You are on page 1of 6

NIMA

49274

ARTICLE IN PRESS
Nuclear Instruments and Methods in Physics Research A ] (]]]]) ]]]]]]

Contents lists available at ScienceDirect

Nuclear Instruments and Methods in


Physics Research A

journal homepage: www.elsevier.com/locate/nima

7
9
11
13
15
17

The role of the Resistive Plate response function in bringing an RPC to a


stationary situation
o b, C.
D. Gonzalez-Diaz a,, M. Morales b, L. Franco Ferreira b, J.A. Garzon b, D. Gonzalez-Castan
c
Pecharroman
a

19

r Schwerionenforschung, GSI, Darmstadt, Germany


Gesellschaft fu
Universidade de Santiago de Compostela, USC, Santiago de Compostela, Spain
c
Consejo Superior de Investigaciones Cientcas, CSIC, Madrid, Spain
b

21
23

a r t i c l e in fo

abstract

Keywords:
RPC
Time-of-ight
Rate capability
Transients
Dielectric response function

We model the transient behaviour of an RPC exposed to intense irradiation by using as input the
dielectric response function of the Resistive Plates measured experimentally. This work generalizes
previous models to the (most common) case where the RPC response cannot be seen as a simple parallel
RC circuit. Actually, the standard simplifying assumption that the corresponding RC constant can be
estimated from the permittivity  measured at high frequencies (1 ) clearly fails at describing the
observed RPC transients, while if using the measured dielectric response function a remarkable good
agreement is obtained.
& 2009 Published by Elsevier B.V.

29

27

25

PR

31
33
1. Introduction

37

In the rst instants when an RPC is suddenly exposed to high


irradiation a self-regulating process takes place, only after which
the average voltage in the gap becomes stable [1,2]. Such
stationary situation is often referred as the DC limit and the
behaviour of the chamber there described on the basis of the DC
model [24]. The essence of this stabilization process is nothing
but the balance between the rate of charge injected in the gap
(from the avalanches) and the rate at which the resistive
electrodes can get rid of it.
Despite these facts are well established, the complexity of the
relaxation process in materials commonly used in RPCs [5,6] is
rather discouraging when aiming at a quantitative description:
RPC modelling at high rates is therefore attempted only after the
pertinent time is awaited [2,7] or, if experimental conditions do
not allow for that, a cautious note is added indicating that results
must not be extrapolated to the case of continuous irradiation
[8,9].
It is not completely fair to say that the rst instants of
irradiation have little practical importance: one can nd a good
example of this at the GSI-SIS where both HADES [10] and FOPI
[11] tRPC walls have to deal regularly with spills of 510 s duration
at a duty cycle of 50%, the same time scale at which relaxation
takes place in glass. In particular, a strong transient behaviour of

49
51
53
55
57
59

TE

EC

47

45

43

41

39

35

61
63
65

Corresponding author.

E-mail address: D.Gonzalez-Diaz@gsi.de (D. Gonzalez-Diaz).

the average time t 0 as a function of the spill duration has been


recently reported [2], even at moderate rates.
This paper addresses transients in RPCs from the theoretical
and experimental point of view, and it is structured as followed:
(i) introduction, (ii) experimental measurement of the dielectric
response function of glass as the one used in the HADES tRPC wall,
(iii) measurement of transient currents in tRPCs under X-ray
irradiation, (iv) theoretical framework, (v) comparison with data
and (vi) conclusions.

67
69
71
73
75
77
79
81

2. Measurement of the glass response function

83

The dielectric response function of any material spans


over orders of magnitude, from the space-charge region at
low frequencies to the optical transition regime at very
high frequencies [12]. Nevertheless, for GHz impulse signals
in homogeneous materials and as long as electrode-related
effects can be discarded [13] the dielectric response function is
mainly concerned with the so-called dipolar relaxation or,
simply, relaxation. This is expectedly the case of oat
glass when the response to an avalanche produced in state of
the art RPCs is to be evaluated [1416]. As a matter of
fact, the real part of the permittivity of oat glass when an
external eld impulse is applied increases as a function
of the time, due to the progressive orientation of its permanent
dipoles, going from a value at high frequencies 0r;1 to a
static (relaxed) value 0r;s [18], with the number of dipoles
roughly proportional to the difference between both

85

0168-9002/$ - see front matter & 2009 Published by Elsevier B.V.


doi:10.1016/j.nima.2008.12.097

Please cite this article as: D. Gonzalez-Diaz, et al., Nucl. Instr. and Meth. A (2009), doi:10.1016/j.nima.2008.12.097

87
89
91
93
95
97
99
101

NIMA

49274

ARTICLE IN PRESS
D. Gonzalez-Diaz et al. / Nuclear Instruments and Methods in Physics Research A ] (]]]]) ]]]]]]

Vs

d4

d3
5

d2
7

[s]

d1

9
11

h
+

13
15

60

17

40

19

25

20

27

49
51
53
55
57
59
61
63

50

60

EC

TE

magnitudes.1 Actually, 1 means in this context the frequency


where the optical transitions become important: as long as the
characteristic frequency of the impulse signal under study is far
from that region, all the relevant informations are conveyed by the
magnitude 0r;1 .
Provided the relaxation of oat glass happens at a time scale of
seconds [18], its dielectric response function can be readily
accessed by a direct measurement in the time domain. Therefore,
rectangular samples of oat glass with thickness h 0:18 cm were
studied through application of steps of increasing and decreasing
HV with the help of an external switch (Fig. 1). In order to reduce
edge effects circular electrodes of diameter d1 2 cm were used.
The induced current Id is given by [12]:
Id t C 1 V 0 dt C s  C 1 V 0 Ct

(1)

where Ct is dened, following the convention of [12], as the


dielectric response function, of integral unity:
Z 1
o 1 s  1
Cteiot dt.
(2)
0

When HV is applied, the ohmic contribution Ic V 0 R is also


present, being I Id Ic . We evaluated the uncertainty (and
reproducibility) of the experimental procedure by taking the
average and rms of groups of three measurements of It as shown
in Fig. 1 for a typical case.
In order to discard electrode-related effects [13] different
electrode materials, treatment and type of contact were tried. The

81
83

300

400

500 600
HV [v]

700

800

900 1000

samples were cleaned with ethanol before attaching the metallic


electrodes and the measurements were performed in air with
temperature and humidity being monitored (but not controlled)
showing averages of 27  C and 40%, respectively, with variations
within 10%. By analogy with a Debye relaxation model and
following a previous work from Fraga [6], we tted the relaxation
curves to a sum of two exponentials:

Ct A1 et=t1 A2 et=t2 .

We note the usual denition o 0 o  i00 o.

(3)

Aiming at a microscopic interpretation of the tting parameters


the use of a stretched exponential is common [17], but this
approach is out of the scope of the present work.
For an instantaneous impulse V 0 , the voltage drop in the glass
can be calculated as
Z
Vt V 0 Ct dt
(4)

and so, after requiring unity integral of Eq. (3) and imposing the
physical condition that the Ohms law is fullled for an innite
train of impulse shots each one with the response given by Eq. (4),
only three relevant parameters are left: 0r;s , t1 and t2 , with A1 and
A2 being function of those. The high linearity measured for the
dielectric response function is illustrated in Fig. 2 both for the
polarization (HV on) and depolarization (HV off) curves as a
function of the applied voltage. Taking the linearity for granted
(that is an expected feature of oat glass at low elds) we evaluate
the uncertainty of the parameters from the dispersion of all the
points, yielding t2 10:7  2 s, t1 1:5  0:5 s, 0r;s 29  5. The
latter value is in close agreement with existing measurements for
oat glass [18]. On the other hand, the determination of the
resistivity is far more accurate, yielding r 3:25  0:1 TO cm and
ohmic behaviour. The response function is completely
characterized only after obtaining 0r;1 7:5  0:5 with an LCRmeter (constant in the range 1 kHz1 MHz), being this magnitude
seldom accessible by measurements in the time domain.
We observed that adhesive Cu tape with or without conductive
rubber behaves very similarly to evaporated Au or Ag, unlike Ag
acetate or Ag polymer.2 Due to the different material used, it is
therefore discarded the presence of electrode-related effects in

65
1

77
79

PR

Fig. 1. Up: setup used for the measurements of the dielectric response function
(the region compressed between d2 and d3 is the guard ring). For the
measurements without guard ring both upper and lower electrode are chosen to
have the same diameter. Down: polarization and depolarization currents
measured for the maximum HV (1 kV); the lines show the t to a double
exponential.

47

40

45

30
time [s]

43

20

41

10

39

75
Open symbols: HV off
Closed symbols: HV on

37

73

Fig. 2. HV dependence of the parameters characterizing the dielectric response


function of glass (t1 , t2 ) as a function of the applied voltage. The electrodes were
made of Cu tape in this case.

40

29

35

71

33

69

23

31

67

20

I [nA]

21

15
14
13
12
2
11
10
9
8
7
6
5
4
3
1
2
1
0
100 200

RS references 186-3600, 321-7295.

Please cite this article as: D. Gonzalez-Diaz, et al., Nucl. Instr. and Meth. A (2009), doi:10.1016/j.nima.2008.12.097

85
87
89
91
93
95
97
99
101
103
105
107
109
111
112
113
114
115
116
117
118
119
120

NIMA

49274

ARTICLE IN PRESS
D. Gonzalez-Diaz et al. / Nuclear Instruments and Methods in Physics Research A ] (]]]]) ]]]]]]

7
9
11

3. Measurements with a small tRPC under X-ray irradiation

15

We built a small 3:4  3:4 cm2 chamber with Cu electrodes and


nylon spacers of 0.3 mm diameter, lled with the standard
mixture [14] at slight overpressure and operated at HV 3 kV.
Signal amplication was performed with a modied version of the
HADES FEE [21] based on a Philips BGA2712 and discriminated at
a threshold of 7 mV. The chamber was irradiated with an X-ray
tube (Fig. 3) that, after attenuation with Cu absorbers, showed a
well dened energy peak of 12 keV and 10% width.
To verify the timing properties of X-ray current modulation
through the tube control, we used a pin silicon diode as a solid
state ionization chamber. The observed current transients were
below 0.5 s over the full dynamic range.
From the measurements of the RPC rate at different HV and
thresholds we concluded that the measured rate is unlikely to be
underestimated by more than 50%. This is, nevertheless, the main
experimental uncertainty a priori when trying to model the
situation, and will be discussed thoroughly in the next section.
Therefore, the best estimate of the actual avalanche ux in the
detector f was the one measured by the RPC itself fRPC (kHz/cm2)
that was observed to be proportional to the current of the X-ray
tube for low irradiation. From the linear region of the curve fRPC
vs. Itube the value of the former was therefore extrapolated to the
highest irradiation case.
The charge spectra of the avalanches were determined by
numerical integration of the signals registered on a digital scope,
verifying an approximate exponential shape either in the presence
of a Na22 source or X-ray photons as previously reported for
narrow RPCs [19]. Then, we proceeded to monitor the current
driven by the chamber when applying steps of current either
increasing or decreasing for a time span as long as 2000 s (Fig. 6).
Being the RPC gap g 0:3 mm and the dimensions of the
electrode A 3:4  3:4 cm2 , the maximum angle of incidence of a
photon that fully crosses the gap is yc 0:6 . Taking into account
that low energy X-rays are strongly attenuated in the electrode
glass, it can be estimated that a misalignment between the RPC
and the direction of incidence of the X-rays by only 2yc reduces
the illuminated area by as much as a factor 2, therefore
questioning the interpretation of any theoretical description.
Experimentally, the aligned position was identied with the
help of an 0:1 resolution goniometer, as an abrupt decrease of
counting rate at yc 0:6 around such position.

35
37
39
41
43
45
47
49
51
53
55
57

33

TE

31

EC

29

27

25

23

21

19

17

4. Theoretical description

59
61

The theoretical description is based on the previously developed MC from [2], adapted for coping with the present situation:
the detector is sub-divided in virtual cells of area Acell following

63
3

65

71
73
75
Fig. 3. Setup used for X-ray irradiation. The RPC was placed inside a gas box with a
thin mylar foil as the entrance window. The relative orientation between the X-ray
tube and the RPC was adjusted with an 0:1 resolution goniometer.

Ref. [22] and each one is assumed to react to every avalanche with
a dielectric response function given by Eq. (3), causing an
instantaneous voltage drop given by Eq. (4), that is taken into
account when the next avalanche is created. The gap capacitance
is merely added to the total parallel capacitance. The avalanche
dynamics is modelled in a simplied way by assuming proportionality between the voltage in the gap V and the average
avalanche charge hqi, in the form hqi aV  V th [19].
We focus on the induced current and distinguish, for each cell,
three contributions: (i) the current induced due to propagation in
the gap, (ii) the current instantaneously induced in the glass and
(iii) the current induced due to the glass relaxation:
Id t Idga t Idgl;inst t Idgl;relax t
N av
X

Idga t gEw

qj t j dt j

The surface conductivity is in general dependent on ambient gas and water


content, so it is not completely rigorous to extrapolate from these measurements
to the actual situation inside an RPC.

77
79
81
83
85
87
89
91

(5)
93
(6)

95

Idgl;inst t 1  gEw

PR

13

69

67

the measured dielectric response function when using Cu tape.


This is important, since this type of electrode is very popular and
in particular is the one used for our investigations. The way the
contact was done proved important for the LCR measurements at
high frequencies but otherwise either soldered or with contact
made by pressure with a Au probe yielded similar results.
Measurements performed together with a Cu guard ring show
compatible results once edge effects were corrected for [20],
therefore surface conductivity in oat glass seems to play a minor
role as compared with bulk conductivity.3

97
N av
X
qj t j

Cs

Idgl;relax t 1  gEw

C 1 dt j

N av
X
qj t j
j

Cs

(7)

99
101

C s  C 1 Ct  t j

(8)

103

being Ew the weighting eld, Nav the number of avalanches in the


given cell and C the total RPC parallel capacitance. In order to
calculate the time dependence of the total current, we evaluate
the discrete functions (6) and (7) as

105

(9)

109

Idga jT fRPC tAgEw hqti


Idgl;inst jT fRPC tA1  gEw

hqti
C1
Cs

111
(10)

with brackets representing the average over all the RPC cells. Eq.
(8), being analytically tractable, is computed directly from the
sum over all the N cells:
Idgl;relax jT

N av
N X
1  gEw C s  C 1 X
qij t ij Ct  tij
Cs
i
j

107

112
113
114

(11)
115
116

4.1. The DC limit


117
The parameters that rule the average avalanche dynamics are
tted to reproduce the DC values of the current that can be
calculated in MC as well as analytically [23]:
IDC

afAV  V th
idark .
1 afAR

119
(12)

The fullment of this consistency condition of MC is actually


better illustrated in Fig. 5 (the analytical value of DV DC is indicated
with dashed lines).

Please cite this article as: D. Gonzalez-Diaz, et al., Nucl. Instr. and Meth. A (2009), doi:10.1016/j.nima.2008.12.097

118

120

NIMA

49274

ARTICLE IN PRESS

11
13

DC limit

15

17

5
19

25

data
simulation

3
4.2. The transitory region

Idark

29

PR

100 200 300 400 500 600 700 800 900 1000
HV [v]

20

45

10

10 Hz/cm2

81
83
85
87

R
O

51

HV [v]

53

50

95
97

103
105
107

0.5

50

0.05

109
111
112

49

93

1.5

100
0.1

47

0
100 150 200 250
time [s]

100 Hz/cm2

50

100 150
time [s]

113

200

114

1000 Hz/cm2

800

4000

600

3000

400

200

115

40
V

glass

DC

2000

20

116

30

/R
DC

117
118

10

1000

61

119

65

79

101

ln1 afrh

150

I [nA]

HV [v]

43

63

77

0.15

EC

15

59

75

and again, a and f appear always together. This behaviour is not


really surprising since the quantity af is proportional to the

0.2

41

57

73

99

TE

1 Hz/cm2

39

afrh

37

55

71

(13)

t eq

Fig. 4. MC description (line) in the DC limit after xing a and V th , and comparison
with data (points).

I [nA]

35

It was argued before that the effect of the nite comparator


threshold can be absorbed in the parameter a when looking at the
behaviour of the current in the DC limit. In fact, the situation is
very similar for the transitory part: resorting to the simple
derivation given in Ref. [2], one can see that the equilibration time
of the voltage drop across the glass is well approximated by

33

69

91

27

31

67

89

23

I [nA]

21

I [nA]

I [nA]

the total current driven by the RPC is not dependent on Acell


provided its value is kept within 0.154 mm2. On the other hand,
the assumption of different charge distributions (box, Poisson,
exponential, 1=q were tried) has also little impact in the predicted
total current.
Therefore, once a and V th are xed from the t in the DC limit,
the parameters relevant for the MC are taken from the dielectric
response function measured in Section 2. Due to the slightly
different temperature at which the RPC was operated T 20 
1 we recalculated the value for r taking the activation energy
previously measured in Ref. [19] and obtained r20 6:2 TO cm.
The inuence of the temperature on t1 , t2 was evaluated by
making the ansatz that the associated mechanisms have the same
activation energy as the one ruling the resistivity [18], resulting in
t120 3 s, t220 20:5 s.
Fig. 5 illustrates the behaviour of the current and voltage drop
in the glass(gap) DV glass as in MC. Two main consistency criteria
are fullled: the asymptotic limit of DV glass coincides with the
analytical expectation DV DC (see Ref. [2], for instance) while the
current in the system is given by DV glass =R in the same limit.
Needless to say that this agreement requires all terms in Eq. (5) to
be considered.

HV [v]

The t of the current in the DC limit is shown in Fig. 4. As can


be noticed from Eq. (12), the actual assumption on the primary
ux is not crucial for obtaining a good t: if a certain inefciency
is assumed so that fRPC ef due to the nite value of the
comparator, it will be re-absorbed in the tting parameter a as
a=e.
There are yet two main uncertainties: the particular value of
the area of inuence of the avalanche Acell and the shape of the
total charge spectra; for the rst Acell 0:5 mm2 was taken while
for the second a box distribution was assumed. The cell area was
chosen so for realism ([2], for instance), however, the box
distribution was a simplifying choice. The remarkable fact is that

HV [v]

D. Gonzalez-Diaz et al. / Nuclear Instruments and Methods in Physics Research A ] (]]]]) ]]]]]]

20
40
time [s]

60

10
time [s]

15

Fig. 5. Behaviour of the current and DV glass as a function of the irradiation time for four characteristic uxes, as obtained from MC.

Please cite this article as: D. Gonzalez-Diaz, et al., Nucl. Instr. and Meth. A (2009), doi:10.1016/j.nima.2008.12.097

120

NIMA

49274

ARTICLE IN PRESS
D. Gonzalez-Diaz et al. / Nuclear Instruments and Methods in Physics Research A ] (]]]]) ]]]]]]

25

20

data
simulation
2

[Hz/cm ]

1000

67

900

69

800

71

700

73

15

600

I [nA]

9
11

500
10

13

400

RPC [Hz/cm2]

RPC

300

15

17

75
77
79
81

200

83

100
19

85

0
21

87

100 200 300 400 500 600 700 800 900 1000 1100 1200
time [s]

23

99

0.6

600

0.5

500

0.4

400

300

0.3

300

200

0.2

200

107

45

100

0.1

100

109

800 850
time [s]

900

700

750

800 850
time [s]

900

47

PR

D
TE
750

700

EC

43

0.8

m/mo

I [nA]

41

700

400

39

59

0.7

500
10

Fig. 7. Close look at Fig. 6. The current is shown in the left gure while the right one shows the relative variation of the gain m=m0 in the same time range, as obtained in
simulation.

average total charge, that expectedly rules both the transitory and
the DC situation. The qualitative behaviour indicated by Eq. (13)
was veried by MC upon variations of the primary ux up to a
factor 4, resulting in little changes of the induced current, if this
factor is re-absorbed in the parameter a. So, we conclude that a
moderate underestimation of fRPC by a constant factor is not
relevant for the quantitative understanding of transients in
current as long as we do not aim at a quantitative interpretation
of the parameter a also (that is not the purpose of this work).

57

97

600

15

37

55

800

700

35

53

95

800

20

33

51

900

0.9

900

31

91
93

1000

RPC [Hz/cm2]

29

49

1000

25

27

RPC [Hz/cm2]

25

89
Fig. 6. Transients on the measured current and comparison with simulation for different uxes (scale on the right). Circles are measured points and line is the output from
the MC.

61
63

5. Comparison with data

65

Fig. 6 shows the current measured directly in the HV supply


(after offset calibration) and the trend predicted by simulation, as

a function of time and for different uxes (values given by


horizontal bars, right axis).
There are a number of situations that deserve attention: (i)
naturally, big transients only appear when the voltage in the gap is
modied substantially with respect to its nominal value, therefore
at low ux the current increases mostly step by step with the ux
(to100 s, FRPC o100 Hz=cm2 ). Otherwise, (ii) the current starts
from a high value and as the next avalanches start to pile-up the
voltage in the gap slowly decreases down to the stationary
situation t 380 s. Even more interesting is the situation (iii)
when the ux is decreased from an already very harsh situation
(t 420 s, for instance), then the current goes suddenly down
allowing for the voltage in the gap to recover slowly according to
the glass time constant. If (iv) a high ux is drastically reduced
down to zero there is only a very little transient, actually coming

Please cite this article as: D. Gonzalez-Diaz, et al., Nucl. Instr. and Meth. A (2009), doi:10.1016/j.nima.2008.12.097

101
103
105

111
112
113
114
115
116
117
118
119
120

NIMA

49274

ARTICLE IN PRESS

D. Gonzalez-Diaz et al. / Nuclear Instruments and Methods in Physics Research A ] (]]]]) ]]]]]]

30

20
I [nA]

charge) as was done in Ref. [6]. The equilibration time is here


dened for convenience as the time it takes for an observable to
reach a fraction 1  1=e of the stationary value [2], and is plot in
Fig. 9 for the current and the voltage drop across the glass(gap) as
a function of the RPC ux. For a ux below 100 Hz=cm2 the gain is
yet little affected so teq is small. At a given point the equilibration
time reaches a maximum and then decreases with ux, exactly as
reported in Ref. [6].

data
Two exponentials (from the measured
dielectric response function)
One exponential, = Cs

25

One exponential, = C

15
10
5

0
980

985

990

995

1000

1005

1010

67
69
71
73
75

time [s]

11
13

Fig. 8. Illustration of the differences between a description based on the measured


dielectric response function and two simplied RC cases.

15

t
t

19

eq
eq

for V

glass

(simulation)

for I (simulation)

21
10

25

teq[s]

23

100

29

27

33
100

101

35

RPC [Hz/cm2]

51
53
55
57
59
61
63

TE

EC

49

47

45

from the third term in the rhs of Eq. (5). A last relevant effect (v)
the second point after high irradiation starts (after a time interval
of 0.5 s) does not reach in some cases the maximum value
expected from simulation by maximum a factor 2. We think that
we may be inuenced by transients of the X-ray tube itself that
may be important below 0.5 s and cannot be detected with our
setup.
The region between 700 and 900 s is zoomed in Fig. 7 shown
together with a plot of the relative variation of the detector gain. It
is apparent that the transient times when ux is increased are
distinct from those when ux is decreased. This is better
illustrated by Eq. (13) and Fig. 9 and arises from the fact that
transient times with increased ux decrease roughly linearly with
the ux itself while transient times with decreased ux are ruled
just by the glass relaxation. Fig. 7 also shows how, as expected, the
total current is a good indicator of the situation in the gap.
As a matter of fact, a description based on an unique RC
constant given by RC 1 ends too short at describing the transient
currents while assuming a time constant RC s provides a reasonable description. This is shown in Fig. 8. Importantly, the very big
discrepancy between 01 and 0s (factor 4) is partially shadowed
once the gap capacitance is considered for calculating the RC
constant.
Finally, we revisited the behaviour of the equilibration time t eq
from the point of view of the total current (equivalently, the total

43

Acknowledgements

Fig. 9. Equilibration time for the current (squares) and for the voltage drop across
the glass (circles) as a function of the RPC ux.

39
41

103

37

102

PR

31

77
79

We extended our previous MC [2] to include the (usual)


situation where an RPC cannot be modelled as a simple parallel RC
circuit. The simulation was driven by the measured dielectric
response function where a simple Debye model with two time
constants was used to t the data. Taking the measured dielectric
response, the simulation describes accurately the transients
observed in a small RPC when exposed to high X-ray luminosity.
Remarkably the only two relevant parameters for the calculation
can be well constrained in the DC limit so the description of the
transients is done without resorting to any extra parameter.
We actually observed that a dielectric response function based
on a simple RC circuit can account also for the transients
reasonably well but not using 01 and rather 0s. This may look
like a simple alternative, but the accurate experimental determination of 0s is generally more complicated than the one of 01 , that
can be accomplished by capacitive measurements.
The framework here introduced provides a theoretical basis for
extrapolating the RPC response from pulsed particle uxes to
continuous ones.

102

17

6. Conclusions

This work is supported by EU/FP6 contract 515876. Special


thanks to Faustino Gomez for many discussions on the X-ray
setup. The authors want to acknowledge also A. Blanco for
providing us with samples of glass from the HADES tRPC wall and
to L. Lopes for many discussions, in particular on Section 2.

81
83
85
87
89
91
93
95
97
99
101
103
105
107
109

References
111
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]

E. Cerron Zeballos, et al., Nucl. Instr. and Meth. A 367 (1995) 388.
D. Gonzalez-Diaz, et al., Nucl. Phys. B (Proc. Suppl.) 158 (2006) 111.
112
G. Aielli, et al., Nucl. Instr. and Meth. A 456 (2000) 82.
C. Lippmann, et al., Nucl. Phys. B (Proc. Suppl.) 158 (2006) 127.
I. Crotty, et al., Nucl. Instr. and Meth. A 337 (1994) 370.
113
M. Fraga, et al., Nucl. Instr. and Meth. A 419 (1998) 485.
C. Gustavino, et al., Nucl. Instr. and Meth. A 527 (2004) 471.
114
A. Blanco, et al., Nucl. Instr. and Meth. A 485 (2002) 328.
P. Colrain, et al., Nucl. Instr. and Meth. A 456 (2000) 62.
Q 1 115
D. Belver, et al., The HADES timing RPC Inner TOF wall, these proceedings.
A. Schuttauff, FOPI new MMRPC-Barrel, these proceedings.
J.M. Albella, Fisica de dielectricos, Marcombo, 1984.
Q 2 116
C. Kim, M. Tomozawa, J. Amer. Cer. Soc. 59 (3) (1976) 128.
P. Fonte, et al., Nucl. Instr. and Meth. 449 (2000) 201.
C. Gustavino, The OPERA veto system, these proceedings.
117
J. Zarzycki, Glasses and the Vitreous State, Cambridge University Press,
Cambridge, 1982.
K. Funke, Prog. Solid State Chem. 22 (1993) 111.
118
H.E. Taylor, J. Soc. Glass Technol. 43 (1959) 124.
D. Gonzalez-Diaz, et al., Nucl. Instr. and Meth. 555 (2005) 72.
119
M. Lisowski, R. Kacprzyk, IEEE Trans. Diel. Electr. Insul. 13 (2006) 139.
D. Belver, et al., Performances of the Front-End Electronics for the HADES RPC
12
TOF wall on a C beam, these proceedings.
120
M. Abbrescia, Nucl. Instr. and Meth. A 533 (2004) 7.
G. Carboni, et al., Nucl. Instr. and Meth. A 533 (2004) 107.

65

Please cite this article as: D. Gonzalez-Diaz, et al., Nucl. Instr. and Meth. A (2009), doi:10.1016/j.nima.2008.12.097

You might also like