only one anode pixel and analyzing the corresponding pulse height distribution (”pulse mode”). When comparing the results of the two modes, the CE is not taken into account, as it is >>>90% for the PHOTONIS MCP-PMTs examined here. Exemplary results of the rate capability obtained with both methods are shown in Fig. 18 for two recent 2-inch PHOTONIS MCP-PMTs with 2 ALD layers. The anode current Ianodesubscript𝐼𝑎𝑛𝑜𝑑𝑒I_{anode}italic_I start_POSTSUBSCRIPT italic_a italic_n italic_o italic_d italic_e end_POSTSUBSCRIPT per area A𝐴Aitalic_A is plotted against the gain relative to low rates. The quantity on the upper axis is the corresponding photoelectron rate R𝑅Ritalic_R \cdot cm-2 assuming a gain of G=106𝐺superscript106G=10^{6}italic_G = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT. The rate capability for each other gain G𝐺Gitalic_G can be found by applying the following trivial formula which is used to calculate R𝑅Ritalic_R

R=IanodeAGe𝑅subscript𝐼𝑎𝑛𝑜𝑑𝑒𝐴𝐺𝑒R=\frac{I_{anode}}{A\cdot G\cdot e}italic_R = divide start_ARG italic_I start_POSTSUBSCRIPT italic_a italic_n italic_o italic_d italic_e end_POSTSUBSCRIPT end_ARG start_ARG italic_A ⋅ italic_G ⋅ italic_e end_ARG (11)

In most cases, both measurement techniques provide the same result, with the gain reduction becoming relevant (<<<90%) at a photoelectron rate of less-than-or-similar-to\lesssim1 MHz/cm2. Some 1-inch MCP-PMTs show a rate capability of up to 10 MHz/cm2 [28, 29, 31]. Sometimes we also observe different results with both measurement methods. This could be due to an inhomogeneity of the resistivity across the active MCP-PMT area.

Refer to caption
Figure 18: Rate capability of the latest 2-inch PHOTONIS MCP-PMTs (SN 9002233 and SN 9002239) with two ALD layers. The results obtained with the ”current mode” and the ”pulse mode” are compared.

5.1 Measurement with Anode Current

Measuring the shorted anode current Ianodesubscript𝐼𝑎𝑛𝑜𝑑𝑒I_{anode}italic_I start_POSTSUBSCRIPT italic_a italic_n italic_o italic_d italic_e end_POSTSUBSCRIPT as a function of photon rate simulates the expectation in a real experiment where the entire PC surface is illuminated with photons. To measure this, we broaden the laser light with a square diffuser ED1-S20-MD from Thorlabs to cover the entire active sensor area. With an ND filter, the light intensity is reduced to 5 - 20 photons per laser pulse and entire PC area. In this type of measurement, all anode pixels are shorted and Ianode@fsubscript𝐼𝑎𝑛𝑜𝑑𝑒@𝑓I_{anode@f}italic_I start_POSTSUBSCRIPT italic_a italic_n italic_o italic_d italic_e @ italic_f end_POSTSUBSCRIPT is measured with a picoammeter at different laser frequencies f𝑓fitalic_f. In parallel, the current of a reference diode Idiode@fsubscript𝐼𝑑𝑖𝑜𝑑𝑒@𝑓I_{diode@f}italic_I start_POSTSUBSCRIPT italic_d italic_i italic_o italic_d italic_e @ italic_f end_POSTSUBSCRIPT is measured with a second picoammeter. This is necessary because the laser intensity does not increase proportionally to the laser frequency and must be corrected. The MCP-PMT is operated with a voltage that provides similar-to\sim106 amplification. The setup is shown schematically in Fig. 19.

Refer to caption
Figure 19: Block diagram of the setup for rate capability measurements with a shorted anode current at different laser frequencies.

A correction for the dark current is made for each measured frequency f𝑓fitalic_f. For example, after measuring of the anode current at a certain laser frequency, the dark currents of the PMT Ianode,darksubscript𝐼𝑎𝑛𝑜𝑑𝑒𝑑𝑎𝑟𝑘I_{anode,dark}italic_I start_POSTSUBSCRIPT italic_a italic_n italic_o italic_d italic_e , italic_d italic_a italic_r italic_k end_POSTSUBSCRIPT and the reference diode Idiode,darksubscript𝐼𝑑𝑖𝑜𝑑𝑒𝑑𝑎𝑟𝑘I_{diode,dark}italic_I start_POSTSUBSCRIPT italic_d italic_i italic_o italic_d italic_e , italic_d italic_a italic_r italic_k end_POSTSUBSCRIPT are measured. In relation to a reference point at a low laser frequency frefsubscript𝑓𝑟𝑒𝑓f_{ref}italic_f start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT, at which the MCP-PMT does not yet saturate, the relative gain Grelsubscript𝐺𝑟𝑒𝑙G_{rel}italic_G start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT can be calculated for each laser frequency f𝑓fitalic_f according to (12).

Grel=Ianode@fIanode,darkIanode@frefIanode,darkIdiode@frefIdiode,darkIdiode@fIdiode,darksubscript𝐺𝑟𝑒𝑙subscript𝐼𝑎𝑛𝑜𝑑𝑒@𝑓subscript𝐼𝑎𝑛𝑜𝑑𝑒𝑑𝑎𝑟𝑘subscript𝐼𝑎𝑛𝑜𝑑𝑒@subscript𝑓𝑟𝑒𝑓subscript𝐼𝑎𝑛𝑜𝑑𝑒𝑑𝑎𝑟𝑘subscript𝐼𝑑𝑖𝑜𝑑𝑒@subscript𝑓𝑟𝑒𝑓subscript𝐼𝑑𝑖𝑜𝑑𝑒𝑑𝑎𝑟𝑘subscript𝐼𝑑𝑖𝑜𝑑𝑒@𝑓subscript𝐼𝑑𝑖𝑜𝑑𝑒𝑑𝑎𝑟𝑘G_{rel}=\frac{I_{anode@f}-I_{anode,dark}}{I_{anode@f_{ref}}-I_{anode,dark}}% \cdot\frac{I_{diode@f_{ref}}-I_{diode,dark}}{I_{diode@f}-I_{diode,dark}}italic_G start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT = divide start_ARG italic_I start_POSTSUBSCRIPT italic_a italic_n italic_o italic_d italic_e @ italic_f end_POSTSUBSCRIPT - italic_I start_POSTSUBSCRIPT italic_a italic_n italic_o italic_d italic_e , italic_d italic_a italic_r italic_k end_POSTSUBSCRIPT end_ARG start_ARG italic_I start_POSTSUBSCRIPT italic_a italic_n italic_o italic_d italic_e @ italic_f start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT end_POSTSUBSCRIPT - italic_I start_POSTSUBSCRIPT italic_a italic_n italic_o italic_d italic_e , italic_d italic_a italic_r italic_k end_POSTSUBSCRIPT end_ARG ⋅ divide start_ARG italic_I start_POSTSUBSCRIPT italic_d italic_i italic_o italic_d italic_e @ italic_f start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT end_POSTSUBSCRIPT - italic_I start_POSTSUBSCRIPT italic_d italic_i italic_o italic_d italic_e , italic_d italic_a italic_r italic_k end_POSTSUBSCRIPT end_ARG start_ARG italic_I start_POSTSUBSCRIPT italic_d italic_i italic_o italic_d italic_e @ italic_f end_POSTSUBSCRIPT - italic_I start_POSTSUBSCRIPT italic_d italic_i italic_o italic_d italic_e , italic_d italic_a italic_r italic_k end_POSTSUBSCRIPT end_ARG (12)

5.2 Measurement with Pulse Height Distribution

Analyzing the pulse height distribution is another method for measuring the rate capability of an MCP-PMT. The photocathode is illuminated with a laser light spot that is attenuated to single photon level. An aperture of similar-to\sim5 mm diameter limits the spot size to one anode pixel, usually at position 44 (see Fig. 3 [left]) or elsewhere. All anode pixels except the illuminated one are shorted, and the charge spectrum is recorded for different laser frequencies f𝑓fitalic_f and intensities Npesubscript𝑁𝑝𝑒N_{pe}italic_N start_POSTSUBSCRIPT italic_p italic_e end_POSTSUBSCRIPT. The pulse height distributions are analyzed in the same way as described in Section 3.1. The fit provides the gain G𝐺Gitalic_G and the number of photoelectrons Npesubscript𝑁𝑝𝑒N_{pe}italic_N start_POSTSUBSCRIPT italic_p italic_e end_POSTSUBSCRIPT, which can be used to determine the anode current per area A𝐴Aitalic_A by

IanodeA=feGNpeAsubscript𝐼𝑎𝑛𝑜𝑑𝑒𝐴𝑓𝑒𝐺subscript𝑁𝑝𝑒𝐴\frac{I_{anode}}{A}=\frac{f\cdot e\cdot G\cdot N_{pe}}{A}divide start_ARG italic_I start_POSTSUBSCRIPT italic_a italic_n italic_o italic_d italic_e end_POSTSUBSCRIPT end_ARG start_ARG italic_A end_ARG = divide start_ARG italic_f ⋅ italic_e ⋅ italic_G ⋅ italic_N start_POSTSUBSCRIPT italic_p italic_e end_POSTSUBSCRIPT end_ARG start_ARG italic_A end_ARG (13)

where f𝑓fitalic_f and e𝑒eitalic_e are the applied laser frequency and the elementary charge, respectively. In contrast to the ”current mode”, this ”pulse mode” method only illuminates a relatively small area of the MCP-PMT to determine the rate capability, which can lead to slightly different results. In principle, the ”pulse mode” method can also be used to determine the rate capability as a function of the active surface. This has been done for a few tubes.

5.3 MCP Recharge Time

It is generally agreed that the rate capability depends on the charge recovery time τ𝜏\tauitalic_τ of the MCP pores. Illuminating the photocathode with LED double pulses with different time delays is a simple way to measure this time. The electrical double signals from a Philips PM-5770 pulse generator are used to drive a blue LED and trigger an oscilloscope. The LED is placed near the input window of the MCP-PMT, attenuated by an ND filter to a moderate light intensity of a few hundred or thousand photons per pulse and ”focused” by an aperture directly in front of the photocathode. The photons in each pulse are distributed over a time interval of typically 1.5 µs. Since the PMT response to each photon is very fast (similar-to\sim1 ns width), at low intensity many of the individual photons can even be visualized with a fast oscilloscope. In Fig. 20 (top) the averaged waveform is shown with two consecutive anode signals for a delay of similar-to\sim1 µs. Normally similar-to\sim1000 of these waveforms are stored, averaged, and further analyzed.

A first and surprising observation is the decay of the pulse heights with a time constant of τd<1subscript𝜏𝑑1\tau_{d}<1italic_τ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT < 1 µs. Since the pulser signal has a rectangular shape with a width of 1.5 µs, this effect must be due to saturation of either the LED or the MCP-PMT. LED saturation was ruled out by performing the same measurement with a Philips XP2262B dynode PMT, which shows neither signal decay nor gain reduction at high rates. On the other hand, the signal decay time τdsubscript𝜏𝑑\tau_{d}italic_τ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT must have a different origin than the MCP-PMT charging time τ𝜏\tauitalic_τ as it is too short. We assume that this decay is caused by space charges in the MCP pores when many photons hit the same pores almost simultaneously. The transit time in the MCPs and from the MCP-Out to the anode is 𝒪𝒪\mathcal{O}caligraphic_O(few ns) for an electron avalanche. Considering that for each LED pulse several thousand Npe/cm2 (\approx 𝒪𝒪\mathcal{O}caligraphic_O(100 Npe/pore)) are evenly distributed over a width of similar-to\sim1.5 µs (see Fig.  20), this multiplies to a time of 𝒪𝒪\mathcal{O}caligraphic_O(few hundred ns), in which the MCP pores and the region after the MCP-Out are filled with many electrons from the multiplication process and could lead to a lower gain due to a reduced electric field. This would mean that the measured decay constant τdsubscript𝜏𝑑\tau_{d}italic_τ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT is roughly comparable to the transit time of the secondary electrons multiplied by the number of photo electrons (Npe). However, more studies are needed to fully understand this signal decay. On the other hand, the rather small increase in charge between the end of the first and the beginning of the second pulse is due to the MCP charging time.

The waveforms are used to integrate the anode charge contained in each of the two pulses. Their charge ratio for a similar-to\sim106 single photon MCP gain is plotted against the time delay of the double pulse. The results for different photon intensities (Npe/cm2subscript𝑁𝑝𝑒𝑐superscript𝑚2N_{pe}/cm^{2}italic_N start_POSTSUBSCRIPT italic_p italic_e end_POSTSUBSCRIPT / italic_c italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT) are shown in Fig. 20 (bottom). The fitted function [1aexp(Δt/τ)]delimited-[]1𝑎𝑒𝑥𝑝Δ𝑡𝜏[1-a\cdot exp(\Delta t/\tau)][ 1 - italic_a ⋅ italic_e italic_x italic_p ( roman_Δ italic_t / italic_τ ) ] provides the pore recharge time τ𝜏\tauitalic_τ, which does not seem to depend on the intensities within the given uncertainties. The weighted mean value for the PHOTONIS MCP-PMT SN 9002085 with 25 µm pores is τ=(1.63±0.06)𝜏plus-or-minus1.630.06\tau=(1.63\pm 0.06)italic_τ = ( 1.63 ± 0.06 ) ms, which is smaller than the expected value τexp=RCsubscript𝜏𝑒𝑥𝑝𝑅𝐶\tau_{exp}=R\cdot Citalic_τ start_POSTSUBSCRIPT italic_e italic_x italic_p end_POSTSUBSCRIPT = italic_R ⋅ italic_C = 7.5 ms from the MCP resistance R𝑅Ritalic_R and the capacitance C𝐶Citalic_C. This roughly agrees with measurements of the recharge time in other tubes where τ=kτexp𝜏𝑘subscript𝜏𝑒𝑥𝑝\tau=k\cdot\tau_{exp}italic_τ = italic_k ⋅ italic_τ start_POSTSUBSCRIPT italic_e italic_x italic_p end_POSTSUBSCRIPT with k𝑘kitalic_k = 0.4 - 2 [32] was observed, where k𝑘kitalic_k should represent the properties of the unknown recharge circuit.

Refer to caption
Refer to caption
Figure 20: PHOTONIS SN 9002085: mean waveform (averaged over 1000 pulses) of an LED double pulse with 1.5 µs width and similar-to\sim1 µs delay (top); measured relative gain for different double pulse delays (bottom), fitted by the function [1aexp(Δt/τ)]delimited-[]1𝑎𝑒𝑥𝑝Δ𝑡𝜏[1-a\cdot exp(\Delta t/\tau)][ 1 - italic_a ⋅ italic_e italic_x italic_p ( roman_Δ italic_t / italic_τ ) ]. It shows the MCP recovery behavior for different photon densities (Npe/cm2subscript𝑁𝑝𝑒𝑐superscript𝑚2N_{pe}/cm^{2}italic_N start_POSTSUBSCRIPT italic_p italic_e end_POSTSUBSCRIPT / italic_c italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT), resulting in an average recharge time of τ𝜏absent\tau\approxitalic_τ ≈ 1.6 ms.

The data demonstrate that the relative gain at 1 µs delay drops more sharply for larger photon densities. In addition, measurements at similar photon densities show that the relative gain decreases with increasing illumination area. It confirms the observation that the rate capability and thus the charging behavior depend on both the photon density and the overall photon intensity (= illuminated area). The more the capacitor charge is emptied when a new photon arrives, the longer it takes for the pores to fully recover, which can lead to lower amplification. The same measurements were repeated with the PHOTONIS MCP-PMT SN 9002228 with 10 µm pores. The measured recovery time in this tube was τ=(2.67±0.16)𝜏plus-or-minus2.670.16\tau=(2.67\pm 0.16)italic_τ = ( 2.67 ± 0.16 ) ms with an expected value of τexpsubscript𝜏𝑒𝑥𝑝\tau_{exp}italic_τ start_POSTSUBSCRIPT italic_e italic_x italic_p end_POSTSUBSCRIPT = 8.7 ms. This tube also shows a pulse decay time of τdsimilar-tosubscript𝜏𝑑absent\tau_{d}\simitalic_τ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ∼ 1 µs. The data indicate that one of the dominating factors for the rate capability is the initial charge of the ”MCP-PMT capacitor” when a new photon is detected.

6 Escalation

Some of the recent MCP-PMTs from PHOTONIS with two ALD-layers sometimes enter a strange and not yet understood state that was never observed in former Planacon-type tubes without or with only one ALD-layer. This ”escalation” state [33, 34] can already start at quite low (few kHz single photons) or even without photon illumination when the MCP-PMT is operated at high gains (greater-than-or-equivalent-to\gtrsim5\cdot106). The latter indicates that the effect is not correlated to the rate capability of the tube. At high photon intensities, the effect can already trigger at an amplification just above 106. The most important observations at the onset of escalation are:

  • 1.

    massive rise of count rate and dark current (Fig. 21 left)

  • 2.

    significant gain drop due to the high rate (Fig. 21 right)

  • 3.

    drop in MCP resistance

  • 4.

    significant production of photons with a white spectrum

  • 5.

    suppression of the effect already in moderate B-fields

Close to the onset of ”escalation”, even a small increase of the MCP voltage by similar-to\sim10 V can cause a massive rise in the anode and MCP currents. When a digital camera or another PMT is placed opposite the ”escalating” MCP-PMT, it can be clearly seen that a huge number of photons are generated in the sensor, which explains the increased count rate. Further tests have shown that these photons are generated at or in the MCP layers or at the electrodes, but do not originate from the PC. At this stage, it is not clear whether the observed ”escalation” condition has a similar origin as the signal-induced noise reported in [35] and other references. By optimizing their production processes, PHOTONIS was able to shift the onset of escalation to non-critical operation voltages. So far, we have not come across an MCP-PMT from Hamamatsu and Photek that also shows this effect, but their available tubes are still without two ALD-layers.

Refer to caption
Figure 21: PHOTONIS 037P541: dark currents for different voltages at the two MCPs and between PC and MCP-In (left). PHOTONIS SN 9002236: fitted pulse height distributions for total PMT voltages UPCAsubscript𝑈𝑃𝐶𝐴U_{PC-A}italic_U start_POSTSUBSCRIPT italic_P italic_C - italic_A end_POSTSUBSCRIPT = 2230 V (top right) and UPCAsubscript𝑈𝑃𝐶𝐴U_{PC-A}italic_U start_POSTSUBSCRIPT italic_P italic_C - italic_A end_POSTSUBSCRIPT = 2240 V (bottom right). Both the steep rise in dark current by similar-to\sim3 orders of magnitude and the sudden drop in gain indicate the onset of ”escalation”.

In our detailed performance screenings of the latest 2-inch PHOTONIS MCP-PMTs, we check each tube for ”escalation”. If it does not escalate to a gain of similar-to\sim107, we assume that the PMT is safe for operation in the PANDA environment. Nevertheless, it would be highly desirable for the cause of the ”escalation” condition to be further investigated and understood.

7 Lifetime

The lifetime of MCP-PMTs is one of the most important issues for many recent applications. Until around 2010 [28, 36], MCP-PMTs were not an option for long-term experiments due to aging of the photocathode caused by feedback ions from the residual gas in the PMT. The prospects have changed since the MCP pores are coated with an ultra-thin layer of aluminum oxide and/or magnesium oxide using an ALD (atomic layer deposition) technique [18]. This drastically reduces PC aging and extends the lifetime of some MCP-PMTs by a factor >>>100 [34]. Meanwhile, MCP-PMTs are an attractive option for many high photon rate applications that require immunity to magnetic fields, very fast timing and high radiation tolerance. For this reason, it is important to measure the lifetime of MCP-PMTs. In our approach, this is achieved by illuminating the photocathode with constant low-intensity light and measuring the QE (and other observables) at regular time intervals. The QE is plotted against the integrated anode charge (IAC), and a decrease in QE indicates the onset of aging of the MCP-PMT. Data examples are presented in several publications [5, 37, 38, 39]. In this paper, we will focus on the description of our approach to measure the MCP-PMT lifetime.

7.1 Illumination Setup

In 2011, the Erlangen group launched an extensive illumination campaign with lifetime-enhanced MCP-PMTs. The aim was to measure the aging behavior of tubes of various designs and from different manufacturers in the same environment. In order to enable the fairest possible comparison of the different PMT types, we decided to set up a test station where the aging behavior of all available lifetime-enhanced MCP-PMTs could be long-term illuminated and measured in parallel [28, 29] and under similar conditions. This setup has already been briefly described several times in other papers [28, 40, 41], but will be presented in more detail here.

All included MCP-PMTs are permanently illuminated with a common LED light source at 460 nm. The special LED driver consists of a combination of an old Anritsu bit-generator and a self-built pulse generator, which generates square-wave pulses of similar-to\sim10 ns width and similar-to\sim70 V height. These electrical pulse signals are sufficient to trigger a very short LED flash. Unfortunately, the rather large and fast pulser signals also cause a considerable amount of stray electromagnetic radiation, which can interfere with all other ongoing measurements. For this reason, double-shielded Aircell7 cables had to be used to conduct the control signals from the pulser to the LED.

Under normal illumination conditions, the blue LED is driven at a frequency of 1 MHz and its intensity is attenuated to single photon level by neutral density filters. This is comparable to the expected boundary conditions at the PANDA DIRCs. At these illumination parameters and 106 MCP-PMT gain, each MCP-PMT accumulates similar-to\sim15 mC/cm2 IAC per day. The stability of the LED light intensity is monitored with a photodiode near the MCP-PMTs during the entire illumination period. In addition, a VME data acquisition system permanently reads out the pulse heights of some selected anode pixels of each PMT and records them to hard disk at a highly prescaled rate (216superscript2162^{-16}2 start_POSTSUPERSCRIPT - 16 end_POSTSUPERSCRIPT). This allows a straightforward determination of the integrated anode charge. A block diagram of the lifetime setup is shown in Fig. 22. For some MCP-PMTs, one half of the PC is masked to allow a direct side-by-side comparison of the aging progress on the same tube. A homogeneously illuminated area of about 30 ×\times× 30 cm2 at a distance of similar-to\sim50 cm from the blue LED is achieved with a square diffuser ED1-S50-MD from Thorlabs, followed by a spherical lens to further broaden the LED light. This currently allows us to illuminate up to sixteen 2-inch MCP-PMTs in parallel within the same setup (see Fig. 23).

Refer to caption
Figure 22: Block diagram with LED driver and readout as well as the DAQ electronics chain of the lifetime setup.
Refer to caption
Refer to caption
Figure 23: Image (left) and schematic representation (right) of the illumination setup [42] with the ED1-S50-MD square diffuser from Thorlabs.

7.2 Measurement Procedure, Analysis, and Results

After the MCP-PMTs have been illuminated for several weeks (or months), measurements of the gain, DCR and QE are performed for each tube. First, the photon rate is reduced to a few kHz, the prescaling in the readout electronics is removed and the distribution of the height of the single photon pulses for some selected anode pixels of all MCP-PMTs in the illumination setup is recorded. This results in the gain estimate of the MCP-PMTs. The DCR is also determined for each tube after the LED has been switched off. In the next step, all MCP-PMTs are transferred to the in-house monochromator [43], which is described in Section 2.2.1 and where a QE wavelength scan is performed in steps of λ𝜆\lambdaitalic_λ = 2 nm from 200 to 850 nm. If no further QE surface scans are planned, the MCP-PMTs are reinstalled in the illumination setup and the pulse height spectra are checked for consistency. After increasing the LED rate to 1 MHz and resetting the prescaling factor, the aging process is restarted by permanently illuminating the MCP-PMTs for several more weeks.

The charge and time distributions of the signals recorded continuously during the illumination are analyzed to determine the integrated anode charge for each MCP-PMT. The additional non-prescaled pulse height spectra are used to determine the gain and the DCR before and after each illumination period. Finally, QE, gain and DCR are plotted against the IAC, which provides the final results. In addition, measuring the entire QE spectrum between 200 and 850 nm allows an investigation of the evolution of the QE relative to a reference wavelength. As an example, the QE aging results are shown in Fig. 24 for the PHOTONIS MCP-PMT SN 9001332. The relative QE clearly shows that PC aging, once it has started, is more severe for red light than for blue light. Although it is not obvious for this MCP-PMT, we have also observed decreasing gain and DCR with progressive aging for some other tubes [29].

Refer to caption
Figure 24: QE and relative QE (normalized to 350 nm) as a function of IAC for the ALD-coated 2-inch MCP-PMT PHOTONIS XP85112 (SN 9001332). The QE is stable up to an IAC of about 10 C/cm2 and then decreases continuously. Obviously, the QE for red photons drops faster than for blue photons. Note that the relative QE for 550 nm and 600 nm is not plotted due to excessive fluctuations (see left graph).

Every few months, a QE surface scan is performed for each MCP-PMT in small steps of 0.5 mm over the entire PC area to identify potential regions on the PC where the QE may decrease faster. The used setup is the same as that described for the spatial QE scans in Section 2.2.2. The measurements are conducted at a wavelength of 372 nm or (in rare cases) 633 nm.

These QE scans provide quite interesting insights. We often observe (see Fig. 25) that the degradation of the QE starts at a certain PC position and grows from that point. The plots show the results of a QE scan and some x-projections at different IACs for the ALD-coated PHOTONIS SN 9001332. The right half of the PC was always covered during illumination. The QE degradation with increasing IAC is much more pronounced in the illuminated half, indicating that ion feedback is the main cause of PC aging.

Refer to caption
Figure 25: QE map (top left) after 15.4 C/cm2 IAC for the ALD-coated MCP-PMT PHOTONIS XP85112 (SN 9001332). The right half of the PC was covered during illumination. The QE x-projections of such maps along three discrete y-positions for different IACs show the evolution of PC aging.

In rare cases, significant ”QE holes” were observed immediately after delivery of the tube, usually on the PC rims and/or corners, which deteriorated within weeks or months, even when the tube was not illuminated. This indicates vacuum micro-leaks in the housing, probably caused by a faulty sealing, and these tubes had to be rejected. An example is shown in Fig. 26 with the Photek MAPMT253 MCP-PMT (SN A1200116), where the first QE scan was made immediately upon arrival of the tube, followed by others several months later. It can be clearly seen that the QE degradation develops from these ”QE holes”. This shows that the rather simple QE scans are quite a powerful tool to detect tiny vacuum leaks in the MCP-PMT housing.

Refer to caption
Figure 26: QE as a function of the PC position (top), recorded at different time points for the 2-inch Photek MAPMT253 MCP-PMT A1200116. The x-projections along different y-positions (bottom) show the evolution of the QE degradation over time, even when the PMT was not illuminated.

8 Position Scans with PADIWA/DiRICH/TRB DAQ

The conventionally assessed PMT performance parameters are QE, gain, time resolution, total DCR and sometimes the rate capability. Only rarely, the CE and lifetime of these tubes, for example, are measured. Especially in multianode MCP-PMTs, it is also very informative to investigate the less easily accessible internal parameters, such as the dark count rate per anode pixel, the spatial and temporal distribution of afterpulses triggered by feedback ions, the contribution of electrons bouncing back from the MCP input and the crosstalk (charge-sharing and electronic) between all anode channels. A more detailed explanation of the latter parameters can be found in the subsections 8.4 to 8.8. After setting up a sophisticated scanning device in combination with a suitable data acquisition system, we were able to investigate these parameters in comprehensive detail.

8.1 Setup

The MCP-PMTs are surveyed in large light-tight wooden boxes whose outer walls are covered with thin copper sheets to shield the interior from electromagnetic radiation. This is essential for clean anode signals from the PMTs, but also because the investigation of some performance parameters requires the measurement of currents in the sub-nanoampere range. A self-built 3-axis stepper (Fig. 27, top left) is installed in one of the dark boxes, which can be moved similar-to\sim40 cm in the x and y directions and similar-to\sim15 cm in the z direction. The large xy range allows us to measure four (and more) 2-inch MCP-PMTs in succession without having to change the setup. The stepper consists of 4 ”C-Beam®” linear actuators made of extruded aluminum, each equipped with a motor controlled by a ”TMC2130 Trinamic” stepper motor driver. The advantage of this driver is that it is equipped with sensorless motor load detection, which recognizes when the stepper motor hits a limit or other obstacle. The driver is controlled by a ”Teensy 3.5” microcontroller, which is programmed in an ”Arduino” development environment [30]. A stepper motor has 200 steps/revolution, while the lead screw used has a feed rate of 8 mm/revolution. By additionally activating a 16-fold microstep option, a theoretical distance of 2.5 µm/step can be achieved. When approaching a position always from the same side, a real positioning accuracy of \leq10 µm was measured with a CMOS camera. This accuracy is sufficient to deliver high-quality scans in our applications.

Refer to caption
Refer to caption
Refer to caption
Figure 27: Important parts of the setup for the position scans: 3-axis stepper (top left) and TRB3 board [47] (top right). The backplane of the DiRICH system shows the front with 3 MCP-PMTs attached (bottom left) and the rear with (from left to right) data concentrator, two DiRICH3 boards and power module (bottom right) [48].

A microfocus lens system is attached to the 3-axis stepper, which is connected to a PiLas laser head outside of the dark box by a single-mode optical fiber. The photocathode of an MCP-PMT is illuminated with ultrafast (σtless-than-or-similar-tosubscript𝜎𝑡absent\sigma_{t}\lesssimitalic_σ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ≲ 15 ps) blue or red laser pulses, and the light intensity is reduced to single photon level by neutral density filters. By adjusting the lens position along the z-axis of the stepper, the light spot can be focused to less-than-or-similar-to\lesssim10 µm FWHM peak width at the PC. In practice, a slightly worse focus is used to avoid gain saturation of the MCP-PMTs at the typically applied photon rates of 5 - 50 kHz.

The signal of each anode pad of a multianode MCP-PMT is read out directly and fed into FPGA-based PADIWA [44, 45] or DiRICH [46] frontend boards with 16 or 32 channels, respectively, which are located inside the dark box. The signals from the frontend boards are forwarded to a TRB3 (Trigger and Readout Board version 3) [45, 47] board, as shown in Fig. 27 (top right), outside the dark box, which takes over the further data management. In our applications, a PADIWA/TRB3 setup is used for reading out MCP-PMTs with a high anode segmentation (e.g., 100×\times×3 pixels) or in spatially limited environments such as measurements inside the Julich dipole magnet, while the standard quality test scans for the 8×\times×8 anode pixel PMTs are performed with a DiRICH/TRB3 setup.

Depending on the setup, the PADIWA boards can either be connected directly to the backplane of the MCP-PMTs or via Samtec HLCD high-speed coaxial cables. Each PMT signal is 7×\times× amplified and processed by an FPGA-based threshold discriminator, which switches an LVDS output as long as the signal is above the threshold. The LVDS signal of each channel is routed to a TRB3 board via simple ribbon cables. In addition to a central FPGA chip, each TRB3 board is equipped with four peripheral FPGAs which, in combination with the PADIWAs, can act as TDC to provide time and time-over-threshold (ToT) signals. In a single central FPGA, the trigger signal is generated with a Central Trigger System (CTS) and then distributed to the central FPGAs of any additional TRB3 boards. The clock signal is either provided by an external source or each TRB3 board generates its own clock, with the latter usually being applied. A TRB3 board has 256 usable data channels if no ToT information is required, and 4×\times×48 channels otherwise. It can process a maximum data readout trigger rate of up to 700 kHz. In principle, by combining many PADIWAs with enough TRB3 boards, it would be possible to read out an ”infinite” number of PMT channels.

In order to significantly reduce the cable clutter caused by the use of PADIWAs, DiRICH frontend boards were developed that can be connected directly to the PMTs via intermediate adapter backplanes (Fig. 27, bottom). One backplane can accommodate up to 4 MCP-PMTs with 64 channels each. A DiRICH channel is equipped with a similar-to\sim18 - 28×\times× (dependent on the pulse height) amplifier, a discriminator and a multihit TDC and provides coarse pulse height information through a ToT technique, all of which is implemented in an FPGA. The intermediate backplane is additionally equipped with a data concentrator and a power supply board. The data concentrator is connected to a TRB3 board outside the dark box via a fiber optic cable for the data and one or two Ethernet cables for the distribution of the trigger (and possibly the clock) signal to the DiRICH boards. The latter signals are generated in the CTS of the TRB3. The power module is connected to a 32 V power supply outside the dark box.

Refer to caption
Figure 28: Block diagram of the setup for the measurements and xy-scans with the DiRICH/TRB3 DAQ system.

A block diagram of the basic components of the scanning setup is shown in Fig. 28. The DiRICH/TRB3 DAQ system has a flexible design and offers a timing accuracy of up to similar-to\sim10 ps per single TDC channel. An internal ring buffer for each channel can (theoretically) store the timing information of up to 32 hits per trigger and handle burst hit rates of similar-to\sim30 MHz. During xy-scans the DiRICH/TRB3 system permanently analyzes the data stream provided by all anode pixels of the MCP-PMT. The lead time (= timing of the pulse leading edge) and the ToT information of each hit and TDC channel (corresponding to an anode pad) are continuously queued in the ring buffer and can be read out and stored within an adjustable time window (e.g. -10 to +10 µs) around a certain trigger time. In our applications, this is usually the trigger pulse of a PiLas laser, which for practical reasons is artificially shifted to t0 = 100 ns for each channel in the analysis.

8.2 Data Taking and Analysis

Before any accurate and rigorous measurement is possible with the DiRICH/TRB3 data acquisition system, the fine time delays of the individual TDC channels must be calibrated. Unfortunately, this calibration is temperature sensitive and needs to be repeated occasionally, especially if the ambient temperature has changed. Otherwise, the time resolution and linearity of the DAQ system can deteriorate considerably. Calibration is performed by triggering each channel with internal random time pulses generated by the CTS. This creates a statistical sample from which the fine time delay is calculated from longer delays and used to calibrate the sampling distances of the time axis. The next step is to set the thresholds for the incoming analog signals on each DiRICH channel. With the PMTs turned off, the offset (or threshold) voltages of the DiRICH channels are varied close to the edge of the noise band until a moderate count rate is measured for each channel. An interesting option in this context is the possibility to measure the pulse height distribution for each anode pixel by simply varying the signal threshold. Counting the hits per pixel for different thresholds gives the distribution shown in Fig. 29 (left). The derivative of this curve provides the analog pulse height distribution measured at each anode pixel (Fig. 29, right). This allows us to examine the average pulse heights arriving at each DiRICH channel. In most of our scans, we apply a common threshold for all channels. In standard measurements, this is typically set to 10 to 25% of the pulse height peak of the single photons in a central anode pixel.

Refer to caption
Figure 29: PHOTONIS SN 9002192 MCP-PMT. Left: measured counts for a fixed anode channel at different thresholds set in the DiRICH board. A threshold of 1000 digits corresponds to similar-to\sim3.5 mV pulse height. Right: derivative of this curve reflecting the pulse height distribution; the noise is clearly separated from the single photon peak [48].

The entire stepper/DAQ setup allows xy-scanning over the active surface of the MCP-PMT in steps down to similar-to\sim20 µm. The laser beam is always directed exactly perpendicular to the PC of the PMT in order to suppress internal reflections in the optical entrance window. Any effects of this kind are not taken into account in the analyses. At each xy position, all data collected for a certain number (usually between 10k and 100k) of laser triggers are read out. The data of each measured position is first stored in files with a special format (.hld) and in a further step converted into root files [49] for the final analysis. In combination with the DiRICH/TRB3 or the PADIWA/TRB3 DAQ system [46, 45, 44], we can read up to 256 anode pixels simultaneously with the DiRICH and more than 300 channels (192 per TRB3) with the PADIWA boards. In practice, the laser positioning along x and y with the stepper, data acquisition for a given number of triggers, file conversion and data analysis at each measured xy point is automated and can be controlled with appropriate Python scripts. The data obtained allows a view “inside the PMT” [50] in 3D (x,y,t) and many performance parameters can be measured as a function of the incident photon position. By performing an xy-scan (for the 8×\times×8 anode MCP-PMTs in steps of 0.5 or 1.0 mm) over the active surface of the MCP-PMT, the information x- and y-position of the laser spot and the lead time, ToT and channel number of each hit within a given trigger time window for each anode pad (channel) and each laser pulse is stored to disk. The spatial dark count distributions, the afterpulse distributions as a function of the xy position and the time after the trigger pulse (quasi the TOF of the feedback ions) and the spatial and temporal recoil electron distributions can be derived from this raw data. By counting the hits in narrow and wide time windows, the charge-sharing and electronic crosstalk behavior can be analyzed. These TRB scans make it possible to separate crosstalk events from those of recoil electrons and afterpulse hits.

Refer to caption
Figure 30: Information obtainable by PADIWA/TRB3 scans for the PHOTONIS XP85012 (SN 9002085) HiQE MCP-PMT [50]. Top: time distribution of all hits within a time window of [-10,1] µs for any anode pixel x4-y6. This time interval contains all hits of interest, e.g. the dark count region at [-10,0] µs. The trigger peak at 100 ns corresponds to photons of the laser pulse. Center: same time distribution, but zoomed to the time window [98,108] ns, which contains the trigger peak and the recoil electron hits just above 100 ns. Bottom: time distribution zoomed to the time range [80,580] ns with afterpulse hits that actually start at t0 greater-than-or-equivalent-to\gtrsim 105 ns (see also Figs. 37 and 38).

Figure 30 (top) shows the lead time distribution of the exemplary anode pad x4-y6 within a [-10,1] µs window for a HiQE PHOTONIS XP85012 (SN 9002085) MCP-PMT summed over all xy illumination points. A laser rate of 10 kHz was used, which corresponds to a delay of 100 µs between two laser pulses. Two distinct regions can be identified: the long flat distribution before the trigger peak at 100 ns contains pure dark count hits, while clearer discrete structures appear later. The DCR is determined by integrating the number of hits between -10 and 0 µs. Two enlarged time window regions are shown in the other plots. In these distributions, the number of hits measured in the anode pad x4-y6 is plotted for laser x-positions between 15 and 34 mm, which also cover the left and right adjacent anode pads. In the middle plots of Fig. 30, the time window of [98,108] ns is zoomed very close around the main trigger peak, which contains recoil electrons and charge-sharing events. At the bottom plots, the time window is zoomed to [80,580] ns, showing that the afterpulse hits arrive much later than the main trigger pulse at 100 ns.

The DAQ system also gives us access to the spatial distribution of different hit classes defined by specific time windows around the trigger peak. In Fig. 31 such distributions are shown for a high-resolution PHOTONIS MCP-PMT (ES440) with a 100×\times×3 anode grid. Shown is the spatial distribution of four classes of hits: (1) the dark count hits between [-10, 0] µs show a flat distribution in space; (2) around the main trigger time ([99.5, 100.6] ns) we obtain, as expected, a narrow peak at the pixel position of direct photoelectrons; (3) in the recoil electron time region ([100.6, 101.5] ns) a well-defined hump about 15 mm wide appears with a central hole due to dead time; (4) events in the afterpulse time region ([104, 500] ns) show a broad scatter with a maximum at the position of the read-out pixel. Please note that in this plot the anode pixel x57-y1 was read out and data for many x-positions from 14 to 44 mm were recorded in steps of 25 µm.

Refer to caption
Figure 31: Spatial distributions (dark count, trigger, recoil, afterpulse) along x for different time windows for the PHOTONIS ES440 MCP-PMT with 100×\times×3 anode pixels. The read-out pixel was x57-y1 while the laser spot was moved along x in 25 µm steps.

8.3 Relative Detection Efficiency

The detection efficiency over the entire active area of an MCP-PMT depends on the QE (and CE), the gain and the threshold applied to the anode signals. To estimate the efficiency that would be obtained in a real experimental setup, we compare the number of expected hits with the measured ones. For each xy illumination point, the number of measured signals within a narrow time window [99.5,100.5] ns around the main peak at 100 ns in the time distribution is divided by the number of expected laser triggers. The ratio obtained for each position is normalized to the highest value within the entire active area, which provides the spatial ”relative efficiency plots” as shown in Fig. 32. It can be seen that the relative efficiency is lower towards the PMT rims and between the anode pixels because (1) the gain is generally smaller at the rims of the tube, probably due to distorted electric fields, and (2) the signal between the anodes is lower due to the sharing of charges. Typically, the same absolute threshold is set to each channel of a tube, and the relative efficiency is highest somewhere in the center of the PMT. At higher threshold values, the efficiency between the pixels and at the rims decreases more. The net effect of different thresholds (10%, 25%, and 50% of the mean single photon pulse height at a central pixel) is shown for the PHOTONIS MCP-PMT SN 9002223 in the first three plots of Fig. 32. The relative efficiency map of a PMT (SN 9002225) with a low QE in some areas is also shown in Fig. 32 (bottom right). A low QE naturally leads to a significantly reduced detection efficiency (see top right and bottom left corners), as fewer photoelectrons per laser trigger exit the PC in this area.

Refer to caption
Figure 32: Relative efficiency as a function of position and signal threshold for two PHOTONIS XP85112 MCP-PMTs: SN 9002223 (top left, top right, bottom left) at different single photon peak height thresholds and SN 9002225 (bottom right) with QE holes in the corners.

8.4 Leakage Current and Dark Count Rate

In addition to dark count events (DCR) caused by thermally emitted electrons at the photocathode, we usually measure a typical DC leakage current of a few nA even when the MCP-PMT is not illuminated. The leakage current comes from a non-infinite resistance between the PC and the MCP-In electrode when a voltage is applied between them. Ideally, the leakage current is in the sub-nA range, because if it is much-greater-than\gg10 nA, this current can become problematic for measuring some parameters such as CE and spectral and spatial QE, where a very small photon-induced current at the MCP-In electrode needs to be accurately determined. Fortunately, the leakage current is not a problem under normal experimental conditions as it does not reach the anode and therefore does not affect the pulse detection of the MCP-PMT.

The thermally induced DCR decreases after the PMT has been placed in the dark. For the series production PHOTONIS MCP-PMTs, the stabilization time is typically 30 to 60 minutes. For this reason, we wait about 1 hour after closing the dark box before starting any measurement. After this period the DCR has usually stabilized at less-than-or-similar-to\lesssim1 kHz/cm2. A by-product of the xy-scans described above is a dark count map for all anode pixels of each investigated MCP-PMT. The DCR for a pixel is derived from the integrated number of hits in the time window from -10 to 0 µs before the actual trigger peak, scaled by 105 and finally divided by the laser rate at each laser position to obtain the DCR in Hz. At a typical laser trigger rate of <<<10 kHz, this time window is free of other background hits such as afterpulses and recoil electrons. As expected and shown in Fig. 30 (top) and Fig. 31, the DCR distribution is uniform in time and space. Fig. 33 shows the measured DCR/pixel for two recent MCP-PMTs from PHOTONIS with 8×\times×8 anodes. It varies between <<<100 Hz and similar-to\sim10 kHz, and we observe that pixels with high DCR are usually found at the rims or corners and not in the center of the anode region. This is the case for most of the scanned MCP-PMTs; a possible explanation could be a distorted electric field due to residues of the sealing material at the PMT rims. With a mean gain of similar-to\sim106 and a signal threshold of similar-to\sim20% of the single photon pulse height at a central anode pixel, the typical DCR, averaged over the entire active area, is between 100 Hz/cm2 and several kHz/cm2 for recently measured PHOTONIS MCP-PMTs.

Refer to caption
Figure 33: DCR maps per anode pixel for the 2-inch MCP-PMTs PHOTONIS XP85112-S-BA SN 9002224 (left) and SN 9002235 (right). The mean PMT gain was set to similar-to\sim106, with a threshold well below the typical single photon pulse height. The color-coded scale (z-axis) corresponds to the DCR/pixel in Hz.

8.5 Time Resolution

8.5.1 Time-over-Threshold (ToT)

As the time stamp for each analog signal is determined using a leading edge discrimination technique in the DiRICH boards, the measured lead time depends on the pulse height of the analog signal and has to be corrected for timewalk. Rough pulse height information is available with the Time-over-Threshold information, which is determined for each signal. The ToT is defined as the time interval in which a signal is above a fixed threshold value. This information is used in the analysis to apply a coarse timewalk correction and derive the true time resolution for an anode pixel. A typical ToT distribution before and after the timewalk correction is shown in Fig. 34. Technically, for each bin along the ”ToT” direction, the lead time distribution is fitted by a Gaussian distribution, which provides the correction for the ”lead time”. The corrected time peak is artificially shifted to 100 ns.

Refer to caption
Figure 34: Typical raw (left) and timewalk-corrected (right) distributions of lead time versus ToT in the region of the main laser peak for signals from a recent PHOTONIS SN 9002223 MCP-PMT. The two visible structures correspond to two classes of hits: The upper bump originates from standard unipolar pulses, the lower bump from bipolar signals, which are mainly caused by hits in close proximity to the edge of the read-out pixel. The latter seems to be a kind of electronic crosstalk between neighboring anode pixels, but the origin is not fully understood. On the other hand, real charge-sharing hits are found at the lower edge of the upper bump, which is reasonable because of their lower pulse heights.

8.5.2 TTS and RMS Time Resolution

It is known that in most MCP-PMTs the transit time spread σTTSsubscript𝜎𝑇𝑇𝑆\sigma_{TTS}italic_σ start_POSTSUBSCRIPT italic_T italic_T italic_S end_POSTSUBSCRIPT is well below 50 ps (see also Section 4). However, the more important parameter for real experiments is the RMS time resolution σRMSsubscript𝜎𝑅𝑀𝑆\sigma_{RMS}italic_σ start_POSTSUBSCRIPT italic_R italic_M italic_S end_POSTSUBSCRIPT within a certain time window around the TTS peak. For PANDA, σRMSsubscript𝜎𝑅𝑀𝑆\sigma_{RMS}italic_σ start_POSTSUBSCRIPT italic_R italic_M italic_S end_POSTSUBSCRIPT in a time window from -0.5 to 2.0 ns around the TTS peak was chosen as a performance criterion. The RMS time resolution depends on both the internal structure of the PMT and the applied voltage divider ratio. Fig. 35 shows pixel maps with the measured time resolutions σTTSsubscript𝜎𝑇𝑇𝑆\sigma_{TTS}italic_σ start_POSTSUBSCRIPT italic_T italic_T italic_S end_POSTSUBSCRIPT and σRMSsubscript𝜎𝑅𝑀𝑆\sigma_{RMS}italic_σ start_POSTSUBSCRIPT italic_R italic_M italic_S end_POSTSUBSCRIPT for one of the latest PHOTONIS XP85112-S-BA (SN 9002224) MCP-PMTs. The values shown are the time resolutions at the corresponding anode pixel, convolved with the resolution of the related DiRICH/TRB3 electronic channel. The electron recoil bump after the TTS peak (Section 4) is mainly responsible for the significantly poorer RMS time resolution. These plots also clearly show that the time resolution is worse for pixels at the rim of the active area. The given time resolutions are the mean of the entire anode pixel area. σTTSsubscript𝜎𝑇𝑇𝑆\sigma_{TTS}italic_σ start_POSTSUBSCRIPT italic_T italic_T italic_S end_POSTSUBSCRIPT is much better (50 - 60 ps) if only a local position (e.g., one laser spot) in the center of the pixel is analyzed.

Refer to caption
Figure 35: σTTSsubscript𝜎𝑇𝑇𝑆\sigma_{TTS}italic_σ start_POSTSUBSCRIPT italic_T italic_T italic_S end_POSTSUBSCRIPT (left) and σRMSsubscript𝜎𝑅𝑀𝑆\sigma_{RMS}italic_σ start_POSTSUBSCRIPT italic_R italic_M italic_S end_POSTSUBSCRIPT (right) time resolutions for each pixel of the PHOTONIS SN 9002224. The applied voltage divider ratio was RPCsubscript𝑅𝑃𝐶R_{PC}italic_R start_POSTSUBSCRIPT italic_P italic_C end_POSTSUBSCRIPT:RMCPsubscript𝑅𝑀𝐶𝑃R_{MCP}italic_R start_POSTSUBSCRIPT italic_M italic_C italic_P end_POSTSUBSCRIPT:RAsubscript𝑅𝐴R_{A}italic_R start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = 4:10:1. With a divider ratio of 1:10:1, σRMSsubscript𝜎𝑅𝑀𝑆\sigma_{RMS}italic_σ start_POSTSUBSCRIPT italic_R italic_M italic_S end_POSTSUBSCRIPT would be about 100 ps worse, while σTTSsubscript𝜎𝑇𝑇𝑆\sigma_{TTS}italic_σ start_POSTSUBSCRIPT italic_T italic_T italic_S end_POSTSUBSCRIPT would remain essentially unchanged.

8.6 Recoil Electrons

There is a certain probability that the photoelectron will not enter a pore directly, but will be scattered back at the MCP-In surface. The recoil electron can then reach the MCP at another location and trigger a delayed electron avalanche. The principle of this effect is sketched in the left part of figure 36.

PCesuperscript𝑒e^{-}italic_e start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPTRecoil esuperscript𝑒e^{-}italic_e start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT\vvE\vv𝐸\vv{E}italic_E
esuperscript𝑒e^{-}italic_e start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPTIonAP esuperscript𝑒e^{-}italic_e start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPTPC\vvE\vv𝐸\vv{E}italic_E
Figure 36: Schematic representation of signal development by recoil electrons (left) and afterpulses from feedback ions (right) [54].

The signals from recoil electrons become visible in the time region directly after the main trigger peak at 100 ns, where a clear structure can be seen in the center-right plot of Fig. 30. The number of entries is approximately constant from 101 to 102.5 ns. The latter time corresponds to photoelectrons elastically scattered normal to the surface. In addition, the number of hits decreases over several orders of magnitude up to similar-to\sim108 ns. In Fig. 37 (top left), the distribution x-position is plotted against time. It shows that the recoil electrons are scattered over an area around the read-out anode pixel. It is instructive to examine the spatial distribution of these hits by shifting the time window. In the other plots of Fig. 37 it can be observed that at a narrow time slice around the trigger peak only the area of the read-out anode pixel x3-y6 is populated, while at later times this pixel is also hit when the laser focus points to the closer vicinity of the pixel position. The spatial and temporal extent of similar-to\sim20 mm and similar-to\sim4.5 ns of these events fit well with expectations for recoil electrons that are backscattered from the MCP-In before being amplified. The events at even later times could be caused by electrons that are backscattered more than once. Fig. 37 (top left) also shows a significant population of the regions adjacent to the read-out pixel at 14 - 21 mm, but the timing is the same as for the main laser peak. These can be attributed to charge-sharing crosstalk hits where the charge cloud extends across the pixel boundaries.

Refer to caption
Figure 37: Temporal and spatial distribution of recoil electrons at the read-out pixel x3-y6 for the PHOTONIS XP85012 (SN 9002085). Top left: x-position vs time, regions where hits of recoil electrons and charge-sharing are expected are marked; top right: xy distribution for hits very close (99.6 to 100.4 ns) to the main trigger peak, corresponding to direct photoelectrons; bottom: xy distributions for temporally shifted events (101.6 to 102.4 ns [left] and 102.6 to 103.4 ns [right]), corresponding to recoil electrons.

8.7 Afterpulses

The DiRICH/TRB3 data acquisition system makes it easy to identify and quantify afterpulse hits. Afterpulses are caused by residual gas ions traveling in the opposite direction to the electrons and hitting the PC of the PMT. They can knock further electrons out of the PC, which are then also amplified in the MCPs, but arrive much later than the photoelectrons from the main trigger. The effect is sketched in the right part of Fig. 36. As can be seen in Fig. 30 (bottom) the arrival times of afterpulse hits range from similar-to\sim10 ns after the trigger peak at 100 ns to similar-to\sim1 µs. The afterpulse ions also reach the surroundings of the illuminated area and can be detected in adjacent pixels.

8.7.1 Afterpulse Time Distribution and Fraction

In Fig. 38 (top) the afterpulse TOF distributions of two different MCP-PMTs are compared. The TOF peaks in these plots can be attributed to light (e.g. H, He, …) and heavy (e.g. Cs and Pb) residual gas ions most likely generated in the first MCP layer. The broad continuum below these peaks is likely caused by ions scattered within the MCP pores, and the later broad hit accumulation originates from ions generated in the second MCP layer. The overall structure of the afterpulse TOF distribution of a single MCP-PMT is often difficult to interpret. However, it is possible to quantify the afterpulse rate for each MCP-PMT. To do this, we calculate the ratio of hits within the time interval from 105 to 505 ns to those within the trigger window from 99 to 104 ns, corrected for DCR in each case. This results in the fraction of afterpulse hits per anode pixel. The fractions shown in Fig. 38 (bottom left) indicate that there is no major difference over the active area. With <<<1%, the afterpulse fraction is rather moderate in the MCP-PMTs shown, but we have also encountered tubes where the fraction increased to much-greater-than\gg10%. In such PMTs, multiple hydrogen ion peaks are seen at constant time intervals (see Fig. 39), which are caused by afterpulses generated by the previous afterpulse ion. By measuring a large number of MCP-PMTs, we hope to get a better picture of the afterpulse behavior, which could help in the development of new MCP-PMTs with even fewer feedback ions.

Refer to caption
Figure 38: Afterpulse time-of-flight (TOF) distributions of the MCP-PMTs PHOTONIS SN 9002085 (non-ALD, top left) and SN 9002224 (ALD, top right). The trigger peak at 100 ns corresponds to direct photoelectrons (p.e.). Some of the other peaks can be assigned to feedback ions from the residual gas, whereas the weak ”retarded laser pulse” in the upper right plot is a laser feature (also visible in (Fig. 39)). Bottom left: afterpulse fraction at each pixel for SN 9002224; bottom right: afterpulse probability as a function of voltage and magnetic field.

It is still unclear whether the afterpulse component depends only on the gain or also on the applied voltage. We were trying to answer this question by taking measurements in a magnetic field. Since the gain is a function of the voltage, we analyzed the afterpulse fraction in a B-field, where the gain decreases at higher fields. The results are shown in Fig. 38 (bottom right). For large voltages (and gains), the data is similar to the gain over the magnetic field (Fig. 14). However, while the gain at 1 T has decreased by a factor similar-to\sim3 compared to 0 T, the afterpulse fraction only drops by a factor <<<2. These data could indicate that the afterpulse probability depends on both the total PMT voltage and the gain. Unfortunately, this argument becomes dubious at lower voltages, as a low gain and comparatively large amplitude afterpulse signal (see Section 8.7.2) passes the detection threshold more easily than a low gain and rather small amplitude photoelectron signal. This also has an effect on the measured afterpulse fractions at high magnetic fields.

8.7.2 Afterpulse Amplitude with Waveform Sampling

An unpleasant feature of the DiRICH/TRB3 DAQ is that a channel is practically dead for similar-to\sim30 ns after a hit is triggered. This means that hits arriving up to 30 ns after the main laser pulse are generally not detected and are strongly suppressed. In addition to recoil electrons, this mainly affects H+, H+2superscriptsubscriptabsent2{}_{2}^{+}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and He+ ions, which arrive about 15 to 30 ns after the photoelectrons. To further evaluate the extent and impact of these missing afterpulses on the performance of the MCP-PMT, several studies were performed by measuring long waveform samples with an oscilloscope and/or a CAEN digitizer. As a bonus, the waveforms also allow the charge amplitude of afterpulse hits to be determined.

An example is shown in Fig. 39, where the afterpulse hits are shown in a time versus signal amplitude plot for the lower-quality PHOTONIS MCP-PMT SN 9002227. While the TOF distribution shows a number of discrete structures, it is obvious that the amplitude of the afterpulse ions (e.g. H at 25 ns) is much higher than that for the photoelectrons (also visible in Fig.4 of [51]). This implies that each ion impact generates several electrons from the PC. An interesting observation is that at the same gain but different magnetic fields, the afterpulse spectrum looks different, both in terms of the time structure and the amplitude height, which is significantly lower at 1 T. For the parameter setting given in Fig. 39, which corresponds to the same gain of similar-to\sim1.5\cdot106, the afterpulse fraction in the time range shown is 27% at 0 T and 9% at 1 T. The different time structure has not yet been clarified, but the lower amplitude indicates that fewer knock-out electrons reach the MCP. This suggests that a high magnetic field reduces the afterpulse fraction and possibly also the damage that each ion does to the PC.

Refer to caption
Figure 39: Afterpulse TOF versus charge distributions for the MCP-PMTs PHOTONIS SN 9002227 at the same gain of similar-to\sim1.5\cdot106; top: 2200 V and 0 Tesla; bottom: 2350 V and 1 Tesla. The insets show the zoomed photoelectron region, which show that the PMT gain is essentially the same in both configurations. The regular time structure, especially in the upper plot, can be attributed to a ”cascade effect” of the feedback ions: an H afterpulse generates another H afterpulse, etc. (peaks at 25, 40 and 55 ns), but also an Mg/Al afterpulse (similar-to\sim85 ns) may induce another H afterpulse at 100 ns. Although the peak labeled as ”O” could also stem from CH4 [35], the abundance of oxygen in the ALD layers makes the latter more likely.

8.8 Crosstalk

Crosstalk between the anode pixels is one of the most disturbing effects of multianode PMTs. In MCP-PMTs, we observe three different types of crosstalk: charge-sharing, electronic, and oscillations. The first two were investigated with the DiRICH/TRB3 DAQ, while the latter was studied with a digital oscilloscope.

8.8.1 Charge Sharing

A special feature of MCP-PMTs is that many electrons emerge from the pores of the second MCP in different directions, creating a charge cloud that grows to a certain lateral extent until it reaches the anode plane. Its width depends on the internal structure of the MCP-PMT, in particular the distance between the MCP-Out and the anode. The magnitude of the electric and magnetic fields also has a significant influence on the width (see Section 9.2). If only one fired pixel per photon (=1 hit) was detected on the entire PMT surface, the full charge cloud was contained in this pixel. With 2 or more hits, the charge cloud is distributed over several anode pixels, being most pronounced between adjacent pixels. This is roughly sketched in Fig. 40.

esuperscript𝑒e^{-}italic_e start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPTesuperscript𝑒e^{-}italic_e start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPTPC\vvE\vv𝐸\vv{E}italic_E
\vvB\vv𝐵\vv{B}italic_Besuperscript𝑒e^{-}italic_e start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPTesuperscript𝑒e^{-}italic_e start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPTPC\vvE\vv𝐸\vv{E}italic_E
Figure 40: Schematic comparison of the signal development from PC to anode without (left) and with the presence of a magnetic field (right): events without crosstalk (green), which only hit one anode pad, and with charge-sharing crosstalk, which hits more than one anode pads (blue). In a magnetic field, the charge cloud is more strongly focused.

By counting the number of anode pixels hit in a narrow time window around the trigger peak at 100 ns, the charge-sharing crosstalk can be investigated. These hits coincide in time, even if the charge cloud is distributed over two or more anode pads. In Fig. 41 (left) we see the 8×\times×8 pixel structure when only one hit is measured in the PMT. With two simultaneous hits in Fig. 41 (right), we observe a population consisting mainly of the pixel borders. Events with more than 2 simultaneous hits would be visible in the pixel corners. The plots were created from xy-scans acquired with a similar-to\sim50 µm diameter laser spot in 0.5 mm steps. Analyzing the distributions of the charge-sharing events also allows estimating the charge cloud width, which varies from >>>1 mm (σ𝜎\sigmaitalic_σ) in absence of to much-less-than\ll0.5 mm in presence of a B-field.

Refer to caption
Figure 41: Charge-sharing crosstalk in the xy distribution for hits within 20 ns around the laser trigger for PHOTONIS XP85112 (SN 9002224): left: 1 anode hit; right: 2 anode hits.

8.8.2 Electronic

The TRB scans also allow a straightforward investigation and quantification of electronic crosstalk effects between the anode pixels. This is unique and easy to accomplish by illuminating a single pixel and then analyzing all other PMT pixels for hits in the time ranges of the main trigger peak (around 100 ns) and the afterpulses (\leq1000 ns). The data from standard position scans can be used for this, as the illuminated position is saved for each xy step. The results for two PHOTONIS MCP-PMTs (SN 9002224, SN 9002227) are shown in Fig. 42, where a threshold value of similar-to\sim20% of the single photon pulse height was selected. The top row shows the fraction of hits in each pixel, normalized to x4-y4, which was illuminated in a central area of 2.5×\times×2.5 mm2. The analyzed time window in the maps includes all hits from 99 to 999 ns. The detected hits in each pixel are corrected by a correspondingly scaled amount of dark events counted in the time window [-10, 0] µs. The pixels directly adjacent to x4-y4 may still be populated by hits from recoil and afterpulse signals, but already in the second column/row from the illuminated pixel onwards, no more real hits are to be expected, except for electronic crosstalk.

Refer to caption
Figure 42: Electronic crosstalk in all other pixels when only the central 2.5×\times×2.5 mm2 of pixel x4-y4 were illuminated. Top: crosstalk maps in % for the PHOTONIS SN 9002224 (left) and SN 9002227 (right) MCP-PMTs. Bottom: time distributions of hits registered in the more distant pixel x6-y5, which should contain pure electronic crosstalk hits. The regular time structure in the bottom right plot is due to electronic crosstalk induced by the much larger afterpulse signals (compare to Fig. 39).

The maps show that the fraction of electronic crosstalk is typically similar-to\sim1% or less, depending on the MCP-PMT and also on the configuration of the voltage divider (1-10-1 versus 4-10-1). In-depth analysis shows that direct photoelectron signals cause only a small amount of electronic crosstalk, while the much larger afterpulse signals lead to significant electronic crosstalk even at the PMT rim. As expected, pulses with a large amplitude tend to induce more electronic crosstalk. The plots in the bottom row of Fig. 42 show the temporal distributions of the signals induced by electronic crosstalk, which can vary substantially from tube to tube.

8.8.3 Oscillations

Another type of crosstalk is the so-called ”coherent oscillations”. Some years ago, J. Va’vra [52] pointed out this ”coherent oscillation problem” in older PHOTONIS Planacon tubes. The phenomenon occurs when several photons hit the PMT at the same time and can lead to fake anode hits. Inspired by this observation, we investigated the effect in various 8×\times×8-pixel MCP-PMTs. A schematic representation of the setup used can be seen in Fig. 43. The active surface of the entire PMT is illuminated with light of different intensities. Using simple masks with several holes across the PC surface exactly at pixel positions, we simulated the arrival of photons with various intensities and at different positions. Three pixels were read out with an oscilloscope.

Refer to caption
Figure 43: Left: block diagram for oscillation measurements and right: pixel map with illuminated (black) and read-out (colored) anode pixels. Masks with different amounts of open pixels were used.

The results obtained are shown in the form of oscilloscope screenshots in Fig. 44. The left image shows the waveforms for an older Planacon MCP-PMT (SN 9002108) with standard backplane, which shows the coherent oscillations when multiple pixels are hit simultaneously. This is particularly annoying because it could lead to fake hits in a real experiment. The number of possible fake hits was also studied with the TRB3 DAQ. In these measurements, half the area of the MCP-PMT was masked, while the open side was homogeneously illuminated with similar-to\sim1 detected photon/pixel. Although hits are only expected on the open half of the PMT, fake hits were also observed on the masked half due to crosstalk and/or the oscillations described above. A more detailed discussion of this fake hit problem can be found in ref. [53].

After we informed PHOTONIS about this problem, they designed a new backplane for their MCP-PMTs. The result for a PMT with modified backplane is shown in the screenshot of Fig. 44 (right). Obviously, ringing is significantly reduced with the new backplane and is no longer a problem for our application. It should be noted that the 2-inch Hamamatsu MCP-PMTs also suffered from oscillation problems at high photon counts, whereas this effect was much less severe with the Photek PMTs from the beginning.

Refer to caption
Figure 44: Oscilloscope screenshots of the photon-induced ringing in comparison between PHOTONIS SN 9002108 (standard backplane) and SN 9002150 (modified backplane) [53]. During the measurements, the entire PC was homogeneously illuminated with similar-to\sim7 photons per pixel and per laser pulse.

9 xy-Scans in Magnetic Fields

9.1 Scanning Setup

The dark box shown in Fig. 12 has been modified to also allow for xy-scans in a magnetic field. For these measurements, the micrometer screws are extended by long threaded spindles stabilized by two metal rails that extend well beyond the range of the B-field, as shown in Fig. 45 (right). The extension rails are connected to the x- and y-axis of the sledge in the dark box on one side and to a stepper motor on the opposite side. This arrangement allows to perform xy-scans in a magnetic field similar to those in the laboratory in absence of magnetic field. The TRB3 data acquisition system is the same as that used in the lab, but the DiRICH boards are replaced by PADIWA frontend cards connected to the MCP-PMT by Samtec high-speed LSHM140 connectors and HLCD coaxial cables and placed outside the B-field. The automatic measurements at many xy positions and the subsequent analyses are controlled via a Python script, which can also be used to set different B-fields and high voltages. This enables very efficient measurements even overnight, for example. The scan range does not cover the full active area due to space limitations within the dark box. Nevertheless, a scan range of almost 25×\times×25 mm2 in x and y direction is possible, which covers about 4×\times×4 pixels of an 8×\times×8 anode MCP-PMT and similar-to\sim50×\times×1 pixels (in x direction) with a 100×\times×3 anode MCP-PMT.

The laser focus is optimized as described in Section 2.2.2 for the QE setup, and the photocathode current is measured with a Keithley picoammeter for different positions. At optimal focus, the PC current is the lowest due to saturation effects in the PC. For scan measurements, we usually place the lens slightly out of focus to avoid saturation of the anode current.

Refer to caption
Refer to caption
Figure 45: Left: image of the dipole magnet with the dark box inserted. Right: image of the open dark box (same as in Fig. 12) with the extension rails in x and y directions and the stepper motors attached to the end of the rails [54].

9.2 Charge Sharing

Fig. 46 shows the crosstalk behavior at different B-fields for the PHOTONIS MCP-PMT SN 9002224 with 8×8 anodes. The high-resolution (50 µm steps) x-scan ran over four inner pixels, and all read-out pixels are summed in these plots. In both the 1-hit and 2-hit distributions, one can clearly identify the regions corresponding to the four anode pads. Only hits within a narrow time window of ±plus-or-minus\pm±0.5 ns around the main trigger peak at 100 ns are selected.

Refer to caption
Figure 46: Distributions of charge-sharing crosstalk within different B-fields, determined with a high resolution x-scan for hits within 1 ns around the laser trigger for PHOTONIS XP85112 (SN 9002224).

The regions between two adjacent pixels are filled with 2-hit events. This reflects the charge sharing crosstalk, and becomes less pronounced and narrower with increasing magnetic fields. The explanation for this is that the width of the charge cloud reaching the anode gets narrower within a magnetic field. The electrons between MCP-Out and anode are constrained onto helicoidal paths, so that at high B-fields there is almost no charge sharing effect. At 1000 mT the pixel boundaries are barely visible, the remaining 2-hit events are probably caused by electronic crosstalk.

9.3 Recoil Electrons

Similar studies can be carried out with the recoil electrons. The effects are shown for the PHOTONIS 946P541 2-inch MCP-PMT with 100×\times×3 anode pixels, which corresponds to a pixel pitch of similar-to\sim0.5 mm. Fig. 47 (left) shows the spatial expansion of recoil electrons with increasing magnetic fields. Obviously, the distribution of recoil electrons becomes much narrower at higher B-fields. This is due to the effect that the electrons are confined within a relatively limited area when a magnetic field is applied. At high fields, the recoiling electron falls back to almost the same position at which it started. If this recoil electron arrives at the same pixel, it can sometimes not be detected due to the similar-to\sim30 ns dead time of the DAQ channel described above (see section  8.7.2). This results in the holes at the position of the read-out pixel that can be seen in the spatial distributions. In contrast, the temporal distribution of recoil electrons is hardly influenced by a B-field. This is reflected in the distributions of Fig. 47 (right), where the slope of the tail at later times on the right side of the main laser peak (from 100.5 to 102 ns) shows only minor changes. The lower intensities for higher fields are due to the fewer recoil electrons from adjacent pixels reaching the read-out pixel x47-y2 as these are spatially more localized.

Refer to caption
Figure 47: Behavior of recoil hits at different magnetic fields for the PHOTONIS 946P541 with 100×\times×3 anode pixels. The anode pixel x47-y2 was read out while the laser position was changed in steps of 25 µm. Left: spatial distribution of recoil electrons within a time window of [100.4,101.5] ns; right: temporal distributions of recoil electrons.

9.4 Geometric and Lorentz Shift

The electrons moving in the MCP-PMT are subjected to the Lorentz forces generated by the electric field along the PMT axis (z) and the magnetic field inside the dipole magnet (e.g. y-z plane). This leads to significant electromagnetic focusing of both the photoelectrons (and any recoil electrons) between the PC and the MCP-In and the electron cloud exiting at the MCP-Out. When the PMT axis is tilted with respect to the direction of the B-field, local shifts of the center of gravity of the charge cloud in the anode plane in the x-direction (Lorentz shift) and in the y-direction (geometric shift) are also observed. The non-negligible magnitude of these effects was simulated for different magnetic fields and compared with our measurements [48]. This will be published in a forthcoming paper.

10 Summary and Conclusions

In this paper, we have presented a systematic approach to the detailed measurement of most performance parameters of MCP-PMTs that can be carried out within a reasonable time frame. In particular, we have focused on the description of the experimental setups and procedures used to measure the different parameters. In parallel, the analyses used and some representative results for each parameter were discussed.

Below is a list of the most important findings and lessons learned from our numerous measurements with MCP-PMTs:

  • 1.

    It is essential that the anode backplane of the MCP-PMT is very carefully designed to limit electronic crosstalk and coherent oscillations between the anode pixels.

  • 2.

    Whenever possible and not absolutely necessary, the illumination spots should not be focused too sharply on the PC during test measurements in order to avoid local gain saturation. This applies to QE scans as well, as the PC current can also saturate at high local intensity.

  • 3.

    QE position scans can be used as a good indicator for vacuum micro leaks in the PMT housing.

  • 4.

    The CE can be significantly increased to >>>90% by recapturing the electrons bouncing back from the MCP-In. However, this requires a higher voltage between PC and MCP-in to optimize the RMS time resolution.

  • 5.

    ALD-coated MCP-PMTs are more susceptible to gain drop in magnetic fields than devices without ALD.

  • 6.

    At high gain and illumination conditions, some of the MCP-PMTs with 2 ALD layers (MgO as secondary electron emitter and Al2O3 as resistive layer) started to emit photons in the visible range, which led to ”escalation” due to photon feedback with the PC.

  • 7.

    The typical rate capability of the investigated 2-inch MCP-PMTs is less-than-or-similar-to\lesssim1 MHz/cm2. A few tubes have shown locally different rate capabilities which may indicate inhomogeneities in the resistive ALD layer.

  • 8.

    The time resolution and the DCR tend to be slightly worse at the rims of the active area of the MCP-PMT.

  • 9.

    The fraction of afterpulses caused by ion feedback varies from much-less-than\ll1% to >>>10%. Normally, the ions identified by the time-of-flight from the MCP to the PC are dominated by H+ ions, but in tubes with a high afterpulse rate we also identify O and Mg/Al ions, presumably released by a defective ALD coating. At high afterpulse fractions, we observe further afterpulses triggered by the feedback ions.

  • 10.

    Electromagnetic focusing reduces the width of the charge cloud arriving at the anode plane, which lowers crosstalk due to charge sharing. The spatial expansion of recoil electron events is considerably diminished. The latter can lead to an increased local electron density, particularly at high photon rates, and impair the tube’s rate capability.

  • 11.

    None of the MCP-PMTs with 2 ALD layers in the lifetime setup has significantly lost gain at IAC \approx 10 C/cm2.

Our hope is that this paper will help to standardize the approaches used to measure the performance parameters of MCP-PMTs. This would facilitate the comparison of results obtained at different sites.

Acknowledgements

This work is supported by the German BMBF and GSI Darmstadt.

References

  • [1] T. Iijima, Status and perspectives of vacuum-based photon detectors, Nucl. Instrum. Methods Phys. Res. A 639 (2011) 137; https://doi.org/10.1016/j.nima.2010.11.127
  • [2] T. Gys, Micro-channel plates and vacuum detectors, Nucl. Instrum. Methods Phys. Res. A 787 (2015) 254; https://doi.org/10.1016/j.nima.2014.12.044
  • [3] P. Hink, Status and perspectives of vacuum-based photon detectors, review talk at RICH2016; https://indico.cern.ch/event/393078/
  • [4] A. Lehmann, Status and perspectives of vacuum-based photon detectors, Nucl. Instrum. Methods Phys. Res. A 1056 (2023) 168568; https://dx.doi.org/10.1016/j.nima.2023.168568
  • [5] A. Lehmann, et al., Latest Technological Advances with MCP-PMTs, J. Phys.: Conf. Ser. 2374 (2022) 012128; https://dx.doi.org/10.1088/1742-6596/2374/1/012128
  • [6] Hamamatsu Photonics K.K., 325-6, Sunayama-cho, Chuo-ku, Hamamatsu City, Shizuoka Pref., 430-8587, Japan, https://www.hamamatsu.com/jp/en.html
  • [7] Photek Ltd, 26 Castleham Road, St Leonards on Sea, East Sussex, TN38 9NS, United Kingdom, https://www.photek.com
  • [8] Photonis Netherlands BV, Dwazziewegen 2, 9301 ZR Roden, The Netherlands, https://www.photonis.com
  • [9] B.W. Adams, et al., A Brief Technical History of the Large-Area Picosecond Photodetector (LAPPD) Collaboration; https://dx.doi.org/10.48550/arXiv.1603.01843
  • [10] Incom Headquarters, 294 Southbridge Road, Charlton, MA 01507, U.S.A., https://incomusa.com/
  • [11] B. Singh, et al., PANDA Collaboration, Technical design report for the PANDA Barrel DIRC detector, J. Phys. G: Nucl. Part. Phys. 46 (2019) 045001; https://dx.doi.org/10.1088/1361-6471/aade3d
  • [12] C. Schwarz, et al., Status of the PANDA Barrel DIRC, JINST 15 (2020) C03055; https://dx.doi.org/10.1088/1748-0221/15/03/C03055
  • [13] F. Davi, et al., PANDA Collaboration, Technical design report for the endcap disc DIRC, J. Phys. G: Nucl. Part. Phys. 49 (2019) 120501; https://dx.doi.org/10.1088/1361-6471/abb6c1
  • [14] I. Köseglu, et al., Status of the PANDA Endcap Disc DIRC project, JINST 15 (2020) C03001; https://dx.doi.org/10.1088/1748-0221/15/03/C03001
  • [15] PANDA Collaboration, Technical Progress Report, FAIR-ESAC/Pbar, 2005
  • [16] PANDA Collaboration, Physics Performance Report, 2009, arXiv:0903.3905v1
  • [17] K. Gumbert, Performance-Messungen von PHOTONIS MCP-PMTs mit 3×\times×100 Anodenpixeln, Master Thesis (in German), Erlangen 2021
  • [18] D.R. Beaulieu, et al., Nano-engineered ultra-high-gain microchannel plates, Nucl. Instrum. Methods Phys. Res. A 607 (2009) 81; https://doi.org/10.1016/j.nima.2009.03.134
  • [19] https://www.nktphotonics.com/products/pulsed-diode-lasers/pilas/
  • [20] https://www.newport.com/f/cs260-uv-vis-monochromators
  • [21] https://www.tek.com/de/products/keithley
  • [22] B. Herold, Studien zur Photodetektion im KM3NeT-Projekt, Diploma Thesis (in German), Erlangen 2008
  • [23] D.A. Orlov, et al., High collection efficiency MCPs for photon counting detectors, JINST 13 (2018) C01047; https://dx.doi.org/10.1088/1748-0221/13/01/C01047
  • [24] https://de.teledynelecroy.com/oscilloscope/
  • [25] https://www.caen.it/
  • [26] E.H. Bellamy, et al., Absolute calibration and monitoring of a spectrometric channel using a photomultiplier, Nucl. Instrum. Methods Phys. Res. A 339 (1994) 468; https://doi.org/10.1016/0168-9002(94)90183-X
  • [27] J. Va’vra, et al., A 30 ps timing resolution for single photons with multi-pixel Burle MCP-PMT, Nucl. Instrum. Methods Phys. Res. A 572 (2007) 459; https://doi.org/10.1016/j.nima.2006.10.291
  • [28] A. Britting, et al., Lifetime-issues of MCP-PMTs, JINST 6 (2011) C10001; https://doi.org/10.1088/1748-0221/6/10/C10001
  • [29] A. Lehmann, et al., Lifetime of MCP-PMTs, JINST 9 (2014) C02009; https://doi.org/10.1088/1748-0221/9/02/C02009
  • [30] https://www.arduino.cc/en/software
  • [31] K. Inami, Recent development of fast timing PMT, 2015, Talk at workshop on forward physics and high-energy scattering at zero degrees; https://indico.cern.ch/event/407524/
  • [32] H. Kobayashi, et al., Evaluation of transient gain-drop and following recovery property on microchannel plate: Comparison between two evaluation methods, Nucl. Instrum. Methods Phys. Res. A 1053 (2023) 168355; https://doi.org/10.1016/j.nima.2023.168355
  • [33] D. Miehling, et al., Lifetime and performance of the very latest microchannel-plate photomultipliers, Nucl. Instrum. Methods Phys. Res. A 1049 (2023) 168047; https://doi.org/10.1016/j.nima.2023.168047
  • [34] S. Krauss, et al., Performance of the most recent Microchannel-Plate PMTs for the PANDA DIRC detectors at FAIR, Nucl. Instrum. Methods Phys. Res. A 1057 (2023) 168659; https://doi.org/10.1016/j.nima.2023.168659
  • [35] M. Andreotti, et al., Characterisation of signal-induced noise in Hamamatsu R11265 Multianode Photomultiplier Tubes, JINST 16 P11030 (2021); https://doi.org/10.1088/1748-0221/16/11/P11030
  • [36] N. Kishimoto, et al., Lifetime of MCP–PMT, Nucl. Instrum. Methods Phys. Res. A 564 (2006) 204; https://doi.org/10.1016/j.nima.2006.04.089
  • [37] T.M. Conneely, et al., Extended lifetime MCP-PMTs: Characterisation and lifetime measurements of ALD coated microchannel plates, in a sealed photomultiplier tube, Nucl. Instrum. Methods Phys. Res. A 732 (2013) 388; https://doi.org/10.1016/j.nima.2013.07.023
  • [38] K. Matsuoka, et al., Nucl. Instrum. Methods Phys. Res. A 876, (2017) 93; https://doi.org/10.1016/j.nima.2017.02.010
  • [39] G. Muroyama, et al. R&D on the Extension of the MCP-PMT Lifetime, JPS Conf. Proc. 27 (2019) 011021; https://doi.org/10.7566/JPSCP.27.011021
  • [40] A. Lehmann, et al., Significantly improved lifetime of micro-channel plate PMTs, Nucl. Instrum. Methods Phys. Res. A 718 (2013) 535; https://doi.org/10.1016/j.nima.2012.11.109
  • [41] M. Pfaffinger, et al., Recent results with lifetime enhanced microchannel-plate photomultipliers, Nucl. Instrum. Methods Phys. Res. A 912 (2018) 123; https://doi.org/10.1016/j.nima.2017.10.084
  • [42] M. Pfaffinger, Performance und Lebensdauer von MCP-PMTs der neuesten Generation hinsichtlich ihrer Einsetzbarkeit im PANDA-Experiment, PhD Thesis (in German), Erlangen 2019
  • [43] B. Herold and O. Kalekin, PMT characterisation for the KM3NeT project, Nucl. Instrum. Methods Phys. Res. A 626-627 (2011) S151; https://doi.org/10.1016/j.nima.2010.04.129
  • [44] C. Uğur, et al., A 16 channel high resolution (<<<11 ps RMS) Time-to-Digital Converter in a Field Programmable Gate Array, JINST 7 (2012) C02004; https://doi.org/10.1088/1748-0221/7/02/C02004
  • [45] A. Neiser, et al., TRB3: a 264 channel high precision TDC platform and its applications, JINST 8 (2013) C12043; https://doi.org/10.1088/1748-0221/8/12/C12043
  • [46] J. Michel, et al., Electronics for the RICH detectors of the HADES and CBM experiments, JINST 12 (2017) C01072; https://doi.org/10.1088/1748-0221/12/01/C01072
  • [47] http://jspc29.x-matter.uni-frankfurt.de/trbweb/uploads/trb3.jpg
  • [48] M. Böhm, Bau und Inbetriebnahme eines automatisierten Messstandes für MCP-PMTs, PhD Thesis (in German), Erlangen 2024
  • [49] ROOT Data Analysis Framework, CERN, https://root.cern/
  • [50] A. Lehmann, et al., Lifetime of MCP-PMTs and other performance features, JINST 13 (2018) C02010; https://doi.org/10.1088/1748-0221/13/02/C02010
  • [51] J. Va’vra, Picosecond timing detectors and applications; J. Phys.: Conf. Ser. 1498 (2020) 012013; https://doi.org/10.1088/1742-6596/1498/1/012013
  • [52] J. Va’vra, PID techniques: Alternatives to RICH methods, Nucl. Instrum. Methods Phys. Res. A 876 (2017) 185; https://doi.org/10.1016/j.nima.2017.02.075
  • [53] M. Böhm, et al., Performance of the most recent MCP-PMTs, JINST 15 (2020) C11015; https://doi.org/10.1088/1748-0221/15/11/C11015
  • [54] S. Krauss, Charakterisierung und Vergleich der aktuellen Microchannel-Plate Photomultiplier hinsichtlich der Nutzbarkeit im PANDA Barrel DIRC, Master Thesis (in German), Erlangen 2020