You are on page 1of 22

Quantum Noise

quantum noise given by the number of the detected photons and the photo-detector
thermal noise.

From: Optical Spectroscopy, 2006

Related terms:

Photon, Saturation, Raman Spectrum, Faraday Effect, Wavelength, Transmittance,


Sum-Frequency Generation

View all Topics

Optical measurements
NIKOLAI V. TKACHENKO, in Optical Spectroscopy, 2006

4.1.4 Noise sources


There are very many sources and types of noises, also three types are the most
common for measurements in the optical spectroscopy applications.

Quantum noise (or photon noise, or short noise, or Poisson noise) is one considered
in detail in the previous section (eqs. (4.11)–(4.16)). In Example 4.1 the light source
has constant emission rate, however, the experimentally available value is the num-
ber of photons measured in a limited time interval, which is the random value by its
nature. This type of noise cannot be eliminated when dealing with quantum objects
(but in some applications can be reduced to negligible level). This type of noise can
also be found in electric circuits at low current values, when the discrete nature of
current carriers (electrons) prevails thermal noise.

Thermal noise (or Johnson noise) has actually the same origin as quantum noise
but at low frequency limit. At temperature above absolute zero the space is filled by
thermal radiation, which has quantum nature, i. e. it fluctuates. When applied to
electric circuits this noise is called Johnson noise. In all circuits the voltage noise in
a spectrum range Δf is

(4.25)
and the current noise is

(4.26)

where R is the circuit active resistance. The power of the Johnson noise is

(4.27)

The noise power does not depend on resistance R.

1/f noise and generation-recombination noise have spectrum density proportional


to the inverse of the frequency (which has given the name to this type of noise).
Usually it dominates in measurements which take a long time. These measurements
are said to be low frequency measurements. Depending on the measuring technique
and devices used, the frequency limit for the 1/f noise domination can be from 10
Hz (0.1 s in time domain) or smaller.

> Read full chapter

Ultrafast fluorescence parametric am-


plification for femtosecond molecular
dynamics research
Shufeng Wang, ... Qihuang Gong, in Femtochemistry VII, 2006

3 Discussion

3.1 Phase matching


The CCE from the NLO crystal came from the amplification of quantum noise inside.
It output angle satisfied the second-order NLO phase-matching condition. From
calculations and experiments, at an angle of 31.5° between pump beam and optical
axis, a phase-matching angle from 500 nm to 750 nm appeared within a very small
angle distribution. It was proven that the seed input corresponding to this condition
was amplified with a full bandwidth.

3.2 Linearity
To experimentally demonstrate the linear amplification dynamic range, supercon-
tinuum white light from a water cell, produced by focused fundamentals, was used
as a seed. This seed can be strongly attenuated in intensity. For the pump beam
of the parametric process, at high pump intensity above the CCE threshold, the
amplification rate increased more dramatically than in non-CCE conditions. At a
fixed weak seed intensity, the output increased linearly towards the pump intensity
as shown in Fig. 1, left. However, if the pump intensity went too high, the system
output became saturated. As shown in the right figure, with variable seed intensity
under a certain fixed pump intensity, the input-output linearity was proven with a
high amplification rate of about 107. However, the stronger seed also was saturated
at output, which means the system was suitable for the weak signal detection.

Fig. 1. The saturation effect towards pump energy at fixed seed intensity (left) and
seed intensity at a fixed pump energy (right)

> Read full chapter

Origin of Non-Quantum Noise and


Time Dependent Thermo Field Dy-
namics
H. Umezawa, in Thermal Field Theories, 1991

1 Introduction
This talk is a review on two topics related to thermal physics. The first topic is a
general analysis of noises in pure states. Recently, the appearance of non-quantum
noise in certain pure states has attracted the attention of physicists in many areas,
such as the squeezed state in quantum optics1–8, the black hole9–11 and thermal
physics12–18. Although each of these subjects can be formulated in many ways thermo
field dynamics (TFD) has provided a common base for all of these subjects2,3,6,9–11,-
19–22. The analysis of this first topic leads to a formulation for time-dependent TFD

which is the second topic.

> Read full chapter

Flash–photolysis
NIKOLAI V. TKACHENKO, in Optical Spectroscopy, 2006

7.2 Time resolution and signal-to-noise ratio


Since the method is applied to study intermediate and essentially unstable states,
the time resolution of the flash-photolysis instrument is one of the most important
questions. There are obvious time limiting parts, namely the the excitation pulse
width and time resolution of the light detection system. The former is not a real
limit at present as femtosecond pulsed lasers are available nova days. Modern
photomultipliers have time resolution as good as 1 ns and there are commercially
available photodiodes with time resolution better than 100 ps. However, 1 ns time
resolution is difficult to achieve in practical measurements. The limiting part is
usually the source of the monitoring light.

There is a relation between the accuracy (or sensitivity) and time resolution of the
flash–photolysis method. Let us consider an ideal instrument which has no noises of
its own, so that only the light quantum noise will be taken into account. Suppose we
would like to measure light intensity with accuracy and with time resolution . This
means, that in time interval the photomultiplier (PM) should detect 1/ 2 photons,
i. e. the photon detection rate must be ( 2 )−1. Accounting for the quantum yield of
the PM, d, and the registration monochromator efficiency, m2, the monitoring
beam photon flux after the sample should be ( 2 d m2 )−1. Then, we have to add
the sample transmittance, T, efficiency of the monitoring monochromator, m1,
and losses in all the optical components (like lenses, monochromator slits and so
on), which will be called optics efficiency o.8 Finally, an estimation for the light
intensity required at the entrance of the monitoring monochromator is

(7.9)

In the derivation of the equation the light detector was assumed to be a photomul-
tiplier and the dark current of the photomultiplier was neglected. If the dark current
cannot be neglected, as in the case of infrared sensitive photomultipliers, or another
type of photo-detector is used, e. g. a photodiode, the equation must be corrected.
Formally this can be done by taking proper value of d. The inverse value, 1/ d, shows
how many photons are needed to obtain a signal equal to the noise, i. e. to reach
signal-to-noise ratio equal one. Therefore d can be evaluated from noise equivalent
power of the detector, see Section 4.2.1.

According to this estimation to reach 1 ns time resolution with signal-to-noise


ratio 100, i. e. accuracy = 0.01, the monitoring light intensity must be > 1 mW
for flash–photolysis measurements at 500 nm as show in Example 7.2. On the
other hand the total emission spectrum density of a tungsten halogen lamp is
roughly 7 mW nm−1 at this wavelength (see Example 1.3 on page 10). This lamp
emission is spread in all direction and less than 1% can be utilized by the monitoring
monochromator (see Example 2.7 on 38). Clearly, tungsten lamps cannot be used
in nanosecond flash–photolysis measurements at least with reasonable spectrum
resolution.9

Example 7.2 Monitoring beam power


Let us estimate the monitoring lamp power (Im1) needed to achieve = 1 ns time
resolution at = 0.01 = 1% accuracy in otherwise favorable conditions for the
measurements. The quantum efficiency of a photomultiplier can be d = 0.1,
which is close to the top value for most photo cathodes. Efficiencies of the
monochromators can be m1 = 0.5 and m2 = 0.5. The losses in optics include
reflections from the sample cuvette walls, reflections from all lenses (L1, L2 and L3 in
Fig. 7.1), and mismatches between the monochromator slits and focused spots. One
can hope to achieve efficiency as high as 50%, thus o = 0.5. Finally, the sample
absorbance can be assumed to be A = 0.5, which gives sample transmittance T = 10−A
≈ 0.3. Substituting all the values in eq. (7.9) we obtain Im1 ≈1 mW at 500 nm. This
estimation does not account for any noises of the photo-detector and monitoring
light source but the quantum noise of the photon flux, and should be treated as
absolute minimum. Real accuracy with 1 mW monitoring intensity and 1 ns time
resolution may be much worse than 1%.

Arc lamps have higher temperature of the emitting area. At 500 nm the emission
spectrum density of a Xe arc lamp can be as high as 0.3 W nm−1 (see Example
1.4 on page 11), and one can expect to collect about 3 mW at the monitoring
monochromator entrance in 1 nm bandwidth. Thus with arc lamps as the source of
monitoring light one can approach the nanosecond time resolution with reasonable
signal-to-noise ratio.

When the time resolution of flash–photolysis method needs to be improved the


following measures can be considered:

1. to use an arc lamp, which has higher working temperature and higher spec-
trum density of the emission in the visible region;
2. to use a pulsed monitoring light source, which allows to increase the temper-
ature of the emitting body even more, as discussed in Section 7.2.1;
3. to decrease spectrum resolution and, thus, to increase the transmission band-
width of the monochromators, thus increasing intensity of the monitoring
light;
4. to use a laser as the source of the monitoring light, which has many folds
higher spectrum density of the emission as compared to lamps.10

High monitoring light intensity needed for the fast measurements creates yet
another problem. Namely, the monitoring light can excite the sample to extend
which cannot be neglected any more. Example 7.3 shows how to estimate the
relative population of the excited state and also demonstrate that if the recovery after
excitation is longer than 1 ms, the monitoring light intensity must be lower than
1 mW in typical experiment conditions. This is in contradiction with requirement
of the high monitoring intensity needed to achieve nanosecond time resolution.
To solve the problem flash–photolysis experiments are carried out with pulsed
monitoring light as discussed in Section 7.2.1.

Example 7.3 Effect of monitoring light on sample


Let us estimate the fraction of the excited molecules of a chlorophyll a solution
under continuous irradiation by a monitoring beam at 430 nm, which is one of
absorption bands maximum. The molar absorption coefficient of chlorophyll at
this wavelength is typical for organic dyes, 1.2 × 105 M−1cm−1, i. e. absorption
cross section is ≈ 4.6 · 1C−16 cm2. The longest living excited state of chlorophyll
a is a triplet state, which has lifetime roughly T ≈ 1 ms.11 Let us assume that the
monitoring light intensity is I = 3 mW (to achieve ns time resolution), and monitoring
beam area is s = 0.1 cm2, thus the monitoring power density is .

For a simple estimation the reaction scheme is , where C and CT are the chlorophyll
ground and triplet excited states, respectively. The kinetic equation for the case is

(7.10)

where Ng and Nt are the number of molecules in the ground and excited states and
Ng + Nt = N = const is the total number of molecules. On the right side of eq. (7.10)
the first summand is responsible for the increase of the ground state population due
to the relaxation of the excited state, and the second summand is responsible for the
decrease of the ground state population due to photo-excitation by the monitoring
beam. In steady state conditions , thus and the ratio of populations is , i. e. under
such conditions 3% of chlorophyll molecules will be continuously in excited state.

Equation (7.9) can also be used to analyze the accuracy of the measurements, , (or
signal-to-noise ratio, which is −1) after some rearrangement

(7.11)

The first term on the right shows dependence on the sample transmittance, the
monitoring light intensity and the wavelength. The second term collects all sources
of the light losses in the instrument (a trivial conclusion is that a better instrument
will give better results). The third term tells that the increase in the time resolution,
i. e. smaller , will give a decrease in the measurement accuracy, .

High sensitivity of the detector is important for high time resolution. In terms
of eq. (7.9) the sensitivity is hidden under the detector quantum efficiency, d.
Decrease in sensitivity will mean proportional decrease in time resolution. Therefore,
although there are photodiodes with time resolution much better than 1 ns, simple
replacement of a relatively slow photomultiplier by a fast photodiode may result
in gradual decrease in the time resolution under other equal conditions, since the
sensitivity of the photodiodes is much worse than that of the photomultipliers (as
was discussed in Section 4.2).

7.2.1 Pulsed monitoring light


The monitoring light intensity is critical for the fast time resolution. The brightness
of a thermal light source, such as arc lamp, depends on the temperature of the
emitting body, e. g. plasma temperature near the lamp cathode. Obviously there
is a limit after which the lamp electrodes will be destroyed.12 However, for a short
time the arc can be overheated by applying a short high voltage pulse. The emission
temperature of the arc may increase to 10000 K, which gives more than 10 folds
increase in emission spectrum density in the blue part of the spectrum as compared
to normal continuous lamp operation with 6000 K cathode area temperature.

In the pulsed mode the lamp is supplied continuously with a relatively low current.
This is the current needed to keep duty ark, but the brightness of the lamp at
this time is at least 10 times lower than in the normal operational mode. A few
microseconds before the excitation flash a high voltage pulse is applied to the lamp.
The voltage pulse has a special temporal profile to generate the light pulse with the
shape as close to rectangle pulse as possible, as illustrated in Fig. 7.6.

Figure 7.6. Utilization of a flash lamp to increase the monitoring light intensity
during a short measuring time interval.

The pulse width can be as long as 1 ms. However, it is difficult to keep constant
emission intensity during this time with accuracy, e. g., 1%. On the other hand
the shape of the pulse can be well reproduced from pulse to pulse. Therefore,
the measurements can be repeated twice: first without excitation pulse, giving the
monitoring pulse temporal profile in the measuring time window, Ubg(t), and then
with the excitation pulse, U(t). The ratio of the measurements, , is normalized
signal with “compensated” monitoring pulse shape, which can be used to calculate
differential absorbance using eq. (7.5) and the procedure described in Section 7.1.2.
This correction procedure is somewhat similar to recording the base line in the
absorption spectra measurements. A drawback of this correction is a decrease in
signal-to-noise ratio, since the noise will be present in the measurements without
excitation (Ubg (t)) and will be added to the noise of the measurements with the
excitation.

At a short time scale, approaching sub nanosecond time domain, the correction
procedure is usually not needed since it is possible to find a flat enough part
of the monitoring pulse where the change in the monitoring intensity inside the
measurement window can be neglected.

The pulsed monitoring light also helps to solve the problem of monitoring light
effect on the sample, e. g. see Example 7.2. Between the measurements the sample is
exposed to a minimum possible monitoring light intensity, which is needed to keep
duty arc. For measurement time window the monitoring light intensity increases
hundred times or more, providing high enough photon flux in the monitoring
beam to attain nanosecond time resolution. To block totally the monitoring light
intensity between the measurements a mechanical shutter can be inserted between
the sample and monitoring monochromator.13 The shutter control unit must be
synchronized with the lamp and excitation laser control systems. Typical time needed
to open a mechanical shutter is a few milliseconds, so it must be triggered first, e.
g. 10 ms before the excitation flash. The arc lamp pulse needs a few microseconds
to reach the working intensity level, and must be triggered, e. g. 10 μs prior to the
excitation flash.14

The flash–photolysis systems equipped with shutters and pulsed monitoring light
sources can be used to study irreversible photo-reactions, such as photo degrada-
tion. The instrument is adjusted with a test sample. When everything is ready for
the measurements the sample is switched to the photosensitive one so that the
monitoring light is opened right before the excitation flash and has no effect on the
sample.

7.2.2 Signal averaging


The signal-to-noise ratio of raw measurements, i. e. measurements of U(t) as
described in Section 7.1.2, can be improved by repeating the experiments a few times
and summing up or averaging the results. For a random noise (which is usually the
case) the signal-to-noise ratio increases as the square root of the number of averaged
data, as was discussed in Section 4.1.3. For example, averaging 4 measurements, one
can improve signal-to-noise ratio twice, averaging 100 measurement will reduce the
noise level 10 times, and so on.

One limiting factor in improving quality of the measurements by averaging is the


time needed to repeat the measurements. Naturally, 100 measurements will take
100 folds longer time then a single measurement. The second limit comes from the
systematic errors, e. g. linearity of the photomultiplier response, – the lower noise
level does not necessary mean higher data quality. Repeating measurements many
times one also has to take care about photo stability of the sample and stability of
the instruments as whole.

Planning the experiments one should realize that averaging does not make any tricks
in sense of eq. (7.11) and the following discussion. Repeating the measurements
N times one increases the number of photons (i. e. in eq. (7.11)) N times, thus
decreases 2 N times, i. e. improves signal-to-noise ratio times. In other words,
eq. (7.11) can be interpreted as the relation between the signal-to-noise ratio and
the number of monitoring photons, , in time interval equal to the instrument
time resolution, . The required number of photons can be collected in a single
measurement or repeating the measurement a few times. Therefore a stronger
intensity of the monitoring beam, when possible, can be a better solution to improve
the measurement results than the averaging.

7.2.3 Spectrum range and spectrum resolution


If only the usable wavelength range is considered, the flash–photolysis method
is similar to the steady state spectrum measurements discussed in Section 5.3.1.
One needs a source of monitoring light and a photo-detector, which are principal
components determining the wavelength range.

The spectrum resolution depends on the time resolution at least in nanosecond time
domain. First of all at constant spectrum density of the monitoring light source
higher spectrum resolution means lower total intensity of the monitoring beam.
Second, the monochromator spectrum resolution is proportional to the size of the
slits, i. e. higher resolution means smaller slits, thus reducing the amount of light
which can be passed into the monochromator at higher resolution. If both these
factors are efficient, one can expect the monitoring light intensity to be propor-
tional to the square of the wavelength resolution, , and, thus, signal-to-noise ratio
to be directly proportional to the wavelength resolution, −1 Δ m, in otherwise
equal conditions (see eq. (7.11)). Similarly, the time resolution is proportional to the
square of the wavelength resolution, .

> Read full chapter


Pump-probe
NIKOLAI V. TKACHENKO, in Optical Spectroscopy, 2006

11.5 Sensitivity
The sensitivity of the methods (both mono- and two-colors) can be rather high. It
depends on inaccuracy of the signal determination, ΔS (Δt). Ideally, calculations
of the signal as the ratio of intensities using eq, (11.5) should make the result,
S(Δt), insensitive to the pulse-to-pulse variation in the energies. The limiting factors
are different types of noises, e. g. quantum noise given by the number of the
detected photons and the photo-detector thermal noise. To improve the accuracy,
one can collect the signal during a relatively long time (thus reducing quantum
noise) or apply modulation–synchronous detection technique in a way similar to
one discussed with the steady state absorption measurements. These all allow one
to achieve sensitivity better than 10−4 (in absorbance units).

A big advantage of the pump-probe method (in comparison to flash-photolysis) is


that one does not even need to resolve individual probe pulses: only the average
pulse energy at a certain position of the delay line is important for the accurate
measurements. Also the sensitivity of the pump–probe does not depend on the time
resolution, since the number of photons collected to measure the light intensity after
the sample does not depend on the widths of the pump or probe pulses.

> Read full chapter

Complementary metal-oxide-semicon-
ductor (CMOS) sensors for fluorescence
lifetime imaging (FLIM)
R.K. Henderson, ... D.-U. Li, in High Performance Silicon Imaging, 2014

Wide-field FLIM systems


In conventional wide-field FLIM systems, photons are captured on a CCD-based
camera. This technology has limitations, including the fact that the capture of
photons on an individual pixel results in an accumulation of charge which must be
amplified. This leads to a background ‘readout noise’ that, in addition to stochastic
‘dark noise’ (which is reduced by the cooling of the device) and the ‘shot noise’
(quantum nature of light), results in a significant threshold below which photons
cannot be detected with confidence. There are different types of CCD devices.
Microchannel plate (MCP) based CCDs require a high voltage, usually 600–900 volts,
applied to the MCP. EMCCDs contain several gain registers between the end of the
shift register and the output amplifier. For EMCCDs, any dark current remaining will
be multiplied up along the readout path. That is why EMCCDs must be cooled down
to −100 ° C, significantly increasing their cost (~£25–30K). CCD-based systems
mainly suffer from readout noise generated by the electronic readout circuitry, which
includes charge transfer noise, amplifier noise and quantisation noise emerging
from digitising signals. In low light situations, readout noise (especially for EMCCDs)
can deteriorate the signal-to-noise ratio (SNR) significantly. Some pixel binning
techniques are provided in some commercial devices to reduce, but do not eliminate,
the readout noise. Moreover, they result in poor spatial resolution. Additionally,
ICCDs and EMCCDs, the microchannel plate (MCP) and the electron multiplying
(EM) register suffer from significant ageing effects.

> Read full chapter

Steady State Absorption Spectroscopy


NIKOLAI V. TKACHENKO, in Optical Spectroscopy, 2006

5.3.3 Sensitivity and absorption range


In this section we will discuss the range of possible absorption measurements.
There are few parameters use to specify light absorption by a sample (see discussion
in Section 1.1.1). They are all equivalent and can be re-calculated one to another.
Therefore the absorbance will be mainly discussed here since it is one of the most
widely used and usually presented to the user by default.

At first let us examen the lower limit of the detectable absorbance. Apparently,
there are many parameters essential for the detection of a very low absorbances,
e. g. photometric stability, reproducibility and noise. To present a common view on
the problem let us discuss it from the point of the smallest detectable signal ΔU
and corresponding smallest detectable absorbance, ΔA, which will be here called
sensitivity.

When the light intensity values Iin and Iout (see Fig. 5.1) are close to each other one
can use approximation

(5.6)
where ΔIs = Iin – Iout. It is clear that the sensitivity (ΔA) depends on how accurately
one can measure the light intensities Iin and Iout. If inaccuracy in the intensity mea-
surements is ΔI, then the sensitivity is . Therefore, the sensitivity, ΔA, is completely
determined by the relative accuracy of the light intensity measurements. Apparently
, where ΔU is the minimum resolved voltage deviation on the photo-detector
output. ΔU depends on many factors such as stability of the light source (which
includes stability of the lamp power supply and the quantum noise of the light), the
noise of the photo-detector and the accuracy of the meter following the detector
(synchronous detection in the case of the scheme in Figs. 5.3 and 5.4). For a clever
designed electronics (detector and meter) the ratio can be smaller than 0.00002.7
If , then the sensitivity of the instrument is ΔA 0.0001. This is a typical value
for a high quality general purpose instruments. Already the value ΔA 0.0001
needs an advanced power supply for the light source and a heavy design of the
mechanical framework of the instrument to achieve proper mechanical stability of
the components in the optical part.

The light source quantum noise can be the limiting factor for the sensitivity at low
intensities of the monitoring light. From the viewpoint of quantum noise statistics,
to provide inaccuracy for the intensity measurements the number of (detected)
photons should be grater than (see square root law Section 4.1.2). Assuming the
photo-detector efficiency to be d and the detection time interval Δt, the minimum
light intensity is , where hv is the photon energy. For example, if , as was considered
above, d = 0.1, which is typical for PMTs at the wavelengths of their maximum
efficiency, Δt = 0.01 s, meaning 100 measurements per second, and hv ≈ 4 × 10−19
J (wavelength 500 nm), then the minimum intensity is I ≈ 2μW. This value looks to be
rather low, but is it so in reality? Typical a tungsten halogen lamp emits 1–5 mW in
1 nm bandwidth in the visible spectrum range (see examples in Section 1.2.1). This
power is emitted in all directions and only few percents can be collected (by mirror
M1 in Figs. 5.3 and 5.4) and focused onto the input slit of the monochromator. Part
of the light will be lost in the monochromator (slits, diffraction grating and mirrors).
After all one can expect to obtain a few micro Watts in 1 nm bandwidth at the sample.
As can be seen 1 nm wavelength resolution is close to the limit when the light
intensity may affect the instrument sensitivity.8 Therefore most spectrophotometers
have highest spectrum resolution colse to 1 nm. The best instruments can provide
the resolution close to 0.1 nm without essiential loss of sensitivity.

The sensitivity estimation has been made for a very transparent sample. The other
extreme is a highly absorbing sample. Then Iout is close to zero but one bit greater
than zero, i. e. ΔU corresponds now to Iout. Thus sample transmittance is . Conse-
quently, A 4.7. This is the maximum optical density which could be measured with
an ideal instrument having detection steps of .9
One may notice, that the increase in the light intensity resolution, i. e. decrease in
ΔU and ΔI, respectively, extends maximum measurable value of A and decreases the
minimum resolved value, ΔA, since they both depends on . Of course, this is true if
there are no other limiting factors, such as thermal stability of the instrument base
line.

> Read full chapter

Laser Architectures for On-chip Infor-


mation Technologies
Charles Cornet, ... Cédric Robert, in Integrated Lasers on Silicon, 2016

5.1.2 Optical computing


Moving from on-chip optical routing toward optical processing is the long-term goal
of the photonics community [WOO 12]. If electronics is the medium of choice for
digital computing, photonics shows appealing advantages such as coherence, large
nonlinearities, spinor or analogic character etc. for the development of alternative
computation paradigms like quantum [LAD 10] or neuromorphic computing [LAR
12]. “More-Moore” microprocessor performances could indeed be improved not
only by hybridizing microprocessors at the hardware level (i.e. with photonic and
electronic layers) but also at the computational level, by allocating a given task to
the best suited computation paradigm [ITR 13]. Many building blocks for on-chip
optical computing have been proposed, among them passive [MAS 15] and active
devices [LAD 10, BEN 11b], but once again, research for specific integrated light
sources is a priority. We can cite in particular quantum-noise-limited sources for
continuous variable entanglement [MAS 15], entangled or single photon sources,
excitable lasers operating as spiking neurons for neuromorphic computing [COO
11].

As for the design of optical networks on chip, light sources for optical computing
should indeed meet precise specifications. As an example, let us sketch the proposal
of Coomans et al. on optical spiking neurons out [COO 11]. The cornerstone of many
optical neuron networks is a network node, which provides a nonlinear response to
stimulation by one or several optical inputs. This response has to be strong enough
to act itself as a stimulus for the next node (so-called cascadability). To address this
challenge, the authors of [COO 11] proposed the development of integrated laser
sources operating in a regime close to bistability, the so-called excitability regime.
The basic scheme of their device is presented in Figure 5.3. The optical neuron
consists in an electrically pumped on-chip microring semiconductor laser with chiral
symmetry. It is coupled to either one or several optical waveguides for the injection
of the stimuli and ejection of the response. Due to the particular symmetry of the
device, stable lasing operation occurs in one direction only, defined by the sign
of the chirality. However, if an optical pulse in the opposite direction enters the
cavity through one of the waveguide couplers, the ring laser can switch into the
counterpropagating state for a limited time, typically a few tens of nanoseconds
before coming back to its stable state. Such an excitability regime only exists in
a small injection current span (close to a Takens–Bogdanov bifurcation) and for
particular values of parameters such as the backscattering coupling phase and the
linewidth enhancement factor [GEL 09]. The two last parameters can be roughly set
during the design of the optical neuron by a careful choice of the position of the
spiral corner with respect to the waveguide couplers [SPR 90] and the choice of active
materials [UKH 04], for example. Specification drift due to fabrication imperfections
will thus be more likely to require individual and even dynamical adjustment of the
operation parameters of the optical neurons, such as the injection current or the
device temperature.

Figure 5.3. a) Sketch of two optical spiking neurons consisting of chiral microring


lasers operating in an excitable regime; b) a stimulation pulse exciting the neuron
from the left waveguide. Neuron a switches from counterclockwise operation to
clockwise operation for about 10 ns. This sends a stimulation along the upper
waveguide. Information is then cascaded to neuron b.Reprinted with permission
from Coomans et al. [COO 11] Copyright 2011 by the American Physical Society

Optical neuron networks operating as aforementioned will most likely require


thousands of microlasers, but this is not the case of time-domain neuromorphic
networks [LAR 12, PAQ 12]. In the latter case, a laser beam coming from a single
source is modulated by the use of a Mach–Zehnder interferometer and sent into
a feedback loop composed of a delay line and a mixer, which combines the input
data with the output of the delay line and in turn drives the interferometer. The
nonlinear nodes of the network are thus consecutive time spots circulating in
the feedback loop. The processed optical data is detected at the end of the delay
line. Such a seemingly simple architecture has been implemented with standard
telecom hardware to demonstrate both voice recognition and time series prediction
based on a 400-node time network. If these designs will most likely be extremely
application-specific, their performances compared to similar functionalities in the
electronic domain could be extremely competitive. An on-chip demonstrator of such
a system still needs to be demonstrated.

> Read full chapter

CO2 Isotope Lasers and Their Applica-


tions in Tunable Laser Spectroscopy*
Charles Freed, in Tunable Lasers Handbook, 1995

7. SPECTRAL PURITY AND SHORT-TERM STABILITY


The output waveform of a stable, single-frequency CO2 laser far above the threshold
of oscillation may be approximated by an almost perfect sine wave with nearly
constant amplitude and frequency. For a laser operating in an ideal environment,
the spectral purity is measured by a linewidth that is determined by frequency
fluctuations caused by a random walk of the oscillation phase under the influence of
spontaneous emission (quantum) noise. In their fundamental 1958 paper, Schawlow
and Townes predicted [57] that the quantum-phase-noise-limited line profile will be
a Lorentzian with a full width between the half-power points (FWHM), which may
be approximated by:

(14)

where a, h, 0, P0, and Qc denote the population inversion parameter, Planck's


constant, the center frequency, power output, and “cold” cavity Q of the laser,
respectively. In a well-designed small CO2 laser the “cold” cavity Q is given by:
(15)

where L, c, and tr denote the cavity length, velocity of light, and mirror transmission,
respectively (diffraction losses are usually negligible compared to output coupling
loss). In a small CO2 laser with L = 50 cm and tr = 5%, Qc is of the order of 107; thus for
a typical power output of 1 to 10 W (which is easily obtainable with a small TEM00q
mode CO2 laser) the quantum-phase-noise-limited linewidth is less than 10−6Hz.
Note that 10−6Hz represents less than 1 part in 1019 of the output frequency ( 0 3
× 1013 Hz) of a CO2 laser. This inherent spectral purity of CO2 lasers can be explained
as follows: The linewidth Δ is inversely proportional to the product of P0 and and
the combination of high P0 and high Qc can be simultaneously achieved with relative
ease even in a small CO2 laser oscillator. Oscillators in the radio-frequency (rf ) and
microwave domain have either high P0 or high Qc but not both together in a single
device.

Laser stabilities are most frequently measured in the laboratory from the results
of heterodyne experiments with two lasers. Laser stabilities can be determined by
either frequency-domain (Fourier spectrum) or time-domain (Allan variance) analysis
of the beat-note spectra of the laser pairs. To establish the spectral purity we can
heterodyne two CO2 lasers of equal high quality so that the resulting beat-note
spectrum can be apportioned equally to each laser. Two problems arise, however, in
trying to measure the Schawlow–Townes linewidth of high-quality CO2 lasers. The
first of these problems is instrumental: The stability of the available instrumentation
itself generally cannot reliably measure spectral purities of 10−6 Hz or better.

The origin of the second problem is that for well-designed CO2 lasers [56] the
so-called technical noise sources dominate over the quantum-phase-noise-limited
Schawlow-Townes linewidth [57]. Examples of technical noise sources are acoustic
and seismic vibrations, and power-supply ripple and noise. These sources can cause
frequency instabilities by perturbing the effective cavity resonance via the sum of
fractional changes in the refractive index n and the opticalcavity length L:

(16)

As an example, a change of only 10−3 Å(about 1/1000 of the diameter of a hydrogen


atom) in a 50-cm-long CO2 laser cavity will cause a frequency shift of approximately
6 Hz. A 6-Hz variation in the approximately 3 × 1013 Hz frequency of a CO2 laser
corresponds to a fractional instability of 2 × 10−13.

Figure 6 shows the real-time power spectrum of the beat signal between two
free-running lasers that were designed and built at Lincoln Laboratory [56]. The
spectral width of Fig. 6 implies a frequency stability at least as good as 2 × 10−13.
The discrete modulation sidebands in the figure were primarily due to ac-power-line
frequency harmonics, cooling fan noise, and slow frequency drift; however, each
spectral line was generally within the 10-Hz resolution bandwidth of the spectrum
analyzer. The measurement of the spectral width was limited to a 10-Hz resolution
by the 0.1-sec observation time that was set by the instrumentation, not by the laser
stability itself.

FIGURE 6. Real-time power spectrum of the beat signal between two free-running
CO2 lasers. The horizontal scale of the figure is 500 Hz/division, which indicates that
the optical frequencies of the two lasers producing the beat note were offset by less
than 3 kHz. The 10-Hz width of the line shown is limited by instrumentation; the
linewidth of the beat note falls within this limit.

Figure 7 shows the real-time power spectrum of the beat note of two ultrastable CO2
lasers that were phase-locked with a fixed 10-MHz frequency offset between the two
lasers and with the unity-gain bandwidth of the servoamplifier set to about 1.2 kHz
[56]. Note that the horizontal scale in the figure is only 2 × 10−2 Hz/division and the
vertical scale is logarithmic, with 12.5 dB/division. Using the results from Fig. 7 and
the equation for a Lorentzian lineshape, we calculate the FWHM spectral width of
the beat note to be about 9 × 10−6 Hz.
FIGURE 7. Spectral purity of beat note between two CO2 lasers that were
phase-locked to each other with a frequency offset of 10 MHz. The FWHM spectral
width of the beat note was about 9 × 10−6 Hz during the 26.67-min measurement
time.

It took 26.67 min of measurement time to obtain just a single scan with the
frequency resolution of Fig. 7. Because tracking even by a very good servosystem
would still be limited by quantum phase noise, the narrow linewidth in Fig. 7 is an
indirect but clear confirmation of the high spectral purity of CO2 lasers, as predicted
by the Schawlow-Townes formula.

The (so far at least) unsurpassed spectral purity and short-term stabilities measured
in the frequency domain and illustrated in Figs. 6 and 7 were also confirmed
by analyzing the signal returns from orbiting satellites that were obtained by a
long-range CO2 radar at the Firepond facility of MIT Lincoln Laboratory [56, 58–62].
Additional confirmation was also obtained at MIT Lincoln Laboratory from extensive
time-domain frequency stability measurements on pairs of ultrastable CO2 lasers
under free-running and phase-locked conditions, and both in acoustically quiet and
in noisy environments [63].

> Read full chapter

Methods for Studying Elementary Reac-


tions
E.T. Denisov, ... G.I. Likhtenshtein, in Chemical Kinetics: Fundamentals and New
Developments, 2003
3.2.4 Intracavity laser spectroscopy (ICLS)
The method is based on the influence of the studied substance on the parameters
of laser radiation. The method is based on that the reactor with the gas is placed
into the laser cavity with a broad amplification contour as it is shown in Fig. 3.2.
The main thing is to select parameters of the active medium of the laser so that
the amplification of the light intensity in it would compensate losses on mirrors and
would not compensate losses related to the studied absorption. These losses differ in
frequency dependence. (The losses on mirrors are broad-band compared to narrow
absorption lines of detected gas molecules.)

Fig. 3.2. Scheme of the kinetic intracavity laser spectroscopy technique: 1, active
medium of the laser; 2, reactor; 3, mirrors; 4, pulse lamps for photolysis of a mixture
and pumping of the laser; 5, feeding source for photolysis; 6, feeding source for the
laser; 7, controlling unit; 8, unit for time delay between photolyzing and laser pulses;
and 9, detector of the spectrum.

For easier understanding of the main principles of the ICLS, let us use the model of
a multimode laser described by the system of balance equations for the number of
photons in each mode of the cavity Mq and population inversion N. Let us assume
for simplicity that the contour of the amplification line in the laser is uniformly
broadened and the interaction of the modes and quantum noises are absent.

Here q is the number of the mode; b are broad-band losses in the cavity, which are
the same for all modes; aq is the absorption coefficient of the intracavity absorption;
P is the pumping power; t is the lifetime of the upper laser level; Bq is the Einstein
coefficient; c is the speed of light; and N is the total number of generating modes.

First let us consider the case where the absorption coefficient is the same for all
modes, Bq = B, the narrow-band absorption is absent, i.e., aq = 0, the approximation
t 1/b is valid. In this case, stationary solutions of the equations written above have
the following form:
Here Pth = b/Bt is the threshold pumping power, and h = P/Pth is the increase of the
pumping over the threshold.

Now let us consider the case where the pumping power aq differs from zero only for
one mode of the resonator qo. In this case, the values of the threshold and population
inversion are determined by modes with the lowest modes and, hence, have the
same value as in the previous case. After switching-on pumping, if the aq value is not
very high, generation appear on all modes. After stationary values of the inversion
No and intensity of the generation Mo are achieved, the intensity in the mode qo
decreases by the exponential law according to Eq. (1) in which the first two terms in
the right part are mutually compensated

It is assumed in the derivation of this formula that a decrease in the photon density
at the frequency wo has almost no effect on the concentration of active particles
N and on the amplification coefficient at the same frequency. It can be shown
that a change in the amplification coefficient at this frequency is proportional to
exp(-2pg/DwP), where g is the width of the uniform amplification line, and DwP is the
absorption linewidth of the analyzed substance. Therefore, we may conclude that the
necessary condition for the case where the derived formula for the time evolution
of the number of photons with the frequency wo is fulfilled can be formulated as
follows:

For gases DwP is usually lower than 1 cm−1, whereas 2pg for condensed active media
lies in the interval from 10 to 103 cm−1. Thus, the ICLS method can successfully be
used for many active media, liquid organic dyes (380—1000 nm), crystalline active
media (1—3 mm).

The number of photons in each mode of the cavity is proportional to the spectral
intensity I(wo). Therefore, we can write

This formula is valid if the whole cavity is filled with an absorbing substance. In many
cases, only a part of the cavity is filled with the substance. When the cavity length is
designated through L and 1 is the length of the absorbing layer, then the coefficient
of cavity filling should be introduced into formula. Then we obtain the final formula

The absorption coefficient ao(wo) (wo is the frequency at which the substance under
study absorbs) at which the generation intensity in the chosen mode decreases by e
times is determined by the duration of the generation tg

Thus, the intracavity spectrometer corresponds to the traditional absorption spec-


trometer with a cell having the equivalent length

(3.9)
It is seen from this formula that the sensitivity depends on the generation time of
the laser. The ICLS spectrum is shown in Fig. 3.7. These spectra are processed in
the same manner as in absorption laser spectroscopy, due to which the value y =
a(wo)(1/L)ct is determined.

Fig. 3.7. General scheme of the technique for studying the dynamics of the transition
state: 1, generator of femtosecond pulses; 2, optical light amplifier; 3, optical line
of the time delay between the pump and probe pulses; 4, devices for changing the
wavelength of light pulses; 5, reactor; and 6, system of detection of fluorescence or
ion current induced by probe pulses.

When organic dyes are used as active media, it has experimentally been shown that
the effective length is proportional to the generation time up to tg is equal to 0.1 s.
At long generation times of the laser, deviations from this law begin. When the rate
of absorption by the sample is comparable with the rate of spontaneous radiation,
an increase in the generation duration already gives no gain. From this condition
we can estimate that the minimum value of the absorption coefficient, which can
be measured by this method, is 3·10−11 cm−1.

The theory shows that expressions (3.8) and (3.9) are fulfilled to times of 1 s. If
we accept that the detected gas fills the whole laser resonator, at a time of laser
generation of 1 s the optical path length is 300 000 km.

Note that the great length of the optical path can be realized at small geometrical
sizes of the reaction vessel. At the same time, the brightness of the light source
remains great, and spectrographs with a high spectral resolution can be applied.
The sensitivity of the ICLS method to selective absorption is restricted by the pres-
ence of spontaneous noises. The theory and experiment show that the absorption
coefficient s[n] can reach 10−11 sm−1.

Laser with the following active media satisfy the necessary conditions of ICLS:
glass activated by neodymium; solutions of organic dyes; alkali-halide crystals with
coloring centers; crystals of the Ti type; sapphire, etc. Lasers operating in both pulse
and continuous modes are used. Lamp pumping, which provides a long generation
time, is often used for the work in the pulse mode. Continuous generation of dye
lasers is performed using ionic argon and krypton lasers for pumping. The typical
scheme of the technique is presented in Fig. 3.2. Most frequently the ICLS method
is applied for studies in a static reactor in combination with pulse photolysis. The
characteristics of ICLS are given in Table 3.2.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like