• Non ci sono risultati.

Superconducting Nanowire Single Photon Detectors based on Magnesium Diboride

N/A
N/A
Protected

Academic year: 2021

Condividi "Superconducting Nanowire Single Photon Detectors based on Magnesium Diboride"

Copied!
93
0
0

Testo completo

(1)

Detectors based on Magnesium Diboride

Simone Frasca

(2)

imply its endorsement by the United States Government or the Jet Propulsion Laboratory, California Institute of Technology.

(3)
(4)

The 2000’s will be remembered as the era of microsatellites. Thanks to

the compactness and lightness of the technology instrumentation, sending

in orbit smaller satellites is going to be the cheapest way to perform space

exploration. However, reducing the dimensions of the satellites is going

to underline even more one of the critical aspects of today’s and future’s

missions: the data transfer. The main constraint is the data rate of these

relatively small satellites. Also, due to planets and moons rotation, the

so-called telecommunication windows for interplanetary missions are pretty

narrow, which make big data transfers hard to perform.

Moreover, with the improvement of technology, the required data rate

will certainly increase, hence finding a way to increment the communication

bandwidth is a primary issue to improve future space mission efficiency.

From these considerations, the idea of optical telecommunication

sys-tems has grown in the last decade, and the Superconducting Nanowire Single

Photon Detectors (SNSPDs [1]–[2]) has proven to be an excellent candidate

to perform the role of the system detector. To date, this technology has yet

(5)

[3].

Optical telecommunication provides some advantages with respect to the

classic RF technology, as higher bit rates and immunity to electromagnetic

interference, i.e. it looks like the ideal technology for deep-space data

trans-mission. If working, an optical telecommunication system would provide 10

to 100 times the state-of-the-art RF system (Ka-band) capability, without

increasing the mission burden in mass, volume, power and/or spectrum.

SNSPDs are prevalently made of niobium nitride (NbN), tungsten

sili-cide (WSi), MoGe, niobium silisili-cide (NbSi), tantalum nitride (TaN).

Cur-rently SNSPDs are state-of-the-art in near IR single photon detection. WSi

SNSPDs have 90% system detection efficiency (SDE), low timing jitter

(50-100 ps), high max count rates (20 Mcps) and have wide spectral sensitivity

(UV to mid-IR).

However, WSi SNSPDs require sub-Kelvin temperatures to operate

op-timally, which means the cryogenics become complex, expensive, large and

resource intensive. To date, no SNSPD made of these materials has been

demonstrated to work above 10K, while NASA has a flight heritage in flying

compact cryocoolers at temperatures greater than 20K.

The purpose of this thesis will be a study about the feasibility of a ”high

critical temperature” material SNSPD, in particular devices made of

mag-nesium diboride (MgB2). Magnesium diboride has a critical temperature of 39K, which makes it one of the best candidates for easy-to-fly optical

(6)

We demonstrated single-photon sensitivity up to 18K for 1.55-µm light, which is the most interesting wavelength for optical telecommunication

pur-poses. However, some unexpected behaviors affect our devices, which will

(7)

This work was developed as my master thesis during my intern at Jet

Propulsion Laboratory, California Institute of Technology. This work was

possible thanks to the help of all the people around me, friends, colleagues

and family. I would like to thank in particular:

My thesis advisor Prof. Alessandro Quarta who has been patient with

me and has supported me during these tough months despite the huge

distance. Big thanks to Dr. Francesco Marsili and Dr. Matt D. Shaw,

who gave me the chance to come at Jet Propulsion Laboratory in the first

place.

My project advisor and mentor Angel E. Velasco, who I had the chance

to work with for almost the whole time at JPL. His continuous explanations

were the main sources for my personal growth during the whole internship

duration. I will never be grateful enough for that.

My team colleagues Emma E. Wollman and Jason P. Allmaras, who

helped me lots of times with their experienced comments and advices. It

was my pleasure to have the chance to work with you all.

(8)

re-member of a day in Pisa without any of the two of you. Of all I accomplished

during these two and a half years, more than just ”a part” of the merit is

yours. Also, a big thank you goes to Francesco, Antonio and Alessandro:

your presence and friendship helped me growing intellectually, and I am

extremely grateful to you all for that.

My family, who supported me every single day of my life, included this

adventure on the other side of the world. Without you I wouldn’t be the

person I am today, so, sincerely, thank you for the job you all did.

But the most special thanks go to my brothers. Federico and Alessandro

are the strongest reason why I realized all I accomplished so far, and for

sure they are the reason why I won’t stop doing my best, to make them

proud of their older brother. I love you.

This research was carried out at the Jet Propulsion Laboratory,

Califor-nia Institute of Technology, and was sponsored by the JPL Visiting Student

Research Program (JVRSP) and the National Aeronautics and Space

(9)
(10)

Abstract I

Acknowledgments III

1 Superconducting Nanowire Single Photon Detectors:

The-oretical Overview 1

1.1 The Hotspot Mechanics . . . 2

1.2 Optical Characterization . . . 4

1.2.1 Statistical Approach of Detection Events . . . 6

1.3 Electrical Characterization . . . 8

1.3.1 The Latching Effect . . . 13

1.4 Uncertainties Characterization . . . 15

1.4.1 The System Dark Count Rate . . . 16

1.4.2 The Afterpulsing Effect . . . 17

2 Devices Fabrication Process 20 2.1 Depositions Techniques . . . 21

2.1.1 MBE . . . 21

2.1.2 HPCVD . . . 23

2.2 Devices Writing Process . . . 25

3 Testing Process 28 3.1 Set-up Description . . . 29

(11)

3.3 The R vs T characteristics . . . 39

3.4 Photo-sensitivity Tests . . . 41

3.4.1 Single- or Multi-Photon Sensitivity Test . . . 43

3.4.2 Device Detection Efficiency and Afterpulsing . . . 44

4 Best Devices Results 50 4.1 v40 S4-3H . . . 52 4.2 v42 2B . . . 58 4.3 v43 5G . . . 62 4.4 Last Devices . . . 67 5 Conclusions 69 Bibliography I

(12)

1.1 The hotspot generation sequence . . . 3

1.2 Example of SNSPDs geometries . . . 5

1.3 Example of Single- and Two-photons detection events . . . . 7

1.4 DDE vs Ibias for a device capable of reaching QE saturation [3] 8 1.5 Pulse’s Rise and Fall times . . . 9

1.6 Example of paralysis on a scope trace . . . 15

1.7 Difference between a not-afterpulsing and an afterpulsing be-havior . . . 17

2.1 AFM image of as-grown (a) and rapidly-annealed (b) films [39] . . . 22

2.2 Surface morphology of films grown on different substrates . . 23

2.3 Image of surface morphology grown at 70 (A), 60 (B), 50 (C) and 40 (D) torr [37] . . . 24

2.4 Surface morphology of films grown at different thicknesses . 24 2.5 Fabrication process . . . 26

2.6 Examples of devices’ geometries . . . 27

2.7 Example of devices’ geometries with the addition of ”dummy wires” . . . 27

3.1 Device to SMP interface . . . 29

3.2 Equivalent electrical circuit of our set-up . . . 31

3.3 I –V curve . . . 33

(13)

3.5 I –V curve measured first increasing and then decreasing the

bias . . . 36

3.6 I –V curve with hotspot plateau . . . 37

3.7 Set-up scheme of the Fast–Isw measurement [44] . . . 38

3.8 Expected scope readout of the Fast–Isw measurement [44] . . 38

3.9 Fast–Isw measurement results . . . 39

3.10 Classic R vs T characteristic . . . 41

3.11 Classic shape of a Pulse . . . 42

3.12 Difference in the photo-sensitivity curves for a single- and a three-photons sensitive detectors. . . 45

3.13 DDE vs Ibias/Isw curve for a single-photon sensitive detector at different temperatures . . . 46

3.14 Classic DDE vs T curve . . . 47

3.15 Afterpulsing usual scope trace . . . 48

3.16 Inter-arrival time of an afterpulsing device . . . 49

3.17 Inter-arrival time of a not-afterpulsing device . . . 49

4.1 Design image of the Mask used for fabrication. . . 51

4.2 v40 S4-3H . . . 53

4.3 I –V characteristic of v40 S4-3H at different temperatures . . 53

4.4 R vs T characteristic of v40 S4-3H . . . 54

4.5 Single-photon sensitivity test at 1.55-µm wavelength for v40 S4-3H, T = 18K . . . 54

4.6 Pulse track of v40 S4-3H at three different temperatures . . 55

4.7 Inter-arrival time measurements of v40 S4-3H at three test temperatures . . . 55

4.8 DDE vs T curve of v40 S4-3H at Ibias/Isw = 99% and 99.5% 56 4.9 Pulses’ Peak to Peak amplitude vs T curve of v40 S4-3H . . 57

4.10 SDCR vs Ibias curve of v40 S4-3H at multiple temperatures . 58 4.11 v42 2B . . . 58

(14)

4.13 DDE vs T curve of v42 2B . . . 60 4.14 Rpolariz vs T measurement of v42 2B . . . 61

4.15 Inter-arrival time measurement of v42 2B at T = 9K . . . . 61 4.16 v43 5G . . . 62 4.17 I –V characteristic of v43 5G at different temperatures . . . 63 4.18 R vs T characteristic of v43 5G . . . 64 4.19 DDE vs T curve of v43 5G . . . 64 4.20 Persistance traces of v43 5G pulse at different temperatures 65 4.21 DDE vs T curve of v43 5G, making the distinction on the

trigger level . . . 66 4.22 Pulses’ Peak to Peak amplitude vs T curve of v43 5G . . . . 67 4.23 Pulse in superconducting state (a) and in slightly resistive

(15)

Superconducting Nanowire

Single Photon Detectors:

Theoretical Overview

Superconducting materials may be the future for space applications. Their

property to present no resistance and very low thermal conductivity below

the so-called critical conditions opened them to plenty of new technology

developments.

These critical conditions exists under the shape of two working-condition

constraints: the first one is the critical temperature Tc, above which the

ma-terial loses all its superconducting properties, while the other is the critical

magnetic field Bc, which behaves the same way. They are related to each

other through the equation:

Bc ≈ Bc(0)  1 − T Tc 2 (1.1)

(16)

wire through the Ampre’s law, we will, for the sake of simplicity, refer to

this property under the form of critical current Ic.

As will be explained later, this parameter is just theoretical. In fact,

for our devices, we will refer to the switching current Isw, which is the

experimental value at which the device switches to the normal state.

1.1

The Hotspot Mechanics

Superconducting Nanowire Single Photon Detectors were first developed by

G. N. Goltsman et al. at Moscow State Pedagogical University in

collabo-ration with C. Williams and R. Sobolewski of the University of Rochester

in 2001 [1].

As they demonstrated, for a superconducting nanowire at operating

tem-perature T lower than its critical temtem-perature Tc, biased at sub-critical

cur-rent Ibias, nominally a single photon, carrying an energy such that ~ω  2∆,

where 2∆ is the superconducting energy gap, impinging on the device, could

generate a perturbation strong enough to break the Cooper’s Pairs of the

superconducting state of the matter due to electron–electron and electron–

Debye–phonon interactions, generating a high temperature electron zone

which, exceeding Tc, creates the so-called hotspot [6]. A scheme of the

hotspot mechanics is shown in Fig. 1.1.

Due to the exceeding of the critical temperature, this part of the nanowire

(17)

(a) The nanowire is biased at subcriti-cal current Ibias

(b) A photon carrying enough energy impinge on the nanowire

(c) The Cooper’s Pairs broke and a re-sistive state portion of the nanowire appears

(d) The current is redirect to the read-out circuit. The nanowire than cools down and returns to the su-perconducting state where the cur-rent starts flowing again trough it (a)

Figure 1.1: The hotspot generation sequence

state. Hence, even if no general theory has been accepted so far [7], the

hotspot starts growing, creating the hotspot region. The growth happens

due to the Joule heating produced by the normal state region.

After the switch, the current is redirected in the readout circuit, which

usually has a resistance Rreadout = 50 Ω, since in theory the resistance

of the hotspot region is much higher than that, usually in the order of

MΩs. The current flowing through the nanowire hence decreases until the

nanowire cools down and restores its superconducting properties, allowing

the current to flow again. This double transition can be seen from the point

of view of the electrical circuit as a fast change of resistance from the zero

value (superconducting state) to the resistive one (which, in case of MgB2 nanowires, is on the order of kΩ). For that reason, when part of the current

(18)

is redirected to the source, generates a voltage pulse reading.

1.2

Optical Characterization

In order to characterize the quality of a SNSPD, some important parameters

need to be now introduced: first of all, the Device Detection Efficiency

(DDE), i.e. the ratio of the number of detected photons and the number that

impinge on it; the System Dark Count Rate (SDCR), which is the number

of recorded counts when the fiber and optics are connected to the SNSPD

and the laser is off, i.e. a sort of parameter which describes the robustness of

the system to external radiation and bias fluctuations; the Maximum Count

Rate (MCR), which depends on the minimum time interval at which each

pulse can be separated; the dead time τdead, which is the time window after

a detection event in which the device is unable to detect other photons;

the timing jitter (∆t), which is the deviation from the true periodicity of a

presumably periodic signal; the spectral sensitivity of the device.

The DDE is defined as the product of the optical absorption coefficient

(η) and the quantum efficiency (QE) of the device. For high-quality

de-vices, this value needs to be optimized by increasing the two parameters

that compose it. The absorption coefficient is an optical property, which

can be enhanced by increasing the thickness of the device or

implement-ing the so-called optical cavities [9]–[11]. The quantum efficiency is instead

(19)

de-(a) Bridge Structure (b) Meander Structure

Figure 1.2: Example of SNSPDs geometries

pendent, and can be enhanced only by fabricating nanowires with smaller

cross-sections.

The SNSPD has to be thought as a ultrathin (5–10 nm) and ultranarrow

(30–200 nm) nanowire, whose length can vary from several hundreds of

nanometers, while in ”bridge” configuration, to the order of magnitude of

millimeters, while in ”meander” configuration. In Fig. 1.2 it is possible to

see the two main geometric styles used during the study.

The nanowire is biased at a sub-critical current, while enlightened by a

photon source (usually a laser). The calibration of the light source has to be

very accurate, since a mis-evaluation of the actual number of photons

strik-ing the nanowire can end up in a wrong evaluation of the device detection

efficiency. The way it was done during the experiments will be explained in

the next chapter.

Let us now consider a light source pointing the device. We will define

Jin as the fluence of photon striking the device, which is the amount of

(20)

define Jabs = JinSdη, which is the fluence of photon which can be detected

by the device, where Sd is the active area of the device, η is the radiation

absorption coefficient of the metallic film given by:

η ≈ 4(Rs/Z0)/[(Rs/Z0)(nsub+ 1) + 1]2, (1.2)

nsub is the index of refraction of the SPD substrate (which presence will be

discussed later in the manufacturing chapter), Rs is the surface resistance

of the nanowire above Tc and Z0 = 377 Ω is the free-space impedance.

If the quantum efficiency reached 100% by fabricating perfect and

ultra-narrow nanowires, the absorption coefficient would be the physical limit of

the device detection efficiency. The goal of all the proven materials SNSPDs

is to find a way to fabricate devices able to achieve 100% quantum efficiency,

and then to design optical stacks (resonant cavities) in order to enhance the

absorption coefficient.

1.2.1

Statistical Approach of Detection Events

The next step is to define a constitutive law for the detection to happen.

This is necessary in order to define the properties of a ”single–photon”

operating condition device. In order to be single-photon sensitive, the

de-vice needs to have a low enough superconducting energy gap ∆(T ) and a

high enough bias current such that just one photon is able to perturb the

(21)

(a) Single-photon impinging on the SNSPD

(b) Two-photons impinging on the SNSPD

(c) Single-photon detection event (d) Two-photons detection event

Figure 1.3: Example of Single- and Two-photons detection events

due to the hotspot generation. When it does not happen, we can appreciate

two-, three- or multiple- photons detection mechanics (see Fig. 1.3).

The dependence on the bias current can be seen from the QE point of

view (see Fig. 1.4). For low bias, the quantum efficiency is low and the

energy of one photon is not enough to cause a detection event to occur.

However, more photons striking the same portion of the device in the same

time may be able to cause a detection event, allowing multi-photons

sensi-tivity. Increasing bias current, the QE increases and a single photon may be

able to carry enough energy to produce the detection event, allowing then

single-photon sensitivity. Then further increasing the current leads to the

QE saturation plateau, where the QE has reached its maximum value.

From a simplified statistical approach, the detection can be considered

as a Poisson distribution [13]–[14].

(22)

Figure 1.4: DDE vs Ibias for a device capable of reaching QE saturation [3]

then the probability for the device to detect n photons is given by:

P (n) ≈ e

−mmn

n! (1.3)

from which, considering that e−m is fully dependent on the photon power input, if m  1 , follows that P (n) ∼ mn/n! and then that the

probabil-ity to detect photons is proportional to m for a single–photon absorption

regime, to m2 for a two–photons absorption regime, and so on.

1.3

Electrical Characterization

After the absorption of a photon, the device switches, acquiring a resistance

value Rn [13]. What happens is that the current in the device then starts

decaying from its initial value to the normal state current

In= Ibias·

Rreadout

(Rreadout+ Rn)

(23)

Figure 1.5: Pulse’s Rise and Fall times

where Rreadout is the resistance of the readout circuit and Ibiasis the biasing

current. This decay happens with a time constant τrise= L/(Rreadout+ Rn),

where L is the inductance of the device. It is called rise time since in the

readout circuit the current increases from zero to its maximum value due

to the redirection of the current into it caused by the switch of the device

(see Fig. 1.5).

The current drop in the device is then stopped at a certain current value

Iret, at which the self-heating of the device due to Joule heating is reduced

enough to allow the wire to recover the superconducting state. At the same

time, the readout circuit sees the peak of the pulse, which has an amplitude

of:

Vpulse = Rreadout· Gampl· (Ibias− Iret) (1.4)

(24)

to its initial value (Ibias) with time constant τfall = L/Rreadout.

From the equations obtained above, it is possible to notice the main role

of the inductance L in the time response of these devices. In fact, in order

to manufacture ultrafast devices, its value should be minimized.

It has been proven [13] that the inductance of the nanowire is prevalently

dominated by the kinetic one, Lk. In order to appreciate the magnitude

difference between the geometric and the kinetic inductance, consider that

a 5 nm thick, 100 nm wide and 5 µm long bridge made of NbN has a geometric inductance Lgeom = 0.005 nH, and, in accordance with the results

obtained by Kerman et al. [13], biased at 11.4 µA it develops a kinetic inductance Lk = 6.10 nH. For that reason, from now on, we will consider

only the contribution of the kinetic inductance.

The kinetic inductance can be considered as the inertia of the charge

carriers, which becomes relevant when the direction of current flow

sud-denly reverses (i.e. during the switch). Its value is proportional to the

room-temperature resistance of the nanowire, since both have the same

de-pendence upon the wire geometry:

Lk = Lk Z wire ds/Acs(s) (1.5) R300K = R300K Z wire ds/Acs(s) (1.6)

where Lk is the kinetic inductivity and R300K is the room temperature

(25)

is the product of the width and the thickness.

An analytical expression for Lkcan be obtained from the one-dimensional

Ginzberg-Landau theory [15]:

Lk = µ0λ2· l/Acs ⇒ Lk = µ0λ2, (1.7)

where µ0 = 4π × 10−7 H m−1 is the magnetic permeability of free-space, l is

the length of the nanowire and λ is the magnetic penetration depth. This

is valid in the approximation of T near Tc and Ibias ≈ 0.

For a temperature independent calculation, we can use the BCS (Bardeen,

Cooper, Schrieffer) theory and obtain a more complex form of the kinetic

inductance [16]: Lk =  l w  Rsqh 2π2 1 tanh2k∆ BT ⇒ Lk = ρh 2π2 1 tanh2k∆ BT (1.8)

where Rsq= ρ/t is the ratio between the resistivity ρ and the thickness t of

the nanowire, h = 6.62607004×10−34m2kg s−1is the Planck’s constant, ∆ is

the superconducting energy gap and kB = 1.38064852 × 10−23m2kg s−2K−1

is the Boltzmann’s constant.

From all this, manufacturing shorter or thicker devices having the same

width implies reducing the kinetic inductance Lk, allowing ultrafast photon

detection. However, decreasing the kinetic inductance is limited by the

latching effect, as will be explained later.

(26)

unusual shape with respect to other technology fields: instead of speaking

of resistance and geometry, usually the devices are described through the

”sheet resistance” and the ”number of squares”. That is mainly because,

in this way, the characteristics of the film are easily recognized (the sheet

resistance can also be considered as a parameter for the thickness of the

original film).

In a regular conductor, the resistance can be written as R = ρl/Acs =

ρl/w · t. Combining resistivity and thickness, obtain:

R = ρ t · l w = Rsq· l w (1.9)

where Rsqis the so-called sheet resistance, and, as shown, takes into account

both the resistive properties of the material and the thickness of the device.

Defining now the number of squares will explain why such a choice. For

a device, the number of squares Nsq is given by:

Nsq =

Z

wire

ds/w(s) (1.10)

which in case of constant width device geometry becomes Nsq= l/w.

Mul-tiplying the number of squares by the sheet resistance gives the overall

resistance. Moreover, more squares also means higher kinetic inductance of

the device, hence slower devices, since the kinetic inductance is proportional

(27)

1.3.1

The Latching Effect

Despite the capability of these devices to reach, nominally, ultrafast

de-tection rates, an important role is played by the so-called latching. As we

explained before, the timing capability of the device is strictly related to

the device’s reset time constant τreset = Lk/R, where Lk is the kinetic

in-ductance and R is the impedance across it.

In order to make the devices ultrafast, it is necessary to reduce as much

as possible the value of τreset. It can be done in two ways: reducing Lk or

increasing R.

The first choice is more disadvantageous then else, since in order to

re-duce Lk we need to design shorter devices, hence meanders with smaller

active area or lower fill-factor. Anyway, even if it is feasible, this approach

leads to a very fast device, but on the other hand also to a very low

de-tection efficiency due to the small active area. Moreover, since the kinetic

inductance is also proportional to the resistance, an decrement of Lk means

also a reduction of R, which does not seem to be the optimal way to develop

these devices.

For that reason the choice made from Yang et al. [19] was to increase

the resistance. In order to do that, they added an on-chip Ti resistor (a

shunt resistor Rshunt [20]) in series with the SNSPD, increasing the total

load’s resistance. Using a shunt resistance of Rshunt = 192 Ω, they were able

to decrease the time constant τreset of a factor 5, leading to the speed-up of

(28)

In theory, using a larger resistor should lead to an even faster device,

but a consideration has to be made in that sense. In fact, increasing the

series resistance means also a decrement in the value of current redirected

after the switch, since the normal state current is given by:

In= Ibias·

Rreadout+ Rshunt

Rn+ Rreadout+ Rshunt

(1.11)

If In > Iret, then the device will not go back to its superconducting state

unless the current is brought back to the Iret value. If, on the other hand,

In < Iret, latching may occur. That happens when the heat generated by

Joule heating equals the heat dissipated into the substrate and along the

wire. That is why the limit imposed by the latching effect is that the reset

time constant of the SNSPD cannot overcome the thermal time constant,

the electron–phonon interaction time τe–ph, of the material. If this condition

is not satisfied, for the latching effect will follow the generation of a

self-sustained normal-state hotspot.

This self-sustained hotspot can be unstable, under the shape of the

so-called paralysis, in which case the hotspot will be cooled down after a while

(see Fig. 1.6), or it can be stable. In the second case, the device will switch

and remain in normal state until the current is then brought back to lower

(29)

Figure 1.6: Example of paralysis on a scope trace

1.4

Uncertainties Characterization

The devices, however, can not be considered as if they were perfectly

manu-factured, since every production process has its limits, which implies defects.

It is then necessary now to introduce the concept of the constrictions [17].

These manufacturing defects were in the beginning considered only as

geo-metrical issues, like a portion of the wire which has a different cross-section

or due to crystal boundaries. More recent studies [18] confirm that the

con-cept of these imperfections is wider than that, since a constriction can also

be a quantum deficit, like a portion of the material which does not behave

as a perfect superconductor, or a chemical defect.

Due to this effects, the actual critical properties of the devices usually

shift from the theoretical values. For example, let us consider a certain

(30)

Ic given by the relation:

Ic(T ) = Jc(T ) · Acs (1.12)

where Jc is the critical current density at a specific temperature. However

in the reality, due to the presence of constrictions, the actual (observed)

critical current of the device appears to be:

Icobs = Ic· C (1.13)

where C ≤ 1 is the so-called constriction factor. This means that the actual

critical current density is not Jc but instead a lower value Jcobs = Jc· C.

From now on, we will refer to the observed critical current as the switching

current, Isw.

1.4.1

The System Dark Count Rate

Another critical aspect of the SNSPDs is the System Dark Count Rate

(SDCR). This is the combination of two effects: the background photons,

which make the device trigger even when the laser is off, and the intrinsic

dark count rate. The origin of the latter phenomenon is not very clear to

date, but it is mainly addressed to two different theories: the theory of

the vortex-antivortex pair (VAP) unbinding due to

Berezinskii-Kosterlitz-Thouless (BKT) transition, and the theory of the vortex hopping from an

(31)

(a) Pulses train while not afterpulsing

(b) Pulses train in afterpulsing case

Figure 1.7: Difference between a not-afterpulsing and an afterpulsing be-havior

The system dark count rate is a parameter of major importance, since

depending on its value the signal to noise ratio (SNR) is altered.

1.4.2

The Afterpulsing Effect

Finally, it is necessary to introduce the concept of afterpulsing. This

phe-nomenon appears under the form of an extra triggering, i.e. one incoming

photon generates more than one pulse. Usually, the extra pulse shows up at

a constant time interval after the ”main” pulse (depending upon the device,

[28]).

It is not clear what is the origin of this phenomenon, since it may be

gen-erated by the device itself due to some mysterious effect or by the coupling

(32)

Fujiwara et al. [25] found out that the presence of afterpulsing was

re-lated to the reflection of the pulse inside the system electrical feed line. In

order to assess that, they analyzed the increment of counts adjusting the

background light, but since the increment was not proportional at the

differ-ent bias currdiffer-ents tested, they realized that it should not be a photosensitive

effect.

Burenkov et al. [28], on the other hand, studied the distribution of

de-tect events of their devices. The system dark count rate pulses should be,

in principle, equally spread in the time domain. From their analyses,

how-ever, they noticed an uneven distribution of dark counts, which according

to them was related to afterpulsing. They found the presence of trains of

pulses, equally time-spaced, of 2, 3, 4 or more pulses. After some

consider-ations, they changed the amplifiers suit to a more low-frequency sensitive

one, and what they found out was no more afterpulsing. Moreover, they

no-ticed that, unexpectedly, after the main pulse the detection efficiency of the

device does not increase monotonically, but presents an overshoot. Hence,

they speculate that afterpulsing, dark counts and detection efficiency of the

device are three correlated effects which arise due to a perturbation in the

bias current through the nanowire.

What has been found so far is, anyway, that afterpulsing is not a

pho-tosensitive effect, i.e. it appears also while the device is not enlightened by

a laser source and all the pulses are generated by the system dark counts,

(33)

current.

Finally, after explaining all the physics that rules these devices, in Tab.

1.1 we collect the most characterizing parameters of the several

supercon-ducting materials used for SNSPDs.

Material, formula Niobium Nitride, NbN Tungsten Silicide, WxSi1-x Magnesium Diboride, MgB2 Bulk Critical Temperature, Tc [K] 16 4.9 39 Electron– Phonon interaction time, τe–ph [ps] ∼5 ∼35 ∼2 Sheet Resistance, Rsq [Ω/sq] (thickness) 900 (5 nm) 380 (5 nm) 80 (7 nm, HPCVD) 150 (10 nm, MBE) RMS Roughness [nm] .35 ? ∼2 (HPCVD) 0.35 (MBE) Atomic Structure

Crystalline Amorphous Crystalline

Table 1.1: SNSPD’s materials characteristic properties

As shown by the table, not only the bulk temperature of the magnesium

diboride is the highest, but also the electron–phonon interaction time τe–ph,

which sets the speed limit of the devices, is lower both than the WSi and

NbN ones, which means that if the devices work properly, they will be a lot

(34)

Devices Fabrication Process

Our fabrication process consists in obtaining the nanowires starting from

MgB2ultrathin films. Since the discovery of superconductivity in MgB2[31], different techniques have been developed and optimized in order to obtain

high-quality thin films, like pulsed laser deposition (PLD), molecular beam

epitaxy (MBE) [32], hybrid physical-chemical vapour deposition (HPCVD)

[33] and reactive evaporation (RCE) [34].

The different growth processes of MgB2 impart different superconduct-ing characteristics for the films. For example, films grown with MBE are

much smoother than the one obtained by HPCVD, which is essential for the

proper fabrication of homogeneous nanostructures, but they show a strong

suppression of the critical temperature with respect to the bulk temperature

(Tc = 20 − 24K for thin films, i.e. a suppression of 39% with respect to bulk

Tc in the best case) even for relatively thick (∼200 nm) films. However,

this issue for MBE has been partially solved by Shibata [39] developing

(35)

30K-33K (i.e. a suppression of ”only” 15% in the best case) for thin films

(∼15 nm).

Moeckly and Ruby developed the reactive evaporation (RCE) method,

which seems to be the most promising for smooth, high critical temperature

thin films. However, to date no ultrathin film has been demonstrated using

RCE technique.

2.1

Depositions Techniques

Our devices were produced by two different types of films: the deposition of

films by MBE developed by Prof Shibata, who deposited the films at NTT

Basic Research Laboratories in Kanagawa, Japan, and the films obtained

by HPCVD developed by Prof Xi’s group from the Department of Physics

of Temple University in Philadelphia, Pennsylvania. Here we will discuss

briefly the fabrication processes of these films.

2.1.1

MBE

Molecular beam epitaxy growth takes place in a vacuum or ultra-high

vac-uum (UHV) chamber (10−8− 10−12torr). Mg and B are coevaporated using

e-guns with a film growth rate of 0.25 nm s−1 at 280◦C. After the deposition, the wafer is passivated with a SiO2coating, which thickness is around 20 nm,

and then annealed with the fast-annealing process described in [39]. The

(36)

(a) AFM of as-grown film. The RMS roughness of the films made in this way widths between 1.2 and 0.4 nm

(b) AFM of rapidly-annealed film. The RMS roughness of the films made in this way goes down to 0.35 nm

Figure 2.1: AFM image of as-grown (a) and rapidly-annealed (b) films [39]

This is because film grown with MBE process are magnesium deficient,

which affects the Tc. Also, and the thinner is the film, the more difficult

is to retain the magnesium, which is why the critical temperature degrades

even further.

However, despite also the critical current density was improved, the

resis-tivity of the films was higher than the one obtained with the other processes.

The root mean square roughness (RMS), however, after the annealing,

was lowered even further. For typical so-grown MBE deposition, the RMS

is usually around 1.2 to 0.4 nm, while after the rapid-annealing process this

value gos down up to 0.35 nm (see Fig. 2.1), which means very smooth

sur-face, of great importance for what concerns the fabrication of homogeneous

nanostructures.

Despite it has been demonstrated by Jo et al. [32] that deposition at

(37)

fur-Figure 2.2: Image of surface morphology grown on 6H-SiC (A and D), MgO (B and E) and stainless steel (C and F). For each substrate, two films grown with different deposition flow rates (upper image is lower deposition rate) [37]

ther the temperature gives even smaller grains, despite growth temperature

is only one of the factors that could affect the Tc. However, it seems that

RCE technique developed by Moeckly and Ruby (deposition at 500◦C) may be the most promising one for MBE-type of deposition.

2.1.2

HPCVD

Hybrid physical-chemical vapor deposition, on the other hand, has proven to

be the most effective technique so far [35]. In this technique, Mg is provided

by thermally evaporating bulk Mg pieces (physical vapor deposition), while

B is obtained from the diborane gas, B2H6 (chemical vapor deposition).

It has been proved [37] that the use of different substrates and the

depo-sition rate, as well as depodepo-sition temperature, diborane mixtures, substrate

(38)

Figure 2.3: Image of surface morphology grown at 70 (A), 60 (B), 50 (C) and 40 (D) torr [37]

Figure 2.4: AFM images of three film thicknesses of 7.5 (A), 10 (B) and 25 (C) nm [37]. At lower thickness the film grows by initial nucleation of discrete islands, then they start coalesce. The thicker is the film, the more homogeneous is the surface.

As general rule, a lower deposition rate implies smaller crystals (see Fig.

2.2). This rate, however, is dependent by the total environment pressure

during the deposition and by the flow rate of diborane. The lower the

pressure and the flow rate, the lower the deposition rate. In Fig. 2.3 it is

possible to notice how the environment pressure affects the morphology of

the surface of the films.

Finally, it has been shown that also the thickness of the film plays an

(39)

film, the higher will be the RMS roughness of the film. That is consistent

with the Volmer-Weber growth mode. As can be seen by Fig. 2.4, the

MgB2 film grows by initial nucleation of discrete islands, which coalesce with increasing thickness. However, in order to realize high-Tc ultrathin

films, the deposition has been made to grow thick films, which were then

etched to sub-10 nm films, retaining the critical temperature of 40K.

For the films we processed, the RMS roughness was around 2 nm.

How-ever, after Ar ion milling etch, the RMS rose to 2–4 nm, which, for 7 nm

thick nanowires, implies a very inhomogeneous nanostructure. However, the

critical temperature usually exceeded 39K, the bulk critical temperature of

MgB2, due to the strains in the film.

2.2

Devices Writing Process

The devices fabrication process started with a 50 nm thick MgB2 film grown on a SiC substrate and passivated with a thin gold layer (∼30 nm). Via

Ar-ion milling [40], the film was thinned down to 7 nm and then passivated

with 30 nm thick SiO2.

Then, via lift-off procedure, the gold contact pads were put in order

to allow the device connection through the passivation layer. A layer of

resist was deposited on the stack and then, using electron-beam lithography

(EBL), the nanowire shape was written. After that, using Fluorine ICP the

(40)

Figure 2.5: Fabrication process

film uncovered in the spacing between the nanowire shape.

Finally, using a Ar ion mill etching process the MgB2 film was written and then passivated with SiO2in order to avoid impurities and deterioration

of the device.

Using this method, we were able to design different geometries, from the

relatively easier bridges to the more complicated meanders structures (see

(41)

mean-(a) 100 nm wide 34 µm long Bridge structure

(b) 100 nm wide 10 × 10µm2 Meander

structure

Figure 2.6: Examples of devices’ geometries

(a) 100 nm wide 34 µm long Bridge structure

(b) 100 nm wide 2 × 2µm2Meander

struc-ture

Figure 2.7: Example of devices’ geometries with the addition of ”dummy wires”

der and bridge structures gave pretty different results. In fact, etching the

meanders used to give much better homogeneity than etching the bridges.

For that reason we decided to add ”dummy wires” around the bridges

ge-ometries, in order to recall as much as possible the meanders. Since the

quality of the bridges improved, we chose to add the dummy wires also to

(42)

Testing Process

After the theoretical and manufacturing process description of the SNSPDs,

let us now focus the attention on the testing process developed to

charac-terize the devices.

As will be shown in the following pages, the identification of the devices

is mainly related to a couple of ”characterizing curves”, which are the I –V

and the R vs T curves. Starting from these two graphics, some of the most

important properties of the devices can be extracted.

After this, the photo-sensitivity analysis is developed, sweeping the bias

of the device in a range close to its switching current Isw both shining light

on it (1.55-µm and 635-nm laser) and without the light (i.e. evaluating the SDCR).

If the device is photosensitive, then further analyses are developed on

it, checking the presence of afterpulsing, scanning the pulse looking for

eventual latching and evaluating the detection efficiency with respect to the

(43)

Figure 3.1: Sample bonded to the device–SMP interface PCBs. The sub-strate chip (10 × 10 mm2) is in contact with the copper holder

with grease and GE varnish. From the devices’ pads, the SNSPDs are connected with the PCB via bonding wires, where from the PCB pads the signal is transfered to SMP connectors.

3.1

Set-up Description

The device was cooled down to a base temperature of 3.15K inside a

Cry-omech 410 pulse tube cryostat. It was mounted on a oxygen-free high

ther-mal conductivity (OFHC) copper support and in contact with some grease

between the interfaces to allow a better thermal contact. GE varnish was

applied on the edges of the 10 × 10 mm2 substrate to firmly anchor it to the copper stage.

On the support, there were mounted two PCBs to allow device–SMP

(44)

be tested (see Fig. 3.1). For the optical coupling, a ThorLabs SMF 28e+

single mode fiber was connected to a ThorLabs ED1-S20-MD (20◦ Square Engineered Diffuser) with the purpose to spread the light on a 20 × 20 mm2

spot size centering the wafer.

Connected to the support there was a thermometer, connected to a

LakeShore 336 Temperature Controller, since the purpose of the project

is to develop an high temperature working device it was important to

prop-erly know the actual operating temperature of the device while testing it.

The electrical connection, from the device, came out of the ICE through

high frequency, coax cables with SMA connectors and was connected to a

Mini Circuit ZFBT-4R2GW+ bias-tee, which has a bandwidth of 0.1–4200

MHz. From the bias-tee, then, the DC termination was connected through

a Stanford Research System SIM900 mainframe to the SIM928 Isolated

Voltage Source with a SIM970 Quad Channel Voltmeter in parallel, while

the AC termination was connected to a suit of room-temperature amplifiers,

in order to increase the signal to noise ratio (SNR) of the device’s pulse.

The amplifiers suit changed depending on the height and sharpness of the

pulses, but we adopted mainly a combination of two of the ones in Tab. 3.1.

Then the feed line was connected to a 225 MHz Agilent 53131A Universal

Frequency Counter or to a 6 GHz LeCroy WaveMaster 8600 oscilloscope,

from which different measurements were carried out. A scheme of the set-up

is shown in Fig. 3.2

(45)

Amplifier Name Bandwidth [MHz] Gain [dB] Noise Figure [dB] Impedance [Ω] ZX60-100VH+ 0.3–100 36 4.0 50 ZFL-500LN+ 0.1–500 28 2.9 50 ZFL-1000LN+ 0.1–1000 24 2.9 50

Table 3.1: Amplifiers adopted to test the devices

Figure 3.2: Equivalent electrical circuit of our set-up

to test our devices, the 635-nm was obtained from a ThorLabs S1FC635

Benchtop Laser Source and the 1.55-µm from a General Photonics TLS-101 Tunable Light Source. From the source, the fiber was first connected in

series to a ThorLabs FPC561 Manual Polarization Controller, an attenuator,

and an optical switch. From the switch, one termination was connected

to a fiber case and then inside the ICE, while the other termination was

connected either to the ThorLabs PM100D (1.55-µm) or the Newport 2931-C (635-nm) optical power meter.

(46)

3.2

The I –V characteristics

The current–voltage characteristic (I –V curves) is, as explained before, one

of the ways to characterize the devices.

In order to carry out the I –V measurement, the amplifiers were usually

dismounted since their presence could, in principle, induce noise which could

suppress the switching current in the devices. The reason why we took such

a precaution will be discussed later in the chapter.

Starting from zero, the voltage source (in series with a bias resistor Rb)

increases the bias across the device by fixed ∆V steps, i.e. increases the

bias current Ibias passing through the device by ∆I = ∆V /Rb. At the same

time, the voltmeter evaluates the voltage across the device, measuring the

resistance of the nanowire at each step. There are some uncertainties related

to the SMA cables or to the bonding between the device and the PCB that

have to be considered. However, the overall resistance of the system never

exceeded the order of magnitude of some Ohms.

In Fig. 3.3 it is possible to see the usual shape of an I –V curve. Since

the device is superconducting at very low bias currents, what happens is

that the I –V curve presents itself as an initially vertical line because the

superconducting nanowire has zero resistance. Once the device reaches the

switching current Isw, a sudden increment in the resistance results as a

reduction of bias current and an increment of bias voltage. That is what

(47)

Figure 3.3: I –V curve for a 100 nm wide, 34 µm long bridge structure constant cross-section and no constrictions, nominally the switching current

should be just one value. Hence, once reached Isw, the whole device should

switch in normal resistive state, and we would observe only one big

transi-tion. However, what happens in reality is that small constrictions cause the

switch of small parts of the device. Then, further increasing the bias

cur-rent, other parts of the device switch, causing a certain number of smaller

transitions.

We noticed that the wider the nanowire, the lower the number of

tran-sitions. That may be due to higher density of constrictions in the thinner

wires, as noticed by Gaudio et al. [18], because of the complications related

to the manufacturing process of ultrathin nanowires, for width value lower

than 150 nm.

(48)

Figure 3.4: I –V curve vs T for a 100 nm wide 5 × 5µm2 meander detector

slightly with respect to the temperature. For example, it is possible to notice

that at lower temperatures, sometimes, the number of small transitions is

lower then the ones at higher temperatures. We think that this effect may be

related mainly to two effects: the first one may be some kind of temperature

dependence behavior of the constriction, which, since their origin in not

completely known and understood, can not be disregarded. The second

one, instead, may be related to the heat generated by the switched portion

of the nanowire: in other words, since the switching current for a wire

section decreases with the increasing temperature, maybe the Joule heating

generated by the nanowire switched portion is not as high as it used to be,

hence the heat power generated only makes switch an even smaller portion

that it used to switch at lower temperatures.

(49)

usually the higher are the switching current and the resistance after the first

transition, the higher is the chance for the device to be photosensitive. That

is because the contribute of quantum efficiency in the device detection

effi-ciency increases with the bias current. Hence, having a damaged nanowire

(constriction) means that the all the other portions of the device is subject

to work at low currents with respect to their critical values. Because of that,

their QE remains low, and the device only works properly in the region near

the constriction.

Also, since Isw will deteriorate with temperature, from Eq. (1.4) one

can notice that having a ”big” pulse at low temperatures is the only way to

hope to have a ”decent” pulse at higher temperatures. A ”sharp” transition

does not mean that the device is going to be photosensitive, but only that

it is a good candidate to be.

From Fig. 3.5 it is possible to see that there are two I –V curves, one

on top on each other, which do not exactly match. That is because the top

curve has been tracked increasing the bias from the superconducting to the

fully normal state, while the bottom one has been obtained reducing the

bias from the fully normal state to the superconducting condition.

This mismatch is given by the so-called hotspot current effect. In fact,

as Skocpol et al. [6] discovered in 1974, usually the switching current of a

bridge is up to 10 times higher than the current required from the device to

self-sustain the normal hotspot.

(50)

Figure 3.5: I –V curve measured first increasing and then decreasing the bias for a 90 nm thick, 1 × 1 nm2 meander structure. The arrows indicate the direction of the bias

3.6, which has been obtained by the I –V characteristic of a WSi SNSPD.

The importance of the hotspot plateau can not be neglected. In fact, in

order to properly bias the device for single–photon absorption, the bias level

should be in the zero voltage state, just below the critical condition. When,

due to photon absorption, the transition takes place, the device follows a

well defined load line and reaches the metastable region. This metastable

region is that between the superconducting branch and the beginning of the

plateau.

For that reason, not having a clear plateau in our I –V curves means

that we do not have an exact idea of the size of this metastable region. This

can be related to the low quality of the films or to the damage caused on

(51)

Figure 3.6: I –V curve with hotspot plateau of a 500 nm wide, 10 µm long bridge structure

3.2.1

Fast–I

sw

measurement

Due to the long time required from the I –V curves to obtain an accurate

estimate of the switching current of the devices Isw, we decided to build in

our set-up a slightly modified version of the fast–Iswmeasurement procedure

developed by Adam McCaughan at MIT [44]. We used an Agilent Wave

Generator in series with a resistance RB to bias the device with a sinusoidal

current signal, Isin(t). Then we used the LeCroy oscilloscope in order to

observe both the bias voltage (channel 1) and the device response (channel

2). The schematics of the set-up is shown in Fig. 3.7.

What happens is that, if the peak voltage of the wave, Vpeak, is such that

Vpeak/RB > Isw for the SNSPD, then the device will undergo a transition

during the bias waving. So on the scope one should see a flat response of

(52)

Figure 3.7: Set-up scheme of the Fast–Isw measurement [44]

Figure 3.8: Expected scope readout of the Fast–Isw measurement [44]

channel 1, while once switched, the device should follow the trend of the

bias voltage, hence jumping from the zero bias to a higher value due to the

sharp transition. What we expect is something like Fig. 3.8.

In order to evaluate the switching current, then, it is necessary to

eval-uate the magnitude of the wave voltage (channel 1) at the instant at which

the device shows the transition (jump on channel 2, say t0) and to divide

it by the bias resistor RB, hence evaluating Isin(t0). A LabVIEW program

was then developed in order to automatize all the process and to collect the

(53)

Figure 3.9: Fast–Isw measurement results

3.3

The R vs T characteristics

The R vs T test is the second test we used to do while ”screening” the

de-vices. Since our goal is to demonstrate working devices at high temperature

(∼20K), it is fundamental for our purposes to test devices which have high

critical temperature.

In order to do the R vs T measurement, we increased the temperature

of the device above 39K (which is the critical temperature for bulk MgB2). Using the Stanford Research Systems SIM921 AC Resistance Bridge, the

device was biased by an AC sinusoidal current (which we usually chose to

be Ipeak = 30 nA) in order to evaluate the resistance of the device itself. We

chose such a small current because, with increasing temperature, the

switch-ing current of the device is reduced, and we did not want the bias current

(54)

Then, simply switching the temperature controller’s heaters off, the

de-vice would start cooling down, a LabVIEW program would read both the

resistance and the temperature of the device and collect the data for the

curve.

It has to be said that the critical temperature of the film and the one of

each device obtained from it may differ very much. In fact, apparently the

production process of MgB2, as well as the one of NbN films (or Nb, MoGe, etc.. [42]), may affect the superconductivity properties of the devices.

On the other hand, knowing the critical temperature of the film before

processing is a good starting point, considering that usually the critical

temperature reduction is about 20% (in the best cases) up to the complete

loss of superconductivity even at base temperature (which usually happens

for very thin, i.e. w ≤ 90 nm, devices).

Again, a distinction between the ideal behavior of the device and the real

one has to be pointed out. Nominally, for a perfect nanowire, the critical

temperature should be homogeneous all over the film, and, consequently, all

over the device. What in reality happens, however, is that the Tc changes

slightly along the nanowire, due to modifications of the superconductivity

properties (film inhomogeneities or fabrication damages).

From Fig. 3.10 it is possible to appreciate a common R vs T curve. From

this curve, we need to focus our attention basically on two parameters: the

first one is the Tc, which is the temperature at which the device reaches

(55)

Figure 3.10: R vs T curve of a 100 nm wide, 5 × 5 µm2 meander structure.

For this device, the critical temperature Tc is around 15.5K,

while the transition width ∆T is ∼2.5K

transition, i.e. the ∆T between the conditions at which the device resistance

is R0.9 = 0.9Rn and R0.1 = 0.1Rn.

This is a very good value to have, since it is a meter of the homogeneity

of the device. The smaller is ∆T , the more homogeneous are the device’s

properties.

3.4

Photo-sensitivity Tests

After the previously described tests, the basic photo-sensitivity test was

performed at the base temperature. This test was basically the last one of

the ”preliminary tests” group.

In order to do that, the AC termination of the bias-tee was connected

(56)

Figure 3.11: Classic shape of a Pulse

very close to the switching current, using the high photon energy 635-nm

laser at high power levels (fractions of a milliwatt), the research of the pulse

was performed.

Before going on, it is necessary to make a little consideration. In order

to avoid latching, to study the photo-sensitivity of the devices we used to

add a series inductance Ls = 6.5 µH to the devices to slow down the decay

time τfall of the response.

If a pulses was present, like in Fig. 3.11, then, setting the voltage

trig-ger level of the Agilent universal frequency counter to half of its height, the

count rate was evaluated while reducing the laser power. That was made

because changing the amount of photons striking on the device should

rep-resent, for a properly working photo-sensitive device, a change in count

(57)

If the device was photo-sensitive, the next step was to check the

polar-ization sensitivity using the manual fiber polarpolar-ization controller. Biasing

at high current and using high laser power in order to maximize the

num-ber of counts per second, manually modifying the polarization of the light

should in principle change the count rate. This because the coupling of the

electromagnetic wave to the electrons increases when they are parallel to

each other. For that reason, when the polarization of the light is parallel to

the nanowire the count rate is maximum, while when it is perpendicular to

it is minimum. Once found the positions of the pads which maximize and

minimize the count rate, after subtracting the system dark counts rate, the

ratio Rpolariz = PCRmax/PCRmin between the photons count rate at the two

conditions was evaluated as reference parameter for the polarization effects.

3.4.1

Single- or Multi-Photon Sensitivity Test

This test was performed in order to check if the device was single-photon

sensitive. First of all, using the temperature controller, the device was

maintained to a fixed temperature, with a maximum fluctuation of the order

of tenths of millikelvins. The device was then illuminated with a fixed laser

power level, while the bias current was swept over a range which had as

upper limit the switching current. During every step, the bias current and

the count rate (CR) were collected. This procedure was repeated for several

laser power levels. After that, the light was shut off, and the system dark

(58)

Once all the data were collected, the first step was to evaluate the actual

photons count rate (PCR) as the difference between the count rate (CR)

and the system dark count rate (SDCR). Then, since the amount of photons

in the laser light flux is proportional to the laser power P, we plotted the

photons count rate divided by P, P2 and P3 vs Ibias. In fact, in accordance

with Eq. (1.3), nominally a single-photon sensitive device should manifest

a linear dependence between the PCR and the amount of photons

strik-ing on the device, a two-photon sensitive one should manifest a quadratic

dependence between PCR and amount of photons, and so on.

In order to be single-photon sensitive, the PCR/P curves evaluated at

different laser powers should in principle coincide. As a consequence, for a

two-photons sensitive device the PCR/P2 should coincide instead, and so on.

In Fig. 3.12 it is possible to appreciate the difference for the results

ob-tained for a single-photon sensitive device and for a three-photons sensitive

one.

3.4.2

Device Detection Efficiency and Afterpulsing

Even if the ratio PCR/P is very useful for qualitative data such as the

discrimination of a single- or a multi- photon sensitive device, it is necessary

(59)

(a) PCR/P for a single-photon sensitive device

(b) PCR/P for a three-photon sensitive device

(c) PCR/P3 for a single-photon sensi-tive device

(d) PCR/P3 for a three-photon sensi-tive device

Figure 3.12: Difference in the photo-sensitivity curves for a single- and a three-photons sensitive detectors. The single-photon sensitive device (Figs. (a)–(c)) was a 90 nm wide, 2 × 2 µm2 meander,

while the three-photon sensitive device (Figs. (b)–(d)) was a 100 nm wide, 1 × 1 µm2 meander

To do that, we used the equation:

PCR = DDE × µ = DDE × P P’ I’ 1 E  × Adevice (3.1)

where µ is the mean number of photons striking on the device per unit time,

P’ and I’ are calibrated power and flux obtained for the optical diffuser, E

is the energy of a photon and Adevice is the active area of the device.

(60)

Figure 3.13: DDE vs Ibias/Isw curve for a single-photon sensitive detector

at different temperatures. The device was a 100 nm wide, 1 × 1 µm2 meander, enlightened with 635-nm laser at 200µW

power. It is possible to notice that the DDE vs T presents a negative slope for 3.15K < T < 6K and a positive slope for 7K < T < 14K

Ibias/Isw), is to slightly decrease with temperature [47]. However, as it was

demonstrated by Inderbitzin et al. [46], sometimes it is possible to see the

trend to move in the opposite direction, i.e. the efficiency may increase

with temperature at fixed reduced bias current. Examples of the possible

representations of a DDE vs Ibias/Isw and a DDE vs T curves are Figs.

3.13–3.14.

Another possible effect which may be hidden behind an unusual DDE

curve could be the presence of afterpulsing. This effect has no actual

expla-nation to date, but it is possible to see whether the device is afterpulsing or

not performing the inter-arrival time (IAT) measurement.

(61)

Figure 3.14: DDE vs T curve for a single-photon sensitive detector, a 100 nm wide, 1 × 1 µm2 meander, at Ibias/Isw = 99% and 99.5%

Ibias and illuminating it with a light source. It is useful to bias at some

relatively high current, since, as we know [28], the afterpulsing effect is

current-related, and having a device which is not afterpulsing at very low

bias, where the count rate is so low that is far from the operative conditions,

is not of any use. That is why we usually biased at the highest ”stable”

current we could reach with the device.

To perform the measurement, there are multiple modes. The one that

we used was made connecting the AC termination of the bias-tee to the

amplifiers suit and then to the oscilloscope, use a high photon energy CW

light source (λ = 635-nm) and, while triggering on the pulse, make an

histogram of the time difference between the main and the extra pulses

which appear on the oscilloscope screen in a time domain of about 5 µs after the main pulse. A correct IAT should present a dead time τdead where

(62)

Figure 3.15: Afterpulsing usual scope trace

the detector measures no counts, followed by a quick rise to a flat region.

Afterpulsing is detected by noting large peaks in the IAT histogram.

The difference between an afterpulsing and a not-afterpulsing detector

(63)

Figure 3.16: Inter-arrival time of an afterpulsing device. The device was a 90 nm wide, 1 µm long bridge structure. The IAT was performed using 635-nm wavelength laser at 200 µW power biasing at 98% of the switching current

Figure 3.17: Inter-arrival time of a not-afterpulsing device. The device was a 100 nm wide, 1 × 1 µm2 meander structure. The IAT was performed using 635-nm wavelength laser at 200 µW power biasing at 98% of the switching current

(64)

Best Devices Results

During the study, we had the chance to face some interesting devices. Here,

we report the most significant ones. The chapter will be built in sections,

each regarding one specific device. It seems a good choice not to overlap

concepts regarding different devices at the beginning. However, in the final

section a more general discussion will be addressed, combining the results

obtained during the whole period of study.

The devices we will discuss come from 4 different films. The properties

of the films are collected in Tab. 4.1.

Wafer No. Production process: Rsq [Ω/sq] TC [K] RMS Roughness [nm] 40 HPCVD 48–60 40 2.1 42 MBE ∼90 31 .4 43 HPCVD 48–50 40.5 1.9 45 HPCVD 85–95 39.5 1.71

Table 4.1: Films properties

(65)

(a) 4 Samples for each Wafer (b) 60 Devices each Sample

(c) Difference between rows A-B, C–H and the extra devices in the corners (R)

Figure 4.1: Design image of the Mask used for fabrication. In dark green there are the pads and the ground plane, in light green the inductors in series with the devices of rows A and B, and in purple the devices. R devices, since are very wide, are in green like the inductors.

(66)

were 4 samples. In each sample there were 8 rows (A to H), 4 columns for

A and B (since the devices were connected in series with inductors) and 8

columns for the other rows. Moreover, there were also 4 extra devices in the

4 corners of the sample, which where assigned the letter R. Usually these

devices were wider than the other SNSPDs, because they were mainly used

as a fabrication process comparison terms, to check if narrower devices had

more suppression in critical current density or critical temperature. The

wafer design is shown in Fig. 4.1(c))

The name of each device was unique. For example ”v35 S3-5F” stands

for ”device of the 35th fabrication run, sample (S) 3, row 5, column F”.

When no S is present in the name, it means that the wafer was designed

with only one sample.

4.1

v40 S4-3H

This device was a 2 × 2 µm2 meander structure made by a 90 nm wide

nanowire with 420 nm spacing (Nsq = 111). The film was obtained by

HPCVD deposition. The device’s geometry is represented in Fig. 4.2.

The device had a sheet resistance Rsq= 57 Ω/sq, which was within the

range of film’s original sheet resistance.

At the temperature of 4K, the switching current Isw was 172.5 µA (see

Fig. 4.3). The critical temperature, as shown by Fig. 4.4, was around

(67)

(a) Design of the device (b) SEM of the device

Figure 4.2: v40 S4-3H

Figure 4.3: I –V characteristic of v40 S4-3H at different temperatures. It is possible to notice how the transitions, with increasing temper-ature, become smoother and then disappear almost completely for 35K

implies, for such a high critical temperature, a pretty homogeneous device.

The device was single-photon sensitive up to the temperature of 18K,

as shown in Fig. 4.5, both at 635-nm and 1.55-µm wavelengths. Anyway, the DDE was pretty low, since biasing at 99.5% of the critical current was

around 5 × 10−7% at 4K and reached a maximum of 10−4% at 16K (see Fig. 4.8). This behavior of increasing DDE with temperature can be explained

since the device efficiency did not saturate. Due to that, increasing the

(68)

Figure 4.4: R vs T characteristic of v40 S4-3H

Figure 4.5: Single-photon sensitivity test at 1.55-µm wavelength for v40 S4-3H, T = 18K

may explain such mechanism [46].

However, at temperatures higher than 18K something unexpected

hap-pened. At 19K, the device started showing a second, smaller pulse from

the oscilloscope, which became the main pulse at 20K (Fig. 4.6). From the

inter-arrival time measurements, we got some interesting results. At 18K

the device does not show any sign of afterpulsing (see 4.7), while at 19K

(69)

(a) T = 18K (b) T = 19K (c) T = 20K

Figure 4.6: Pulse track of v40 S4-3H at three different temperatures

(a) T = 18K (b) T = 19K (c) T = 20K

Figure 4.7: Inter-arrival time measurements of v40 S4-3H at three test tem-peratures

usual afterpulsing shape. It may just be an increment in DDE.

We have no reasonable explanation for that, only some hypotheses. The

most possible guess could be that the device is becoming more efficient,

since now two or more constrictions of the detector are working together to

detect photons.

The reason why it only happens at a certain temperature has been

guessed in this way: let us say that the device shows two constrictions.

Constrictions induce a suppression both in switching current and critical

temperature. If the first constriction has lower switching current and higher

critical temperature of the second one, sooner or later, due to the Isw–T

re-lationship, the latter will take over, becoming the prevalent. Hence, instead

Riferimenti

Documenti correlati

I beni comuni rurali nelle aree montane – come le foreste di proprietà o uso collettivo, i pascoli, le malghe – e le risorse in essi contenute subiscono in modo specifico

introduction &amp; objectives: To our knowledge the impact of venous tumor thrombus (VTT) consistency in patients affected by renal cell carcinoma (RCC) has

As expected we found a difference in depressive symptoms between set-shifting groups but only interoceptive awareness as measured by the EDI-2 differed between intact versus

Several of the analytical methods mentioned are discussed in further details in the case study section, where the obtained results are used to demonstrate the opportunities

ABC, ATP binding cassette; AE: adverse event; CCyR: complete cytogenetic response; CHR: complete hematological response; CML: chronic myeloid leukemia; CTC-AE, Common

Come si può notare, le immagini consegnate dalle due cobles sono sostanzialmente identiche, sebbene sia diverso l’oggetto che subisce l’azione dei venti: nel primo

Concerning the opportunity to use the social networks to promote public health intervention programs, it would be desirable that campaigns like IBC could be performed not