arxiv:2112.01691v1 [astro-ph.im] 3 dec 2021

41
Characterization of Low Light Performance of a CMOS Sensor for Ultraviolet Astronomical Applications Timothee Greffe a , Roger Smith a , Myles Sherman b , Fiona Harrison b , Hannah Earnshaw b , Brian Grefenstette b , John Hennessy c , Shouleh Nikzad c a Caltech Optical Observatories, California Institute of Technology, Pasadena, California, USA b Cahill Center for Astronomy and Astrophysics, California Institute of Technology, Pasadena, California, USA c Jet Propulsion Laboratory, California Institute of Technology, Pasadena, California, USA Abstract. CMOS detectors offer many advantages over CCDs for optical and UV astronomical applications, espe- cially in space where high radiation tolerance is required. However, astronomical instruments are most often designed for low light-level observations demanding low dark current and read noise, good linearity and high dynamic range, characteristics that have not been widely demonstrated for CMOS imagers. We report the performance, over tempera- tures from 140 - 240 K, of a radiation hardened SRI 4K×2K back-side illuminated CMOS image sensor with surface treatments that make it highly sensitive in blue and UV bands. After suppressing emission from glow sites resulting from defects in the engineering grade device examined in this work, a 0.077 me - /s dark current floor is reached at 160 K, rising to 1 me - /s at 184K, rivaling that of the best CCDs. We examine the trade-off between readout speed and read noise, finding that 1.43 e - median read noise is achieved using line-wise digital correlated double sampling at 700 kpix/s/ch corresponding to a 1.5 s readout time. The 15 ke - well capacity in high gain mode extends to 120 ke - in dual gain mode. Continued collection of photo-generated charge during readout enables a further dynamic range extension beyond 10 6 e - effective well capacity with only 1% loss of exposure efficiency by combining short and long exposures. A quadratic fit to correct for non-linearity reduces gain correction residuals from 1.5% to 0.2% in low gain mode and to 0.4% in high gain mode. Cross-talk to adjacent pixels is only 0.4% vertically, 0.6% horizontally and 0.1% diagonally. These characteristics plus the relatively large (10μm) pixel size, quasi 4-side buttability, electronic shutter and sub-array readout make this sensor an excellent choice for wide field astronomical imaging in space, even at FUV wavelengths where sky background is very low. Keywords: CMOS,SRI mK×nK, ultraviolet, dark current, glow suppression, dual-gain. *Timothee Greffe, [email protected] 1 Introduction For decades the detector of choice for ultraviolet (UV) space applications has been the microchan- nel plate (MCP) coupled to various kinds of 2D charge sensors. Space missions such as GALEX, Neil Gehrels Swift, and Hubble Space Telescope Cosmic Origins Spectrograph adopted MCP im- agers because their sensitivity extends into the far-UV (FUV), while the wide bandgap of their photocathode depresses sensitivity to optical photons and allows them to be operated at room tem- perature. However, MCPs have relatively low quantum efficiency (QE) of only 10% at 1500 ˚ A [1] [2][3], require high voltages in space where electrical breakdown is a concern, and are damaged 1 arXiv:2112.01691v3 [astro-ph.IM] 29 Mar 2022

Upload: others

Post on 16-Apr-2022

9 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

Characterization of Low Light Performance of a CMOS Sensor forUltraviolet Astronomical Applications

Timothee Greffea, Roger Smitha, Myles Shermanb, Fiona Harrisonb, Hannah Earnshawb,Brian Grefenstetteb, John Hennessyc, Shouleh Nikzadc

aCaltech Optical Observatories, California Institute of Technology, Pasadena, California, USAbCahill Center for Astronomy and Astrophysics, California Institute of Technology, Pasadena, California, USAcJet Propulsion Laboratory, California Institute of Technology, Pasadena, California, USA

Abstract. CMOS detectors offer many advantages over CCDs for optical and UV astronomical applications, espe-cially in space where high radiation tolerance is required. However, astronomical instruments are most often designedfor low light-level observations demanding low dark current and read noise, good linearity and high dynamic range,characteristics that have not been widely demonstrated for CMOS imagers. We report the performance, over tempera-tures from 140 - 240 K, of a radiation hardened SRI 4K×2K back-side illuminated CMOS image sensor with surfacetreatments that make it highly sensitive in blue and UV bands. After suppressing emission from glow sites resultingfrom defects in the engineering grade device examined in this work, a 0.077me−/s dark current floor is reached at160 K, rising to 1me−/s at 184 K, rivaling that of the best CCDs. We examine the trade-off between readout speedand read noise, finding that 1.43 e− median read noise is achieved using line-wise digital correlated double sampling at700 kpix/s/ch corresponding to a 1.5 s readout time. The 15 ke− well capacity in high gain mode extends to 120 ke−

in dual gain mode. Continued collection of photo-generated charge during readout enables a further dynamic rangeextension beyond 106 e− effective well capacity with only 1% loss of exposure efficiency by combining short and longexposures. A quadratic fit to correct for non-linearity reduces gain correction residuals from 1.5% to 0.2% in low gainmode and to 0.4% in high gain mode. Cross-talk to adjacent pixels is only 0.4% vertically, 0.6% horizontally and0.1% diagonally. These characteristics plus the relatively large (10µm) pixel size, quasi 4-side buttability, electronicshutter and sub-array readout make this sensor an excellent choice for wide field astronomical imaging in space, evenat FUV wavelengths where sky background is very low.

Keywords: CMOS,SRI mK×nK, ultraviolet, dark current, glow suppression, dual-gain.

*Timothee Greffe, [email protected]

1 Introduction

For decades the detector of choice for ultraviolet (UV) space applications has been the microchan-

nel plate (MCP) coupled to various kinds of 2D charge sensors. Space missions such as GALEX,

Neil Gehrels Swift, and Hubble Space Telescope Cosmic Origins Spectrograph adopted MCP im-

agers because their sensitivity extends into the far-UV (FUV), while the wide bandgap of their

photocathode depresses sensitivity to optical photons and allows them to be operated at room tem-

perature. However, MCPs have relatively low quantum efficiency (QE) of only 10% at 1500A [1]

[2] [3], require high voltages in space where electrical breakdown is a concern, and are damaged

1

arX

iv:2

112.

0169

1v3

[as

tro-

ph.I

M]

29

Mar

202

2

Page 2: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

by over-illumination. The latter issue has restricted many astronomical UV telescopes from view-

ing fields containing bright stars, particularly in the FUV where over-illumination causes severe

depletion of the photocathode.

The drawbacks of MCPs have motivated the development of Silicon-based UV imagers (CCD

and CMOS) for astronomy, where higher QE provides more sensitivity for a given telescope size,

enabling powerful observatories on moderate-sized platforms. The ability to view fields containing

bright stars and regions where high dynamic range is required, such as the Galactic plane and Mag-

ellanic Clouds, is another significant scientific advantage, especially for wide-field of view (FoV)

systems. On the technical side, avoiding high voltages and being able to match pixel sizes to the

specific application are additional positive factors. For wide FoV applications, CCDs and CMOS

detectors can be assembled in close butted mosaics with ∼85% fill factor, supporting imaging

applications on a larger scale than MCPs[2].

The Ultraviolet Explorer (UVEX) mission [4], which motivates the investigation documented

here, requires two 12K×12K focal planes, imaging in FUV(1390 − 1900 A) and NUV (2030 −

2700 A) passbands at a plate scale of 1.02′′ per 10µm pixel, and a high efficiency Long Slit Spec-

trometer covering the 1150 − 2650 A wavelength range, at 0.4 A/pixel sampling with a single

detector. The sky in the FUV is so dark that even a small amount of readout noise dominates

photon shot noise from sky in a typical 900s exposure, so the ideal sensor will have very low read

noise (∼ 2 e−) and low dark current (< 10−3 e−/s) at temperatures readily accessible with radiative

cooling (170K).

While CCDs can meet some of these requirements at beginning-of-life, CMOS image sensors

have very significant advantages [5]. Since charge is converted to voltage within the pixels, there

are negligible charge transfer losses, a problem which is particularly severe in CCDs when back-

2

Page 3: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

ground is low, and gets worse with radiation damage, requiring complex post processing based

on frequent recalibration as demonstrated on HST[6][7]. Also, CMOS imagers do not require

a shutter. While CCDs can also be operated without a shutter, the required subtraction of aver-

aged parallel overscan lines adds shot noise and delivers poor cancellation of image smear in the

presence of pointing jitter.

In the CMOS architecture studied here, the 5-transistor pixel design (Figure 2) eliminates dead

time between exposures. Charge transfer to the sense node defines the end of one exposure and

the start of the next, allowing charge collection to continue during readout without charge smear-

ing. We use Rolling Shutter mode which requires one frame time to clear charge at the start of a

sequence of exposures. Subsequent readouts between exposures of the same scene do not repre-

sent any overhead. An important application of this is to extend the dynamic range by combining

consecutive short and long exposures, e.g. pairing 3 s and 300 s exposures extends dynamic range

by 100 without dead time between exposures due to readout.

Another advantage of CMOS imagers is the flexibility they afford for subarray readout. Since

pixels are addressed directly, subarrays can be read at high cadence without disturbing signal in-

tegration on the remainder of the sensor. This feature further extends the dynamic range to bright

targets which would saturate in the normal frame time, and enables high time resolution studies

of rapidly varying sources. For space applications where guide stars are required, subarray read-

out can provide rapid centroiding of bright stars without interfering with the use of the rest of the

device for deep science exposures.

3

Page 4: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

Fig 1: The Mk×Nk Imager reticle consists of four modular building blocks which are photo-composed (“stitched”) to create custom dimensions in 1k increments. The sizes shown have beensuccessfully produced, for ground and space missions.

In this paper we report on the characterization of an SRI Nk×Mk CMOS sensor[8] made by

Sarnoff International for scientific imaging from space. This radiation tolerant design (Table 1)

has significant flight heritage (Parker Solar Probe, Solar Orbiter, SOLO HI EIS) [9] [10]. It can be

fabricated in 1k increments up to 4k×4k from the same reticles (Figure 1) and the compact address

logic allows 4k×4k tiles to be mosaicked with high fill factor, enabling large focal plane areas. The

device used in this work has a 4K×2K format, as used on the Europa Clipper mission. Ultimately

for large FOV astronomical applications such as UVEX a 4K×4K size is attractive. Such devices

have been made successfully for Lawrence Livermore National Laboratory’s National Ignition Fa-

cility [8]. All sizes have very similar electrical characteristics since they are fabricated using the

same reticles, at the same foundry, using the same processes.

Total Ionizing Dose > 100 KradNon-Ionizing Energy Loss > 3× 1011 protons/cm2 @ 63 MeV protons

Single Event Latchup > 100 MeV cm2/mg

Single Event Upset > 25 MeV cm2/mg

Table 1: Radiation tolerance of SRI Nk×Mk, reproduced from data sheet.

4

Page 5: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

2 Provenance of Sensor Under Test

The 4K×2K device used for this study was produced together with identical wafers for the Eu-

ropa Clipper mission [8]. The Jet Propulsion Laboratory’s Advanced Detectors, Systems, and

Nanoscience Group at JPL’s Microdevices Laboratory (MDL) applied their end-to-end post fab-

rication processing including bonding, thinning (to 7 ± 0.3µm), and delta-doping process using

Molecular Beam Epitaxy [11] [12] [13] [14] at wafer level. This process has been demonstrated

over the years on a number of CCD and CMOS formats to provide high and stable quantum ef-

ficiency across the UV and visible spectrum [13]. Every photon not reflected or absorbed in the

anti-refection (AR) coating is both detected and collected into pixels [12].

JPL applied various AR coatings and metal dielectric filters (MDFs), returning several die to

SRI for packaging and screening tests. The in-band UV QE and out-of-band rejection obtained

with the delta doping and direct deposition of MDFs will be described in a companion paper. An

earlier broadband AR-coated version was measured for Europa-Clipper band [15]

Unfortunately, pinhole defects in the photo-resist (that protects the device during the etching

of back surface to reveal the bond pads) caused damage at random locations on the delta doped

wafer. An engineering grade device without any coatings, but with tens of these ”etch-through”

defects was provided to Caltech for evaluation at lower temperatures than previously explored. At

these temperatures several strong glow sources became apparent.

We will show that the relatively simple on-board circuitry for row/column decoding (no timing

generator or ADCs) allows for sufficient glow suppression to meet the demanding dark current

requirement with ample margin (§5.5), while read noise rivals the best CCDs (§5.7). Linearity is

good (§5.2), inter-pixel correlation is low(§5.8), and dynamic range is exceptional when CMOS

5

Page 6: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

capabilities for dual gain (§4.3.1) and dual exposure duration are harnessed. Readout overheads

are insignificant.

3 SRI Mk×Nk CMOS detector

3.1 Pixel Topology

Each pixel is made up of a light-sensitive pinned photodiode, a transfer gate, and 4 transistors

to reset the sense node and select it during readout (Figure 2). A pixel read cycle begins by

resetting the sense node to the reference potential, PIX VREF by closing transistors RESET and

MIMSELECT. If MIMSELECT is high during the time charge is sensed, the MIM capacitance

is in parallel with the sense node increasing the conversion gain (e−/µV) and well capacity at the

expense of increased read noise (in e−). By reading out the voltage before and after connecting the

MIM capacitor, one can read the same signal in high then low gain, to provide low noise then high

signal capacity. Charge transfer and reset operations are performed for an entire row and cannot

be executed on single pixels. This prevents CCD-style Correlated Double Sample (CDS) timing

if using off-chip digital subtraction. To support CDS with a conventional short delay between

baseline and signal samples, sample-and-hold circuitry is provided at end of each column. This

allows analog subtraction of an entire row to be performed concurrently when reading out in rolling

reset mode.

The detector has the ability to reset and transfer charges from the photodiodes to their sense

nodes over the whole array concurrently, when a signal labelled SNAP is pulsed. This “Global

Shutter” mode results in concurrent exposure start and end for all of the pixels of the array.

6

Page 7: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

RESET

TRANSFER GATE

n+

MIM SELECT

PIX VREF

MIM CAP

PIX SUPPLY

N

NN

p-EPI

p-SUBSTRATE

p+

ℎ𝜗 BufferedVideo Signal

p-PINNING

N-Diodep-Well

SENSE NODE

p-Well

RowSelect

Fig 2: Five Transistors with Pinned Photo-Diode Pixel (5TPPD) Diagram. Reproduced from [16].

3.2 Output Buffer

The RAW video outputs are buffered by p-type transistors in a Source-Follower topology. We use

3.3 kΩ pull-up resistors to 3.3 V as the load. No other processing occurs: we do not invoke the

optional RC filtering.

Alternatively, the CDS video output pins of the detector can be buffered by n-type transistors,

also in a Source-Follower topology. In this case an analog CDS circuit located on the bottom of

every column, stores the reset levels of a row before subtracting the signal obtained after charge

transfer. This subtraction is done using switched capacitors resulting in a 4 e− noise floor (on

comparable pixel designs [16]). We prefer Line-wise Digital CDS since it delivers several times

lower noise floor ( §5.7) .

4 Experimental Setup

4.1 Cryostat

A cryostat with Liquid Nitrogen (LN2) cooling is configured for operation of the detector at tem-

peratures down to 130 K. A large printed circuit board, dubbed the Vacuum Interface Board (VIB),

7

Page 8: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

is located between the front and back sections, sealed with O-rings on each side. The VIB trans-

ports all required signals into the vacuum on internal layers as well as supporting electrical compo-

nents in vacuum (such as the zero-insertion-force socket housing the detector) and in atmosphere

(FPGA, connectors to detector controller, test points, and switches that support trouble-shooting).

The detector socket is thermally strapped to the nitrogen tank at 77 K, while the VIB has cut-outs

to minimize conduction to the outside world at room temperature. We tuned the thermal link

to the liquid nitrogen tank for 120 K equilibrium temperature without active heating. Detector

socket temperature is servo controlled to 140 K with millikelvin stability. Detector electronics are

mounted directly on the cryostat in close proximity to the VIB to limit the length of the video path.

Archon Controller

Xilinx FPGA

Vacuum Interface Board

Cold LN2 Bench at 77K

Detector ZIF Socket at 140K

SRI CMOS detector

Cryostat

Fig 3: CMOST Camera. The open cryostat is shown with detector inserted into its zero-insertion-force socket. The controller sits on top of the cryostat, and the Vacuum Interface Board providesrow and column address generation and routing of video signals.

8

Page 9: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

4.2 Camera Controller

The detector controller is an “Archon” made by Semiconductor Technology Associates Inc1. It

offers versatile configuration including clock and bias drivers, LVDS outputs, and 4 slots, each

with 4 video processing channels. An intuitive waveform definition language has been added by

Caltech Optical Observatories (COO).

Direct addressing of rows and columns is supported by the detector to allow efficient subarray

readout or user selectable raster scanning direction. This requires a mix of direct selection of the

1K pixel blocks and binary coding for row selection within those blocks. As this complexity is not

supported by the Archon controller, we have added an external FPGA to generate the addresses, a

Xilinx Zynq SoC on Picozed board which plugs into the atmospheric side of the Vacuum Interface

Board.

4.3 Readout Mode

The detector offers a large range of possible readout schemes thanks to its versatile addressing

topology [16]. In True Global Shutter mode, one must read out a whole frame after Global Reset

to acquire baseline pixels and then digitally subtract a newly acquired image after Global Charge

Transfer. This frame-wise digital CDS suffers from the long time interval between frame acqui-

sitions, and thus is susceptible to low frequency noise (offset drift). An alternative is to record a

single image then subtract the average of many dark frames to remove offset patterns. While the

averaged reference frame is less noisy, it fails to remove ∼ 12e− of kTC noise. To avoid these

additional noise sources, all performance measurements reported here have used Rolling Shutter

(line by line reset) and line-wise digital CDS.

1http://www.sta-inc.net/archon/

9

Page 10: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

Rolling Reset is widely used in CMOS detectors. Light is accumulating at all times, so the

transfer of charge from photodiodes to their sense nodes, defines both the end of one exposure

and start of the next. Each line is addressed sequentially. The sense nodes for all pixels in one

line are reset concurrently. In our ”line-wise digital CDS” scheme, baseline samples are digitized

sequentially for the entire line, then charge is transferred from photodiodes to all sense nodes

concurrently for the same line. The resulting signal levels for each pixel in the line are then

digitized sequentially. The process repeats for the next line. The difference between the baseline

and signal samples is calculated in the host computer post-facto. 1.5 s is required to scan a frame

with 2048 lines at 700 kpix/s, with 16 channels being read in parallel each servicing a block of 256

columns.

The 1.5 s frame time results in a small skew in exposure start times across the image area but

this has negligible effect in the long exposures anticipated in our application. When acquiring a

sequence of frames there is a 1.5 s overhead to reset the pixels at start of sequence (e.g. after

telescope slew) but no dead time between frames within a sequence. Table 2 and Figure 4 describe

a multi-frame sequence for a hypothetical 4 × 4 pixel array.

With exposure duration being defined by the time between charge transfers, it follows that

minimum exposure time is one frame time and longer exposures are executed by inserting a delay

between frame scans. Faster cadence exposures can be executed when reading a subset of rows

usually scanning between two preset pointers.

The Archon version used for these experiments was originally developed for CCD readout and

thus has an AC-coupled differential video input. The negative input is grounded while positive

input is connected to either the RAW or the CDS output of the detector selected by a mechanical

switch. (We prefer RAW, as noted.) The DC level of the JFET input stage (in the Archon, post AC

10

Page 11: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

Step Action Output1 Reset Sense Node2 Read Sense Node Baseline Image Low Gain3 Transfer charge from Photodiode to Sense Node4 Read Sense Node Signal Image High Gain

Table 2: Row readout sequence in Rolling Reset mode.Each line is read twice: immediately afterreset to capture the baseline and after charge transfer to capture the signal. Timing is shown inFigure 4

coupler) is set through an analog switch typically once per row. This level is chosen so that the

differential input voltage is near the beginning of the ADC input range.

Each channel is digitized by a separate 16 bit ADC running at 100 MHz. An analog filter limits

bandwidth into the ADC only to the extent required to prevent aliasing. The high sample rate

allows each channel to be operated in digital oscilloscope mode for diagnostic purposes. When

executing digital CDS, noise bandwidth is set by the number of consecutive 100 MHz samples

which are averaged per pixel. This averaging occurs concurrently in FPGAs on each video card.

The line-wise digital CDS arrangement avoids the kTC noise incurred when resetting the ana-

log CDS capacitors. The fact that baseline and signal have identical signal path assures good gain

matching for optimal rejection of low frequencies. Surprisingly, the 366 µs interval between base-

line and signal sample proves to be short enough that low frequency noise sources (bias voltage

drift; 1/f noise) do not negate these benefits.

PIXEL

RESETVCLK

Transfer

Frame 0

Expose

Frame 1

ExposeExpose

Tintegration

row

1

row

2

row

3

row

4

row

1

row

2

row

3

row

4

row

1

row

2

row

3

row

4

ResetSense Node

Charge Transfer

Row 1 Pixels: 1,2,3,4,5 1,2,3,4,5

Row 2 Row 3 Row 4

Irrelevant frame

Pixels: 1,2,3,4,5

1,2,3,4,5

Row 1

Reset level Signal levelReset level Signal level

Row 2 Row 3 Row 4Pixels: 1,2,3,4,5

Row 1 Row 2 Row 3 Row 4

Tintegration

1

2

3

4

Fig 4: Readout Diagram of the Rolling Reset Mode. Hypothetical 4 ×4 pixel array. The trueexposure time comprises the readout time that is fixed and a parametric exposure time Tintegration. VCLK is the Vertical Clock pulse that increments the Row Pointer to the next row.

11

Page 12: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

4.3.1 Dual Gain mode

For best noise performance the smallest possible sense-node capacitance is desired as this increases

the voltage change in response to a given charge. However the signal range is then limited by

available voltage swing. To allow greater signal range, a MIM capacitor in each pixel can be added

in parallel to the sense node capacitance by closing the appropriate analog switch (Figure 2). Since

the readout of the voltage does not affect charge on the sense node, it is possible to read the same

charge packet first with MIM capacitor isolated then again after it is connected. The sequence of

steps for this Dual Gain readout mode is detailed in Table 3.

The key here is to establish separate CDS baselines for both gain states. There is no kTC noise

problem since any charge trapped when opening the gain-select switch is measured at Step 4 and

vanishes when the switch is closed again at Step 7.

Frame time is almost doubled since each line is digitized twice as many times. Due to the

added time delay, the difference between the frames produced at Steps 9 and 2 provides a CDS

sample which is inferior to the difference between frames produced at Steps 6 and 4. However,

this slightly elevated read noise is irrelevant since the low-gain sample is only used for large signals

that are shot-noise limited.

5 Performance

5.1 Photon Transfer Curves and Well capacity

To infer well capacity and conversion gain (e−/ADU), pairs of identical exposures with increasing

integration time are acquired with temporally stable flux and moderate spatial uniformity. We then

subtract the two frames of each pair to remove common mode patterns due to illumination or fixed

12

Page 13: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

Step Action Output1 Reset Sense Node in Low Gain2 Read Sense Node in Low Gain Baseline Image Low Gain3 switch to High Gain4 Read Sense Node in High Gain Baseline Image High Gain5 Transfer charge from Photodiode to Sense Node6 Read Sense Node in High Gain Signal Image High Gain7 Switch to Low Gain8 Transfer charge left behind from Photodiode to Sense Node. KEY STEP9 Read Sense Node in Low Gain Signal Image Low Gain

Table 3: Row readout sequence to generate a dual-gain image. Each row is read four times togenerate two pairs of row per gain. A second charge transfer must be performed after switching tolow gain to recover the full signal range.

electrical patterns, leaving only the random noise component of the signal. Spatial variance is then

calculated for a region chosen to be free of bad pixels or other artifacts such as ROIC glow. This

procedure is repeated for both low and high gains.

100 101 102 103 104 105 106

Median Signal (e-)

100

101

102

103

Noise

(e-)

FWC

= ~1

5 ke

-

FWC

= ~1

20 k

e-

Readout Noise = 2.1 e-

Readout Noise = 10.1 e-

Readout Mode: DEFAULTTemperature: 140.0Kred + : Low gain mode 8.5 e-/adublue * : High gain mode 1.2 e-/adu

Photon Transfer Curve in Rolling Reset Mode

Fig 5: Photon transfer curves of low and high gain modes. Readout noise, system gain and FullWell Capacity can be directly read from this plot.

Conversion Gain (e−/ADU), the inverse of the slope of the the linear fit, is calculated separately

for each gain mode. These values are then used to convert signal and variance to values displayed

in electrons, for both gain states on the same Photon Transfer Curve (Figure 5). This plot format

is convenient to immediately highlight the signal level at which shot noise dominates.

When only spatial variance is used in photon transfer curves, total noise is underestimated due

13

Page 14: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

to interpixel correlations, so system conversion gain (e−/ADU) is over estimated. The correct

conversion gain is obtained when the covariance with adjacent pixels is included in the photon

transfer curve . Such interpixel correlations were observed only in low-gain mode (§5.8). The 8.0

e−/ADU conversion gain in low-gain mode has been corrected for Inter-Pixel Correlation (IPC)

throughout this paper resulting in a 6% reduction in dark current and read noise compared to that

inferred from simple photon transfer method using variance alone. The photon transfer curve then

yields readout noise of 2.1 e− and 9.5 e− and a Full Well Capacity of 15 ke− and 113 ke− in the

high gain and low gain modes, respectively.

5.2 Linearity

Linearity has been measured in both high and low gain modes by varying the exposure time while

the detector is illuminated with a constant flux. To do this, a red LED is attached to the LN2 tank

on a separate thermally regulated platform, operating slightly above 77 K. A bench current source

injects a current ranging from a few tens to hundreds of nanoamperes leading to a total power

dissipation well below a microwatt, resulting in very little efficiency variation due to junction self-

heating.

We infer that detector self-heating is responsible for some gain drift as readout cadence influ-

ences the level of the video signal. Sensor current consumption peaks at ∼ 70 mA during readout,

dissipating hundreds of milliwatts. This highlights the need for a more thermally conductive pack-

age to minimize the change in detector temperature due to self heating. The package in use is far

from optimal in this regard so significant performance improvement is probably possible. It is also

clearly beneficial to adopt clocking schemes which maintain the readout cadence even when idle

to minimize temperature drift.

14

Page 15: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

An initial transient in the signal is seen in Figure 6 after turning on the detector and contin-

uously reading it at a constant cadence. This transient appears only where there is illumination:

not in dark areas under a mask. This demonstrates that the CDS processing indeed corrects for

offset drift of the RAW video levels but cannot prevent the gain change. We explore this depen-

dence of gain on temperature in more detail in §5.3, finding that the video level settles to 1e− after

approximately 30 min.

0 500 1000 1500 2000 2500 3000Relative Time (s)

807.5

810.0

812.5

815.0

817.5

820.0

Mea

n Si

gnal

(e-)

Settling time and stability post-CDS series of 20s exposures

Dark area + offset for referenceIlluminated area

Fig 6: Initial transient after starting the readout of the detector at a constant (and fastest) framerate. Each data point is the average of a clean 128 × 128 window. The transient is only visible inthe illuminated area, implying that this is a gain change and not an offset.

In order to mitigate gain changes due to self-heating during the linearity experiment, the detec-

tor was read out at a very low duty cycle (at least 5 min per frame), so that the effect of self-heating

was negligible.

We estimated the non-linearity by fitting an affine function or a quadratic with non-zero value

at origin. We acknowledge that this is not ideal, but fixing a zero offset leads to a large mismatch

with our fit. In order to optimize both the residuals during our fitting process and the non-linearity

itself, we included a weighting factor, wi, in the estimator E, which is defined as

E =∑

w2i (p(xi) − yi)

2 =∑(

p(xi) − yiyi

)2

(1)

15

Page 16: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

where yi is the measured signal at exposure time xi, and p(xi) is the value of the linear fit at

exposure time xi. This prevents the gain correction from increasing steeply at the lower end of the

non-linearity curve when the denominator becomes smaller.

Figure 7 summarizes the linearity behavior in high gain (1.2 e−/ADU) mode. A linear fit

results in an RMS deviation of 1.4%, with larger disagreement for higher signal. The residuals

show clear curvature, motivating a quadratic fit to the data. Better agreement is achieved with

a quadratic fit, yielding an RMS of 0.2% and flatter residuals. We perform an F-test and find

that the quadratic fit is statistically required, with the null hypothesis of a linear fit rejected with

p = 2.2 × 10−13. This is consistent with typical CMOS sensors where non-linearity originates

in both the voltage-dependency of the PN junction capacitance that contributes to Sense Node

capacitance, and in the non-linearities of the source-follower itself [17]. The inset plot shows

linearity behavior for faint signals (obtained at reduced flux), demonstrating that the fit improves

to an RMS deviation of 0.1% and projects to 0 signal for null exposure time. (We must extrapolate

to zero-length exposure since minimum exposure time is the frame scan time which is 1.5 s when

reading at 700 kPix/s/ch.)

Figure 8 summarizes the linearity behavior in low gain (8.0 e−/ADU) mode. This shows sim-

ilar behavior: the linear fit yields an RMS value of 1.5%, while the quadratic fit gives better

agreement with RMS deviation 0.4% and is similarly a statistically-justified improvement over the

linear fit. Note that the faint signal is neglected since low gain mode will not be utilized for very

faint signals.

16

Page 17: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

0 500 1000 1500 2000 2500 3000 3500 4000 4500Exposure time (s)

0

5000

10000

15000

20000Si

gnal

(e-)

Linear fit R2=0.9985

Signal VS Exposure TimeLinear FitQuadratic FitData

0 20000

500Linear fit R2=0.999997

Faint Signal

0 2000 4000 6000 8000 10000120001400016000Signal (e-)

3

2

1

0

1

Devi

atio

n fro

m th

e fit

(%) Linearity

Linear Fit. RMS = 1.40%Quadratic Fit. RMS = 0.20%Faint Sig. Linear Fit. RMS = 0.09% 250 500 750

101 RMS = 0.09%

Faint Signal

Fig 7: Linearity in high gain mode. Signal is the median of a 256x256 window. Faint signalranges from 0 e− to 1000 e−. An F-test comparing a quadratic fit to a linear fit returns a p-value =2.2 × 10−13, confirming that the quadratic model provides a statistically improved fit.

0 500 1000 1500 2000 2500 3000 3500Exposure time (s)

0

25000

50000

75000

100000

Sign

al (e

-)

Linear fit R2= 0.9995

Signal VS Exposure TimeLinear FitQuadratic FitData

20000 40000 60000 80000 100000Signal (e-)

2

1

0

1

Devi

atio

n fro

m th

e fit

(%) Linearity

Linear Fit. RMS = 1.55%Quadratic Fit. RMS = 0.42%

Fig 8: Linearity in low gain mode. Signal is the median of a 256x256 window. Note that thefaint signal is neglected since low gain mode will not be utilized for very faint signals. An F-test comparing a quadratic fit to a linear fit returns a p-value = 2.6 × 10−12, confirming that thequadratic model provides a statistically improved fit.

5.3 Temperature dependence of the gain

Temperature stability requirements in flight will probably be driven by the temperature dependence

of the gain. To evaluate this, we recorded identical flat fields at moderate intensity for tempera-

tures varying between 140 K to 180 K. Light intensity was held constant while regulating LED

temperature to millikelvin precision. We allowed a settling time of 30 min between each 5 K step.

Figure 9 shows that the temperature dependence of gain is not linear but gradually flattens as the

temperature increases. Around 165 K, the gain varies by 0.15 %/K. It is not possible to determine

from this experiment whether the gain dependence on temperature is caused by variation in QE,

sense node capacitance, Source-Follower transconductance, or load resistance.

17

Page 18: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

It would be interesting to reproduce this experiment to characterize the temperature dependence

of the offset. While the per-row CDS processing removes any slow DC-offset variation, one could

track the baseline samples, however, the AC-coupling of our video acquisition board currently

prevents such measurement.

150 155 160 165 170 175 180Temperature (K)

7200

7300

7400

7500

Mea

n Si

gnal

(e-)

Temperature dependance of the gain

Gain VariationMean Signal of a 128x128 sub-window 0.05

0.10

0.15

0.20

Gain

Var

iatio

n (%

/K)

Fig 9: Temperature dependence of the gain between 150 K and 180 K. Around 165 K, the gainvaries by 0.15 %/K. The signal is computed as the mean of a 128x128 window.

5.4 Multiplexer (Mux) glow

Electronics surrounding the image area (primarily logic and pixel drivers on the left edge of the

chip) act as glow sources that become visible in long exposures. Figures 10 shows Dark Maps

generated from the difference of 18 hour and 1 hour exposures to suppress fixed patterns and

reveal me−/s effects.

It can be seen that light is emitted by the row select logic to the left of the image and the column

select logic well below the bottom of the image. The device under test also had several strong glow

sources in the image area to the right and below of the area displayed.

We suppress photo-emission from the address logic by reducing the levels of bias voltages dur-

ing exposure, where not required in the charge collection process. The only remaining biases are

VDD driving the logic and the Transfer Gate that keeps the charges in the photo-diode. Experience

shows that the three biases V Ref, V Reset High, and V MIM High must be kept on to ensure

18

Page 19: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

a valid bias of the Sense Node before readout. Some settling time was added after turning the

other biases on again and starting a readout of the image area. This settling time, though dozens of

seconds, is negligible compared to the long integration times required for glow to be visible. This

glow-mitigation technique greatly increased the area where dark current is less than 1 me−/s.

5.5 Dark current

Dark current has been measured at temperature between 140 K and 240 K. One hour was allowed

for settling between each temperature change. No hysteresis from heating or cooling the detector

is observed. We evaluate dark current within a sub-window in the region of the detector minimally

affected by glow. In spite of the very low signal, good signal to noise ratio is obtained by using a

large number of pixels and very long exposure times.

McGrath et al. [18] list different mechanisms as possible sources of dark current, each having

their own temperature dependence. We do not attempt to identify the dominant source, noting

only that the dark current decrease follows the expected Arrhenius law, until a a floor is reached at

0.08 me−/s for temperatures below 165 K.

Given 2 e− read noise and 900 s exposure time, the dark current can rise to 1 me−/s before

increasing total noise by 10%. These requirements imply that the operating temperature could be

as high as 184 K before shot noise on dark current becomes significant.

The striations to the right in Figure 10 (Bottom ”low-glow” Map) are due to light scattering

from a bright defect at the lower edge, which we assume will not be present in science grade

devices. A dozen or so larger, approximately semicircular white spots are are not due to cosmic

rays but appear to represent weak glow sites of order 2 me−/s/pixel at peak, affecting fewer than

1% of pixels (Figure 11).

19

Page 20: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

0 250 500 750 1000 1250 1500 1750 2000

0

200

400

600

800

1000

Dark Current Map - CMOS SRI MkxNk - 140K - 1me-/s contour - Biases ON

10 1

100

101

102

me-

/s

0 250 500 750 1000 1250 1500 1750 2000

0

200

400

600

800

1000

Dark Current Map - CMOS SRI MkxNk - 140K - 1me-/s contour - Biases OFF

10 1

100

101

102

me-

/s

Fig 10: Top Figure: 18 hours exposure in the dark while keeping electronics in its default state.Bottom Figure: “Low glow” 18 hours exposure in the dark while shutting down support electronics.Dark Maps of the top left quadrant of the detector show fine structures and glow sources in andaround the image area. Most of the glow coming from the left edge of the chip, where the logicand pixel drivers are located, is turned off by lowering the biases during the exposure. Chargecollection is unaffected.

5.6 Optimizing Speed and Noise Performance

The STA Archon controller digitizes pixels at a rate of 100 MHz and supports sample averaging in

real time for each channel to provide low-pass filtering. This is the digital equivalent of the analog

20

Page 21: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

Semi Circular Bright Spots visible on the array

Fig 11: Some bright spots show a peculiar semicircular shape with the same orientation.

140 160 180 200 220 240Temperature (K)

10 2

10 1

100

101

102

103

Dark

Cur

rent

(me-

/s)

e^(-Et/kT)3 e-/h threshold

Floor: ~0.077 me-/s

Dark Current vs Temperature

Fig 12: Dark current for a wide range of temperatures. A simple fit of an Arrhenius law shows theexpected temperature dependency.

integrator found in traditional CDS processors.

Readout noise tends to increase with higher pixel frequency, since noise bandwidth increases

when there is less sample averaging to reject high and mid frequency noise. Conversely, when pixel

rate is low the increased interval between baseline and signal samples results in poorer rejection

of low frequency noise processes which begin to dominate. To find the noise minumum between

these extremes, we explored multiple pixel frequencies from 700 kHz to 1.4 MHz for numerous

averaging window widths.

For any given pixel time, when the averaging window is too wide, the sampling of the video sig-

21

Page 22: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

nal begins before it settles sufficiently, with adverse effects on gain and linearity behavior. There-

fore we check Gain and linearity using low level flat illumination to infer changes in system gain

and thus the true noise in electrons.

20 40 60 80 100 120 140Window Size (Sample)

1.8

2.0

2.2

2.4

2.6

2.8

3.0

3.2

Noise

(e-)

Readout Mode: DEFAULTTemperature: 139.996K

Pixel Speed-Noise OptimizationPixel speed | Optimum Noise | Window Width (sample) 1.41 MHz | 2.10e- +/- 0.03e- | 33 +/- 2 1.30 MHz | 2.06e- +/- 0.03e- | 37 +/- 2 1.20 MHz | 1.99e- +/- 0.03e- | 42 +/- 2 1.10 MHz | 1.97e- +/- 0.03e- | 50 +/- 2 1.00 MHz | 1.92e- +/- 0.04e- | 56 +/- 4 0.90 MHz | 1.95e- +/- 0.02e- | 67 +/- 2 0.80 MHz | 1.89e- +/- 0.03e- | 76 +/- 3 0.70 MHz | 1.90e- +/- 0.03e- | 94 +/- 3

Fig 13: Pixel Speed Noise optimization. The noise is defined as the spatial standard deviation ofa 100x100 sub-window of the array. For each pixel frequency, an optimum number of samples isfound by fitting a parabola to the data.

When the window becomes too wide, the gain decreases and the noise rises. For each pixel

frequency, there is therefore an optimal window width at the bottom of each curve in Figure 13. In

this paper, we only use the optimum sample width for any given pixel frequency.

5.7 Read noise

CMOS sensors can exhibit significant spatial variation in readout noise, predominantly due to sin-

gle electron traps close to the channel of the pixel buffer MOSFET. On some pixels, this creates

a bistable video level known as Random Telegraph Noise (RTN), whose amplitude and character-

istic time constant vary from pixel to pixel, with characteristic switching rate being dependent on

temperature.

In Figure 16 a time series of a pixel exhibiting strong RTN can be seen jumping between three

22

Page 23: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

states after CDS processing, representing a negative transition between CDS samples, a positive

transition, or no transition.

The noise map in Figure 14 was created by computing the (temporal) standard deviation of

each pixel from 94 short dark exposures.

0 500 1000 1500 2000 2500 3000 3500 4000

0

250

500

750

1000

1250

1500

1750

2000

Temporal Noise Map over 94 100ms dark frames, high gain mode . Fpix = 1.1 MHz

0.0

0.5

1.0

1.5

2.0

2.5

3.0

Noise

(ele

ctro

ns)

Fig 14: Noise Map of the detector in high gain. Excess noise is visible in the bottom rows, next tothe amplifiers. The bad column generates noise in its surrounding as well as four distinct clusters.

The histogram in Figure 15, was created from data in subset at top left of the temporal noise

map, chosen to exclude defective areas. Instead of following the ideal Gamma distribution, the

measured distribution is skewed with many pixels being noisier than expected.

Given this highly skewed distribution, it may not be obvious what is the best metric for noise

performance. The modal noise is 1.33 e−, the median noise is 1.43 e−, and the mean is 1.70 e−.

We propose that a better metric is the square root of the average variance across all pixels not

excluded as “too noisy to be used”. The square of this “RMS Noise” times the number of pixels in

23

Page 24: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

a photometric aperture then represents expected total variance which can be used for making SNR

projections. Figure 17 shows the RMS Noise as a function of the number of pixels included. As

more pixels are classified as “bad” the RMS Noise improves but at the expense of lost information

where interpolation across bad pixels is employed.

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24Noise (electrons)

100

101

102

103

104

105

106

Pixe

l Cou

nt

Pixel Noise distributionPixel speed

700 kHz1.1 MHz1.4 MHzgamma pdf

0

20

40

60

80

100

Frac

tion

of p

ixel

s (%

)

Fig 15: Noise distribution of pixels in a clean sub-array of the detector. The ideal gamma distri-bution describes distribution if noise were distributed normally in the detector. The dashed curvesrepresent the fraction of pixels (right y-axis) below a given noise level (x-axis)

0 20 40 60 80 100frame #

20

10

0

10

20

leve

l (ad

u)

RTN pixel

Fig 16: Fluctuating level of a pixel affected by Random Telegraph Noise. The amplitude of the stepis much higher than the read out noise. Note that the size of the step is constant while transitiontiming is random.

24

Page 25: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85 90 95 100Fraction of pixel used (%)

0.50

0.75

1.00

1.25

1.50

1.75

2.00

2.25

2.50

RMS

Noise

(e-)

RMS Noise vs fraction of pixel used - Sub-windowPixel speed

700 kHz1.1 MHz1.4 MHz

Fig 17: Square root of average noise variance for all pixels included in window, after exclusion ofnoisiest pixels, plotted for three different pixel rates.

5.8 Inter-Pixel Correlation (IPC) - Cross-correlation

Pixels may couple to their neighbors via two main effects [19][20]: 1) Brighter Fatter Effect (BFE)

[21] [22] is the tendency for previously accumulated charge to repel incoming charge, allowing

neighboring pixels to collect additional electrons; 2) Inter-Pixel Capacitance leads to voltage cou-

pling between neighboring pixels [23]. The method described below is used to evaluate the total

contribution of inter-pixel effects (BFE and Inter-Pixel Capacitance) to cross-correlations with-

out distinguishing them. This total effect will be referred to as the Inter-Pixel Correlation (IPC).

A method to distinguish underlying inter-pixel phenomena using correlations in flat fields is de-

scribed in papers by Hirata et. al. [19] and Choi, et. al. [23].

We extracted the impulse response from a difference of two flat frames using a 2D auto-

correlation method. To limit errors due to illumination non-uniformity and imperfections within

the device itself, the area we used for processing was limited to well chosen 256 × 256 pixel sub-

arrays. This is only enough to provide a ∼ 0.4% error bar with our results according to [19]. We

performed a Monte Carlo simulation using fake shot-noise limited frames with no IPC to confirm

25

Page 26: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

this error bar. Forty-eight pairs of frames were then combined to reduce the errors. The normalized

kernel of the inferred impulse response is shown in Figure 18.

H Pixel

0 2 4 6 8 10V Pix

el

02

46

810 No

rmal

ized

Auto

corre

latio

n (%

)

0.00.51.01.52.02.53.0

Normalized Auto-correlation of a difference of flat frames

3 4 5 6 7 8H Pixel

3

4

5

6

7

8

V Pi

xel

0.0 0.0 0.0 0.0 0.0

0.0 0.1 0.4 0.1 0.0

0.0 0.6 97.5 0.6 0.0

0.0 0.1 0.4 0.1 0.0

0.0 0.0 0.0 0.0 0.0

Impulse response derived from Auto-correlation

Fig 18: Left: Auto-correlation of the difference of two flat frames. The noise is reduced by aver-aging 48 256 × 256 pixel patches. Right: Impulse response or IPC reconstructed from the Auto-correlation in low gain mode. In High-Gain mode, no significant inter pixel coupling is measured.

We reproduced this analysis for a range of illumination levels (Figure 19). We define the

horizontal (vertical) correlation as the ratio of the first horizontal (vertical) neighbors over the

central peak. Here, the noise associated with each sample is evaluated by computing the standard

deviation of the correlation coefficient of pixels far away from the peak. Both horizontal and

vertical directions display a similar rate of IPC increase with signal, a characteristic associated

with decreasing slope in the variance versus mean as signal increases. Both effects are typically

associated with BFE but confirmation awaits spot projection tests in which the dependence of

lateral charge diffusion on signal contrast can be measured directly. It is relevant to note here that

the pixel clock was slowed down to 1.25 kHz to eliminate any inter-channel coupling or bandwidth

limit coming from the video acquisition chain.

The classical variance-versus-mean method [24] assumes pixels are uncorrelated and therefore

26

Page 27: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

0 20000 40000 60000 80000 100000 120000 140000Signal (e-)

0.50

0.25

0.00

0.25

0.50

0.75

1.00

1.25

1.50

IPC

(%)

R^2 = 0.558 Slope (95%) = 4.4e-06 %/e +/- 1.1e-06

Inter Pixel Coupling VS Signal in Low Gain ModeHorizontal IPCVertical IPC

Fig 19: Horizontal and Vertical Correlation of first neighbors in low gain mode.The error reportedon the slope justifies the linear fit versus a constant.

underestimates the shot noise of the incident photons due to the smoothing effect of the spatial

correlation. Figure 20 shows such a Photon Transfer Curve. A similar variance-versus-mean

curve that integrates the power distributed in the wings of the impulse response reveals the true

conversion gain. Before correction for correlations, the conversion gain is overestimated by ∼ 6%.

The corrected Conversion Gain has been used throughout this paper wherever units are in e−.

While inter-pixel coupling can be accurately measured in low-gain mode, no significant cou-

pling was measurable in high-gain mode (Figure. 21).

2000 4000 6000 8000 10000 12000 14000 16000Mean (adu))

0

250

500

750

1000

1250

1500

1750

2000

Var (

adu^

2)

Classic Janesick Method: Slope (95%) = 8.5 +/- 0.1 e /aduIntegrated PSF

i, jPSF(i, j)/size2: Slope (95%) = 7.9 +/- 0.1 e /adu

Photon Transfer Curve using auto-correlation features

Classic Janesick method or PSF(0, 0)/size2

Classic Janesick method or PSF(0, 0)/size2

Using integrated PSF i, j

PSF(i, j)/size2

Using integrated PSF i, j

PSF(i, j)/size2

Fig 20: Photon Transfer curve obtained in Low-Gain mode by taking into account the pixel cor-relation. The total conversion gain drops by 6% from 8.5 e−/ADU to 7.9 e−/ADU in low-gainmode.

27

Page 28: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

0 2000 4000 6000 8000 10000 12000Mean (adu))

0

2000

4000

6000

8000

10000

12000

14000Va

r (ad

u^2)

Classic Janesick Method: Slope (95%) = 1.08 +/- 0.01 e /aduIntegrated PSF

i, jPSF(i, j)/size2: Slope (95%) = 1.07 +/- 0.03 e /adu

Photon Transfer Curve using auto-correlation features

Classic Janesick method or PSF(0, 0)/size2

Classic Janesick method or PSF(0, 0)/size2

Using integrated PSF i, j

PSF(i, j)/size2

Using integrated PSF i, j

PSF(i, j)/size2

Fig 21: Photon Transfer curve obtained in High-Gain mode. The pixel correlation is insignificant.

5.9 Cosmic Ray Statistics

We have examined the distribution of cosmic rays incident on the detector from three 18 hour

dark exposures of the CMOS detector in order to confirm that event rates and number of pixels

affected are as expected. This information is useful for understanding how particle event sizes

(track lengths) are likely to scale if thickness is changed, and to check that the detector package

and AR coatings do not produce additional events. Bad pixels were masked to avoid spurious

detections, but this masking did not include the noisiest pixels.

Cosmic rays were identified as any contiguous region above a threshold set to 50 e-, well

above the median read noise. This resulted in 31,877 events, or 1.45 CR/min/cm2, which is 45%

higher than the expected 1 CR/min/cm2 [25]. The excess might be caused by radiation from our

concrete building. Figure 22 shows four examples of cosmic ray events identified with this method.

We estimate the measured length distribution by taking the diagonal length of the bounding box of

each cosmic ray. This crude method creates a large fractional error on the length of smaller cosmic

rays.

For this experiment, the detector was facing the horizon. The incident angles of incoming

28

Page 29: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

0.0 2.5 5.0 7.5 10.0 12.5 15.0 17.5

0

2

4

0 1 2 3

0

1

2

3

0 1 2 3 4

0

1

2

3

0.0 2.5

0

2

4

6

Fig 22: Example of various length cosmic ray events identified by finding contiguous regionsabove 50 e− in dark frames.

cosmic rays, θ and φ, were then calculated to follow the distribution of Equation 2 that accounts

for the typical angular distribution of muons at the ground according to [25]. The resulting angle

distribution is shown in Figure 23 (Right). A more in-depth analysis would need to account for

radiation from the concrete room.

P (θ, φ) = cos2(θ − π/2)) ∗ cos(θ) ∗ cos(φ) (2)

A diagram of the incident angle in reference to the cosmic ray and detector facing the horizon

is shown in Figure 23 (Left). The length l of each Cosmic Ray track can then be related to the

incident angle θ and φ and the thickness T of the detector.

Using this geometric model, we are able to use the cosmic ray length distribution to infer the

thickness of the detector, previously measured optically to be 7.0 ± 0.3µm. This was done by

29

Page 30: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

z

x

y

ThicknessT

Detector

High incidence muonfrom zenith

θΦ

Low incidence muonfrom horizon

θ

Fig 23: Left Figure: Reference diagram for incident angle θ and azimuthal angle φ.Right Figure: Expected angular distribution of cosmic rays hitting the detector.Muons and cosmic rays are preferentially coming from the zenith while our camera system ispointing towards the horizon. Note muons coming with a high incidence onto the detector leave alonger trace than those coming with a normal incidence.

running a Monte Carlo simulation for the incident angles of cosmic rays, θ and φ, drawn from the

distribution in Figure 23 (Right). We calculated the distribution in length of the simulated cosmic

ray using the angle and thickness of the detector for a range of thicknesses from 5µm to 20µm. The

length distribution agrees with the expectation that shorter events are more likely, as longer cosmic

ray tracks correlate with steeper incident angle. The mean track length is 3.6 pixels. The results

of the Monte Carlo simulations are plotted alongside the measured length distribution shown in

Figure 24 and compared using a least-square method to generate a graph showed in Figure 25,

with the error calculated from a parabolic fit to the residual. The implied detector thickness is

7.3 ± 0.5µm, which overlaps with the direct optical measurement.

We determine from this length distribution that the total rate of pixels affected by cosmic rays is

4.1 pixel/min/cm2 so fractional area contaminated is 4.1 × 10−6 per minute.

Figure 26 shows the charge distribution for all events and just multi-pixel events for compari-

son. We observe excess counts near the minimum charge deposition, suggesting that this is due to

some pixels having sufficient Random Telegraph Noise (RTN) to exceed the CR detection thresh-

old. We assume the multi-pixel event distribution is the true distribution, and will return to masking

30

Page 31: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

0 5 10 15 20Length (pixels)

100

101

102

103

104

105Fr

eque

ncy

Distribution of muon trail lengthMonteCarlo Simulation T = 5.0 umMonteCarlo Simulation T = 10.0 umMonteCarlo Simulation T = 15.0 umMonteCarlo Simulation T = 20.0 umFrom Dataset

Fig 24: Comparison of cosmic ray track length distribution to Monte Carlo simulation. The errormade when evaluating track length on our dataset is shown. To improve quality and smoothness ofthe simulated curves, trial numbers were boost up by 100x in each Monte Carlo simulations.

6 8 10 12 14 16 18 20Detector thickness (um)

6

7

8

9

10

log(

resid

ual)

<- T

hick

ness

= 7

.3um

+/-

0.5u

m

Residual vs thickness

residual of data against modelParabolic fits

Fig 25: The residual is defined as ε =∑ (data−MonteCarlo)2

MonteCarloThe minimum of this residual indicates

the most likely detector’s thickness. The error on this measurement is evaluated from the curvatureof a fitted parabola around the minimum.

pixels with excess RTN in a future work. The modal value of the multi-pixel charge distribution is

175 e−/pixel with estimated width 309 e−/pixel.

98% of cosmic rays are expected to be muons which create linear tracks with known deposition

31

Page 32: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

0 200 400 600 800 1000Charge (e-)

0

50

100

150

200

250Fr

eque

ncy

Charge per Pixel (Single Pixel Events)

0 200 400 600 800 1000Charge (e-)

0

200

400

600

800

Freq

uenc

y

Charge per Pixel (Multi-Pixel Events)

Fig 26: Cosmic Ray Charge Distribution for (left) all events and (right) multi-pixel events. Theexcess counts at lowest charge levels (at left) may be due to RTN and must be investigated further.

rate[26]. To measure this deposition rate, we plotted the total charge deposited versus track length.

This was measured in pixels, and thus is projected onto two dimensions rather than being the true

three dimensional track length for each event. The contribution of device thickness is significant

only for track lengths less than ∼ 100µm. Figure 27 shows an average 66 e−/µm charge deposition

rate in the longest tracks region (> 100µm), which is higher than the 27 e−/µm reported in [26]

for comparably thin silicon, but very similar to deposition rate measured with thick CCDs.

100 200 300 400 500Track Length ( m)

10

0

10

20

30

40

50

Char

ge k

e

66 e / m ± 6e-/ m

Charge vs. Length for Cosmic Ray Events

Fig 27: Charge vs. track length for cosmic ray events. The estimated slope is printed in ke−/µm,with uncertainty estimated from the covariance matrix.

6 Summary

The SRI 4K×2K described here is an “engineering grade” device produced from a wafer that

was part of a production run for the Europa Clipper mission. In spite of a post-processing error

32

Page 33: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

which produced several severe glow sites, dark current and read noise in the unaffected areas are

excellent, rivalling the best CCDs. This is critical for UV spectroscopy and far UV imaging where

sky background is extremely low.

In addition to being much more radiation tolerant than any CCD, this sensor offers higher ex-

posure duty cycle, electronic shuttering, extended Dynamic Range (dual gain, dual exposure), and

the ability to read subarrays (for guiding and/or photometry of very bright stars) while continuing

long exposures on the remaining pixels. UVEX puts all of these features to good use. For example,

several subarrays within each of its nine NUV sensors are located over bright stars which are read

at 10 Hz to provide the precise guiding needed to keep a target centered on its 2 arcsec spectrograph

slit, while also allowing photometry of these bright sources for cross-calibration against shallower

legacy surveys.

The device exhibits excellent linearity. Low correlation between adjacent pixels and the weak

dependence of this correlation on signal bodes well for future measurements of brighter-fatter

effect. The low thickness in relation to pixel size minimizes cosmic ray track length, while the nar-

rowness of the tracks suggests low lateral charge diffusion, though this has not yet been carefully

measured.

While this sensor meets all key requirements for UVEX, further work is desirable to quantify

how far the photometric and astrometric accuracy can be pushed. Future tests will include projec-

tion and scanning of thousands of sub-pixel spots to characterize charge diffusion, to investigate

calibration systematics arising from errors in pixel boundary location (e.g. brighter-fatter effect),

and to verify that photometry does not vary appreciably with sub pixel spot motion. More sensitive

image persistence tests are also planned.

Meanwhile work is proceeding (at JPL) to validate coating performance, optimizing both in-

33

Page 34: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

band QE and out-of-band rejection. An essential part of this work is the validation of Quantum

Yield (the production of more than one electron per photon) and the study of how QY statistics

affect SNR as a function of fluence, a property shared by all silicon detectors.

7 Acknowledgments

We thank the SRI team for packaging multiple detectors and performing screening tests, and for

their advice during our detector characterization effort. We acknowledge the support of JPL and

Caltech’s President and Director’s Research and Development Fund (PDRDF Program) which

supported the majority of this investigation. We also gratefully acknowledge the collaborative

agreement between APL-SRI-JPL that made wafers available for this work.

References

1 C. Martin, T. Barlow, W. Barnhart, et al., “The Galaxy Evolution Explorer,” in Future

EUV/UV and Visible Space Astrophysics Missions and Instrumentation, J. C. Blades and

O. H. W. Siegmund, Eds., 4854, 336 – 350, International Society for Optics and Photonics,

SPIE (2003).

2 O. H. W. Siegmund, J. B. McPhate, T. Curtis, et al., “Development of UV imaging detectors

with atomic layer deposited microchannel plates and cross strip readouts,” in X-Ray, Optical,

and Infrared Detectors for Astronomy IX, A. D. Holland and J. Beletic, Eds., 11454, 274 –

284, International Society for Optics and Photonics, SPIE (2020).

3 S. Nikzad, M. E. Hoenk, F. Greer, et al., “Delta-doped electron-multiplied ccd with absolute

quantum efficiency over 50% in the near to far ultraviolet range for single photon counting

applications,” Appl. Opt. 51, 365–369 (2012).

34

Page 35: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

4 S. Kulkarni, F. Harrison, B. Grefenstette, et al., “Science with the ultraviolet explorer (uvex),”

(2021).

5 E. R. Fossum, “Active pixel sensors: are CCDs dinosaurs?,” in Charge-Coupled Devices and

Solid State Optical Sensors III, M. M. Blouke, Ed., 1900, 2 – 14, International Society for

Optics and Photonics, SPIE (1993).

6 R. Massey, T. Schrabback, O. Cordes, et al., “An improved model of charge transfer ineffi-

ciency and correction algorithm for the hubble space telescope,” Monthly Notices of the Royal

Astronomical Society 439, 887–907 (2014).

7 R. Massey, “Charge transfer inefficiency in the hubble space telescope since servicing mission

4,” Monthly Notices of the Royal Astronomical Society: Letters 409, L109–L113 (2010).

8 J. Janesick, T. Elliott, J. Andrews, et al., “Fundamental performance differences of CMOS

and CCD imagers: part VI,” in Target Diagnostics Physics and Engineering for Inertial Con-

finement Fusion IV, J. A. Koch and G. P. Grim, Eds., 9591, 1 – 17, International Society for

Optics and Photonics, SPIE (2015).

9 S. P. Plunkett, R. Howard, D. H. Chua, et al., “The Wide-Field Imager for the Parker Solar

Probe Mission (WISPR),” in AGU Fall Meeting Abstracts, 2017, SH23D–2693 (2017).

10 R. A. Howard, A. Vourlidas, C. M. Korendyke, et al., “The solar and heliospheric imager

(SoloHI) instrument for the solar orbiter mission,” in Solar Physics and Space Weather In-

strumentation V, S. Fineschi and J. Fennelly, Eds., 8862, 155 – 167, International Society for

Optics and Photonics, SPIE (2013).

11 D. M. Burrows, A. Wolszczan, and A. Moore, The WSPC Handbook of Astronomical Instru-

mentation, vol. 2,3, World Scientific (2021).

35

Page 36: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

12 M. E. Hoenk, P. J. Grunthaner, F. J. Grunthaner, et al., “Growth of a delta-doped silicon

layer by molecular beam epitaxy on a charge-coupled device for reflection-limited ultraviolet

quantum efficiency,” Applied Physics Letters 61(9), 1084–1086 (1992).

13 S. Nikzad, A. D. Jewell, M. E. Hoenk, et al., “High-efficiency uv/optical/nir detectors for

large aperture telescopes and uv explorer missions: development of and field observations

with delta-doped arrays,” Journal of Astronomical Telescopes, Instruments, and Systems 3(3),

036002 (2017).

14 J. R. Janesick, Scientific charge-coupled devices, vol. 83, ch. Delta-doped CCDs. SPIE press

(2001).

15 J. J. H. Shouleh Nikzad, Michael E. Hoenk and L. D. Bell, “High performance silicon and iii-

nitride-based uv and uv/optical imaging detectors,” Journal of Instrumentation , 3–33 (2021).

16 J. Janesick, T. Elliott, J. Andrews, et al., “Mk x Nk gated CMOS imager,” in Target Diagnos-

tics Physics and Engineering for Inertial Confinement Fusion III, P. M. Bell and G. P. Grim,

Eds., 9211, 8 – 20, International Society for Optics and Photonics, SPIE (2014).

17 F. Wang and A. Theuwissen, “Linearity analysis of a cmos image sensor,” in Electronic

Imaging, Electronic Imaging, 84–90, Society for Imaging Science and Technology (2017).

2017 IS&T International Symposium on Electronic Imaging ; Conference date: 31-01-2017

Through 02-02-2017.

18 D. McGrath, S. Tobin, V. Goiffon, et al., “Dark current limiting mechanisms in cmos image

sensors,” Electronic Imaging 2018(11), 354–1–354–8 (2018).

19 C. M. Hirata and A. Choi, “Brighter-fatter effect in near-infrared detectors. i. theory of flat

autocorrelations,” Publications of the Astronomical Society of the Pacific 132, 014501 (2019).

36

Page 37: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

20 A. C. Moore, Z. Ninkov, and W. J. Forrest, “Interpixel capacitance in nondestructive focal

plane arrays,” in Focal Plane Arrays for Space Telescopes, T. J. Grycewicz and C. R. Mc-

Creight, Eds., 5167, 204 – 215, International Society for Optics and Photonics, SPIE (2004).

21 P. Antilogus, P. Astier, P. Doherty, et al., “The brighter-fatter effect and pixel correlations in

ccd sensors,” Journal of Instrumentation 9, C03048–C03048 (2014).

22 A. A. Plazas, C. Shapiro, R. Smith, et al., “Laboratory Measurement of the Brighter-fatter

Effect in an H2RG Infrared Detector,” pasp 130, 065004 (2018).

23 A. Choi and C. M. Hirata, “Brighter-fatter effect in near-infrared detectors. II. autocorrelation

analysis of h4rg-10 flats,” 132, 014502 (2019).

24 J. R. Janesick, Photon Transfer, SPIE (2007).

25 J. Beringer, J. F. Arguin, R. M. Barnett, et al., “Review of particle physics,” Phys. Rev. D 86,

010001 (306–307) (2012).

26 D. Groom, “Cosmic rays and other nonsense in astronomical ccd imagers,” Experimental

Astronomy 14(1), 45–55 (2002).

List of Figures

1 The Mk×Nk Imager reticle consists of four modular building blocks which are

photocomposed (“stitched”) to create custom dimensions in 1k increments. The

sizes shown have been successfully produced, for ground and space missions.

2 Five Transistors with Pinned Photo-Diode Pixel (5TPPD) Diagram. Reproduced

from [16].

37

Page 38: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

3 CMOST Camera. The open cryostat is shown with detector inserted into its zero-

insertion-force socket. The controller sits on top of the cryostat, and the Vacuum

Interface Board provides row and column address generation and routing of video

signals.

4 Readout Diagram of the Rolling Reset Mode. Hypothetical 4 ×4 pixel array. The

true exposure time comprises the readout time that is fixed and a parametric expo-

sure time Tintegration . VCLK is the Vertical Clock pulse that increments the Row

Pointer to the next row.

5 Photon transfer curves of low and high gain modes. Readout noise, system gain

and Full Well Capacity can be directly read from this plot.

6 Initial transient after starting the readout of the detector at a constant (and fastest)

frame rate. Each data point is the average of a clean 128 × 128 window. The

transient is only visible in the illuminated area, implying that this is a gain change

and not an offset.

7 Linearity in high gain mode. Signal is the median of a 256x256 window. Faint

signal ranges from 0 e− to 1000 e−. An F-test comparing a quadratic fit to a linear

fit returns a p-value = 2.2× 10−13, confirming that the quadratic model provides a

statistically improved fit.

8 Linearity in low gain mode. Signal is the median of a 256x256 window. Note

that the faint signal is neglected since low gain mode will not be utilized for very

faint signals. An F-test comparing a quadratic fit to a linear fit returns a p-value =

2.6× 10−12, confirming that the quadratic model provides a statistically improved

fit.

38

Page 39: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

9 Temperature dependence of the gain between 150 K and 180 K. Around 165 K,

the gain varies by 0.15 %/K. The signal is computed as the mean of a 128x128

window.

10 Top Figure: 18 hours exposure in the dark while keeping electronics in its default

state. Bottom Figure: “Low glow” 18 hours exposure in the dark while shutting

down support electronics. Dark Maps of the top left quadrant of the detector show

fine structures and glow sources in and around the image area. Most of the glow

coming from the left edge of the chip, where the logic and pixel drivers are located,

is turned off by lowering the biases during the exposure. Charge collection is

unaffected.

11 Some bright spots show a peculiar semicircular shape with the same orientation.

12 Dark current for a wide range of temperatures. A simple fit of an Arrhenius law

shows the expected temperature dependency.

13 Pixel Speed Noise optimization. The noise is defined as the spatial standard devia-

tion of a 100x100 sub-window of the array. For each pixel frequency, an optimum

number of samples is found by fitting a parabola to the data.

14 Noise Map of the detector in high gain. Excess noise is visible in the bottom rows,

next to the amplifiers. The bad column generates noise in its surrounding as well

as four distinct clusters.

15 Noise distribution of pixels in a clean sub-array of the detector. The ideal gamma

distribution describes distribution if noise were distributed normally in the detector.

The dashed curves represent the fraction of pixels (right y-axis) below a given noise

level (x-axis)

39

Page 40: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

16 Fluctuating level of a pixel affected by Random Telegraph Noise. The amplitude

of the step is much higher than the read out noise. Note that the size of the step is

constant while transition timing is random.

17 Square root of average noise variance for all pixels included in window, after ex-

clusion of noisiest pixels, plotted for three different pixel rates.

18 Left: Auto-correlation of the difference of two flat frames. The noise is reduced

by averaging 48 256 × 256 pixel patches. Right: Impulse response or IPC re-

constructed from the Autocorrelation in low gain mode. In High-Gain mode, no

significant inter pixel coupling is measured.

19 Horizontal and Vertical Correlation of first neighbors in low gain mode.The error

reported on the slope justifies the linear fit versus a constant.

20 Photon Transfer curve obtained in Low-Gain mode by taking into account the

pixel correlation. The total conversion gain drops by 6% from 8.5 e−/ADU to

7.9 e−/ADU in low-gain mode.

21 Photon Transfer curve obtained in High-Gain mode. The pixel correlation is in-

significant.

22 Example of various length cosmic ray events identified by finding contiguous re-

gions above 50 e− in dark frames.

23 Left Figure: Reference diagram for incident angle θ and azimuthal angle φ. Right

Figure: Expected angular distribution of cosmic rays hitting the detector. Muons

and cosmic rays are preferentially coming from the zenith while our camera system

is pointing towards the horizon. Note muons coming with a high incidence onto

the detector leave a longer trace than those coming with a normal incidence.

40

Page 41: arXiv:2112.01691v1 [astro-ph.IM] 3 Dec 2021

24 Comparison of cosmic ray track length distribution to Monte Carlo simulation.

The error made when evaluating track length on our dataset is shown. To improve

quality and smoothness of the simulated curves, trial numbers were boost up by

100x in each Monte Carlo simulations

25 The residual is defined as ε =∑ (data−MonteCarlo)2

MonteCarloThe minimum of this residual

indicates the most likely detector’s thickness. The error on this measurement is

evaluated from the curvature of a fitted parabola around the minimum.

26 Cosmic Ray Charge Distribution for (left) all events and (right) multi-pixel events.

The excess counts at lowest charge levels (at left) may be due to RTN and must be

investigated further.

27 Charge vs. track length for cosmic ray events. The estimated slope is printed in

ke−/µm, with uncertainty estimated from the covariance matrix.

List of Tables

1 Radiation tolerance of SRI Nk×Mk, reproduced from data sheet.

2 Row readout sequence in Rolling Reset mode.Each line is read twice: immediately

after reset to capture the baseline and after charge transfer to capture the signal.

Timing is shown in Figure 4

3 Row readout sequence to generate a dual-gain image. Each row is read four times

to generate two pairs of row per gain. A second charge transfer must be performed

after switching to low gain to recover the full signal range.

41