Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Tunable diode laser absorption spectroscopy for measurements in a gas

turbine combustion environment

Pavel Kolos
31 august 2023

Nomenclature

IR-LAS Infrared laser-absorption Ni Absorbing-species column density


spectroscopy n Number density
A Integrated absorbance P Gas pressure
a Modulation depth PLIF Planar laser-induced fluorescence
c Speed of light Q Internal partition function
DA Direct absorption R Two-color ratio of integrated
absorbance
DFB Distributed-feedback RMS Root mean square
E′′ Lower-state energy S Transition linestrength
f Modulation frequency SLM Standard liter per minute
FDML Fourier-domain mode-locked SNR Signal-to-noise ratio
FFT Fast Fourier transformation SWDA Scanned-wavelengths direct absorption
h Planck’s constant T Temperature
H Harmonic component of TARS Triple annular research swirler
transmission coefficient Fourier
series
HPST High-pressure shock tube TDL Tunable diode laser
HT Hyperspectral tomography Tni Absorbing-species number-density-
weighted path average temperature
𝐼 Incident light intensity To Reference temperature
𝐼0 Average laser intensity WMS Wavelength modulation spectroscopy
i2 First term of nonlinear laser α Absorbance
intensity modulation amplitude
io Linear laser intensity modulation ν Optical frequency
amplitude
𝐼 Transmitted light intensity 𝜈̅ Center laser frequency
k Boltzmann constant νo Linecenter frequency
L Path length through absorbing gas φ Equivalence ratio
LAT Laser absorption tomography ϕ Lineshape function
LBO Lean blowout χ Mole fraction of absorbing species
LOS Line of sight 𝜏 Transmission coefficient

1
1 Introduction

Over the last 40 years, a wide range of laser diagnostics has been developed and deployed for
characterizing combustion via gas temperature, pressure, velocity, and composition measurements.
Implementing the advanced sensors, working based on laser diagnostic, into a jet engine control
system will expand the engine's possibility to maintain high efficiency and low emissions over a wide
range of operational regimes. Modern aviation engines don’t have the equipment for real-time
monitoring of combustion temperature, equivalence ratio, and emission species concentration. The
environment in gas turbines is too harsh for implementing conventional sensors.
One of the modern techniques, that are developing today, is infrared laser-absorption
spectroscopy (IR-LAS). Modern researchers are putting efforts into creating sensors applied for
measurements in gas turbines and based on IR-LAS techniques. However, for many combustion
applications, capturing the spatially resolved structures in the flow field is needed, and planar laser-
induced fluorescence (PLIF) and other pulsed imaging diagnostics are naturally preferred. Similarly,
coherent anti-Stokes Raman scattering remains the preferred technique for providing point
measurements of temperature and species.
IR-LAS techniques offer cost advantages over PLIF due to the affordability of diode, quantum-
cascade, and interband-cascade lasers used, as well as lower operational costs for laser detectors
compared to anti-Stokes Raman detectors and PLIF cameras [1]. Additionally, IR-LAS lasers are more
compact, robust, and energy-efficient than current laser-based imaging systems. IR-LAS can provide
high-bandwidth, species-specific measurements of thermodynamic conditions (e.g., temperature,
pressure) and a wide range of molecular species, occurring in gas turbine thermodynamic cycles, such
as H2O, CO2, CO, CH4, and NOx. All mentioned advantages make the IR-LAS techniques most
preferable for gas turbine environments.

2 Measurement Techniques

Direct Laser Absorption Techniques

The most common laser absorption technique (LAS) is scanned-wavelengths direct absorption
(SWDA). In SWDA, the laser wavelength is tuned over the majority of an absorption feature to
measure the integrated area and lineshape directly. The Fundamentals of the IR-LAS modeling take
the following forms [2].
The absorption of photons at optical frequency 𝜈 can be related to gas properties using the
Beer-Lambert relation given by Eq. 1:

2
𝐼
= exp (−𝛼 ) (1)
𝐼
The absorbance due to a single quantum transition can be related to gas properties using
Eq. 2:

𝛼 = 𝑆 (𝑇) 𝑛 𝜙 (𝜐, 𝑇, 𝑃, 𝜒)𝑑𝑙 (2)

In combustion gases, the lineshape function is usually modeled by the Voigt Profile (VP)
owing to it accounting for both Doppler and collisional broadening. However, in certain cases the VP
can lead to substantial errors, thereby motivating the use of more advanced lineshape models. The
linestrength is given by Eq. 3:
𝑄(𝑇 ) ℎ𝑐𝐸 1 1
𝑆 (𝑇) = 𝑆 (𝑇 ) 𝑒𝑥𝑝 − −
𝑄(𝑇) 𝑘 𝑇 𝑇
(3)
ℎ𝑐𝜈 ℎ𝑐𝜈
× 1 − 𝑒𝑥𝑝 − 1 − 𝑒𝑥𝑝 −
𝑘𝑇 𝑘𝑇
The lineshape function is defined such that its integral over optical frequency is unity Eq. 4:

𝜙𝑑𝜈 = 1 (4)

As a result, the integrated absorbance (i.e., area), A cm −1, is given by Eq. 5. This integrated
absorbance can be measured by laser absorption technique:

𝐴≡ 𝛼 𝑑𝜈 = 𝑆 (𝑇)𝑛 𝑑𝑙 (5)

Which reduces to Eq. 6 for uniform gas conditions over the line-of-sight (LOS):
𝐴 = 𝑆 (𝑇)𝑛 𝐿 (6)
The temperature can be inferred from the ratio of integrated areas of two transitions with
different lower-state energies. If the gas is uniform along the optical path, this ratio, R 2λ, A, reduces to
the two-color ratio of transition linestrengths and is given by Eq. 6:
𝐴 𝑆 , (𝑇)𝑛 𝐿 𝑆 , (𝑇)
𝑅 , ≡ = = (7)
𝐴 𝑆 , (𝑇)𝑛 𝐿 𝑆 , (𝑇)
If the path length and gas temperature are known, the absorbing-species number density can
be found from the integrated area of a single transition using Eq. 6. If the gas pressure is also known
and an appropriate equation of state exists, the absorbing-species mole fraction can also be found from
the area. A schematic illustration of SWDA is given in Figure 1.[1]

3
Figure 1 – Schematics for scanned-wavelength direct absorption [1]

Hyperspectral direct absorption

Strictly speaking, SWDA is typically used to describe cases where the laser’s
wavelength/frequency is tuned over a narrow spectral range (0.1 to 10 cm -1) to resolve one or several
absorption transitions. Hyperspectral direct absorption (DA) spectroscopy is used for spanning a
region of the spectrum greater than 10 cm-1. All hyperspectral DA techniques are used to measure a
large portion of an absorption band, frequently at very high rates (10 to 100 kHz). The linewidth of
these sources varies greatly, from the order of kHz to near-GHz (e.g., Fourier-domain mode-locked
(FDML) sources).
This unique combination of features has mandated the use of a variety of unique data
processing and spectral-fitting techniques to overcome challenges associated with large datasets and
the need to accurately model significant portions of an absorption band, at times, convoluted with an
estimated instrument function. Despite these complexities, hyperspectral DA techniques have
demonstrated the ability to overcome most of the challenges limiting conventional SWDA, and as a
result, have emerged as one of the most promising LAS technologies for characterizing harsh
combustion environments.

Wavelength-modulation spectroscopy

Wavelength modulation spectroscopy (WMS), as an extension of absorption spectroscopy, is


a well-known technique for improving the signal-to-noise ratio (SNR). WMS has been used
extensively to study a wide range of harsh combustion environments particularly due to several
experimental and theoretical advancements enabling robust, calibration-free WMS techniques.

4
Typically, the first and second harmonic signals are used to calculate gas properties. The Fundamentals
of the WMS technique take the following forms [3], [4].
The laser wavelength is rapidly modulated as Eq. 8:
𝜈(𝑡) = 𝜈̅ + 𝑎 × cos (𝜔𝑡) (8)
The incident laser intensity is modeled by Eq. 9:
𝐼 (𝑡) = 𝐼 ̅ [1 + 𝑖 cos(𝜔𝑡 + 𝜓 ) + 𝑖 cos(2𝜔𝑡 + 𝜓 )] (9)
The transmission coefficient can be expanded in a Fourier cosine series by Eq. 10:

𝜏 𝜈(𝑡) = 𝐻 (𝜈̅ , 𝑎)cos (𝑘𝜔𝑡) (10)

The second harmonic Fourier component is given by Eq. 11:


𝑆(𝑇)𝑃 𝐿
𝐻 (𝜈̅ , 𝑎) = − 𝜙(𝜈̅
𝜋 (11)
+ 𝑎 × cos (𝜔𝑡)) × cos(2𝜔𝑡) 𝑑(𝜔𝑡)
Lock-in amplifiers are used to measure the second and first harmonics (1f and 2f) for WMS
detection. This technique is sensitive to absorption line shape curvature and is insensitive to low-
frequency noise. Assuming an optically thin (absorbance < 0.05) line-center measurement of an
isolated absorption transition and linear intensity modulation with a phase shift of π, the 1f normalized
2f signal simplifies to Eq. 12:
2𝑓 𝐻 𝑆(𝑇)𝑃𝜒 𝐿
≈ = 𝜙(𝜈̅
1𝑓 𝑖 𝑖 𝜋 (12)
+ 𝑎 × cos (𝜔𝑡)) × cos(2𝜔𝑡) 𝑑(𝜔𝑡)
Taking the ratio of the 2f/1f signals from two different absorption features cancels the direct
dependence on species concentration and relies only on temperature. In the following Eq. 13, the ratio
of i0,2/i0,1 terms is a constant, the ratio of line strengths depends solely on T, and the ratio of lineshape
integrals depends weakly on T and on χi and on P for cases where the Voigt or Lorentzian profiles are
used:
𝑖 , 𝑆(𝑇) ∫ 𝜙 (𝜈̅ + 𝑎 × cos (𝜔𝑡)) × cos(2𝜔𝑡) 𝑑(𝜔𝑡)
𝑅𝑎𝑡𝑖𝑜 = (13)
𝑖 , 𝑆(𝑇) ∫ 𝜙 (𝜈̅ + 𝑎 × cos (𝜔𝑡)) × cos(2𝜔𝑡) 𝑑(𝜔𝑡)
A schematic illustration of WMS is given in Figure 2.

5
Figure 2 – Schematics for wavelength modulation spectroscopy [4]

Solutions for nonuniform flows

Laser-absorption spectroscopy provides a path-integrated measurement. In cases where the


gas conditions vary along the measurement line-of-sight (LOS), the measured integrated absorbance
can’t be given by Eq. 6 and is given only by Eq. 5. The path integral in Eq. 5 complicates the
relationship between the measured absorbance and gas properties. However, several techniques were
developed to provide accurate measurements of path-averaged temperature and concentration in a
non-uniform environment, for instance, the two-color absorption spectroscopy strategy [2].
It was shown that by using two absorption transitions with strengths that scale linearly with
temperature over the domain of the temperature nonuniformity, absorbing-species column density (Ni)
and absorbing-species number-density-weighted path average (Tni) can be calculated from the
integrated areas of two transitions. Furthermore, the absorbance spectra observed across a nonuniform
LOS can be accurately compared with simulations performed with a uniform LOS, effective
lineshapes, and absorbing-species number-density-weighted path-average gas conditions
(temperature, pressure, and absorbing-species column density), the example of spectra is shown in
Figure 3. As a result, measured SWDA and WMS signals can also be directly translated to the
absorbing-species number-density-weighted path average gas conditions without knowing how the
gas conditions vary along the LOS.

6
Figure 3 – Simulated WMS-2f Ú 1f spectra for two water-vapor transitions [2]

Hyperspectral tomography

A wide range of laser absorption tomography (LAT) techniques have been developed and used
to reconstruct 2D chemical species, temperature, and pressure fields in combustion gases [1]. In
tomography, multiple LOS is used to sample a heterogeneous measurement volume from unique
viewpoints, and the measured projections (i.e., path-integrated absorption) are used to reconstruct the
2D (or even 3D) field. In most LAT applications, the 2D field is reconstructed from measured
projections by solving linear systems of equations, Ax=b; where the m x n matrix A specifies the
known path length of each ray through each pixel, the m-vector b contains the measured projections
for each LOS, and the n-vector x specifies the unknown (i.e., to-be-solved for) absorbance or species-
concentration in each pixel. A multitude of algorithms have been developed to solve for x (e.g., in a
least-squares sense or through minimization of an objective quantity) and address the ill-posed nature
of this problem (i.e., multiple solutions often exist or the solution is unstable and sensitive to input
noise).
Recently, hyperspectral tomography (HT) techniques have been developed and deployed to
reconstruct 2D temperature and H2O concentration fields [5]. This method utilizes many absorption
transitions, thereby adding increased spectral information to the reconstruction process. In practice,
the temperature and species-concentration fields are reconstructed by minimizing the function given
by Eq. 14:

𝑝 𝐿 ,𝜆 − 𝑝 𝐿 ,𝜆
𝐷(𝑇 ,𝜒 )= (14)
𝑝 𝐿 ,𝜆
where pm is the measured projection at location Lj and a wavelength λi, pc is the corresponding
projection calculated using the reconstructed T and 𝑐ℎ𝑖𝑛 𝑝𝑟𝑜𝑓𝑖𝑙𝑒, and J and I are the total number of
wavelengths and projection locations, respectively.

7
Broad and Blended spectra

As pressure increases, the corresponding increase in collisional broadening leads to broad,


overlapping transitions. Broad and blended spectra complicate quantitative absorption measurements
for three primary reasons: (1) The absorbing baseline complicates the determination of the incident
laser intensity, (2) overlapping transitions complicate the relationship between the absorbance
spectrum and gas conditions, and (3) the broad spectrum leads to smaller WMS-2f signals.
The first complication only pertains to DA techniques, and this challenge has been overcome
by using an additional non-resonant light source or broadly tunable light sources.
In high-pressure gases, the absorbance at a given wavelength is commonly influenced by many
overlapping transitions that complicate the analysis. As a result, to address the second complication
and fully understand how the WMS signals vary over a broad range of temperatures and pressures, a
complete and accurate WMS model is needed.
The last complication pertains to WMS, and this challenge has been overcome by using large
modulation depths and, most recently, by combining large modulation depths with strong mid-infrared
absorption.

3 Applications

Equivalence ratio measurements in gas turbine combustors

Nowadays, modern research in gas turbine measurements develops techniques for producing
measurements inside combustion chambers. One of the reasons is to reduce pollutant emissions by
maintaining lower flame temperatures, significantly reducing NO x. It can be reached by operating at
lean air/fuel ratios (equivalence ratio). In addition, these lean operating conditions reduce engine
maintenance, because the lower combustion temperatures increase the lifetime of engine components.
However lean premixed can cause combustion instabilities in the form of thermoacoustic oscillations
and lean blowout, which reduce efficiency, increase emissions, and pose significant safety hazards.
To maintain safety and, at the same time, low NOx emissions, the monitoring of the equivalence ratio
has to be implemented.
A rapid (2 kHz) tunable diode laser (TDL) absorption sensor has been developed for real-time
measurements of equivalence ratio in gas turbine combustors [3]. The real-time TDL sensor for
equivalence ratio is based on near-IR absorption of methane around 1.65 μm (2ν 3 band). This
wavelength is free of interference from the absorption of other species in the air. Figure 4 shows the
simulated methane absorption spectrum (with HITRAN 2004 database) for typical lean premixed gas

8
turbine conditions. The R(3) line around 1653.725 nm is well isolated at atmospheric pressure and
thus is selected. The WMS technique is used to provide real-time measurement.

Figure 4 – Simulated absorption spectra for CH4


(1 and 15 atm, 6% CH4 in air, L = 7 cm, 700 K) [3]

Figure 5 shows the schematic of the real-time TDL sensor for equivalence ratio measurements
in gas turbine combustors. The fiber-coupled distributed-feedback (DFB) laser operating near 1653
nm is mounted in a commercial laser mount and maintained at a constant temperature. The laser
wavelength is modulated by a 10 kHz sinusoidal current waveform to generate a modulation depth of
0.08 cm-1. The laser beam is divided into two paths by the fiber splitter. For the first path, the light is
collimated into free space and transmitted through a static cell filled with calibration gas (CH 4 and N2
mixture). For the second path, the light goes through a set of 25 mm sapphire windows which are
mounted flush with the walls on the air plenum and gas turbine combustor. The laser beam is then
focused on the detector. The free space light paths are exposed to air since there is negligible
interference absorption. The detector signal is demodulated by a lock-in amplifier to simultaneously
recover the 1f and 2f signals with a time constant of 0.5 ms. The sensor bandwidth may be improved
by using two lock-in amplifiers or a software lock-in. A fast PC equipped with an NI-DAQ board
records signals and provides real-time equivalence ratio output at 2 kHz.

9
Figure 5 – Schematic of the real-time TDL sensor for equivalence ratio [3]

The TDL sensor is used to measure equivalence ratio fluctuations in a dry low NO x gas turbine
combustor to study the flame characteristics. The atmospheric pressure, swirl-stabilized gas turbine
combustor is fueled with natural gas and is shown schematically in Figure 6. The flow is from left to
right. The natural gas and air flow rates are controlled by valves and measured with calibrated flow
meters. The natural gas is delivered to the combustor's end cover and injected into the airflow through
the holes on the swirler vanes.

Figure 6 – Schematic of the atmospheric pressure test rig with a gas turbine combustor [3]

Combustion tests with steady conditions are conducted to demonstrate the TDL sensor
accuracy and response time. Figure 7 shows the measured WMS-1f and -2f signals and -1f normalized
WMS-2f signal R. The measured ratio is not steady, probably due to unsteady turbulent mixing.
Finally, the measured WMS-2f / lf ratio R as a function of equivalence ratio for fixed air flow rate was
derived and also presented in Figure 7. The bar in the figure represents the measured ratio fluctuations
for different test conditions. The gas conditions are P = 1 atm, T = 297 K, air= 0.177 kg/s,
Eq. Ratio = 0.65.

10
Figure 7 – Measured 1f, 2f, and R (left) and the measured R as a function
of equivalence ratio (right) [3]

Sensing and control of combustion instabilities

As mentioned above, lean premixed combustion can cause thermoacoustic instabilities and
lean blowout (LBO). Intensive experimental and theoretical work has been performed to understand
the driving mechanism of thermoacoustic instabilities and to develop effective approaches to suppress
them. It was shown that typical instabilities can be reduced by passive or active control in industrial
combustors. An important part of any control strategy is a robust sensor to measure a meaningful
control variable.
The single TDL was proposed as a sensor to detect thermoacoustics instability and the
proximity to LBO along an optimized LOS in a swirl-stabilized combustor, which serves as a model
of a gas turbine combustor [6], [7]. The TDL temperature sensor is based on absorption from two
neighboring near-infrared transitions of water vapor near 1.4 μm. The selected line pair is free of
interference from the absorption of other major combustion products. The measurement technique is
WMS. The simulated spectra of the chosen transition pair and WMS-2f peak ratio as a function of
temperature are shown in Figure 1. It should be noted that the use of the TDL sensor for precise
temperature measurement may be complicated by the assumption of uniform gas composition and
temperature along LOS. For the control application temperature changes and fluctuations of the flame
are more important than the absolute values of temperature.

11
Figure 8 – Simulated H2O WMS-2f spectra at 300, 1000, 1500, and 2000K; P=1atm;
10% of H2O in air; L =15cm; modulation depth a = 0,047 cm-1 [6]

The atmospheric pressure, swirl-stabilized dump combustor, used for the experimental study,
is propane-fueled and is shown schematically in Figure 9. The combustor configuration was designed
as a model of a lean, partially premixed turbine combuster that includes a triple annular research
swirler (TARS). The flow is from bottom to top. A combination of honeycomb and mesh screens is
used to create uniform airflow. Identical loudspeakers (75 W each) are mounted in the air-conditioning
chamber for thermoacoustic stability control. The TARS has three air passages and the fuel injection
points are uniformly distributed between the outer and intermediate swirlers. The combustion chamber
is bounded by a quartz tube, which permits the uncooled operation of the combustor and provides
optical access for the TDL sensor. The fuel flow and airflow rates are independently controlled by
valves and measured using calibrated flow meters.

Figure 10 – Schematic diagram of the real-time TDL temperature sensor and the swirl-stabilized
combustor [6]

12
The TDL sensor's details are generally the same as in a previous application example. A DFB,
fiber-coupled diode laser operating near 1.4 μm is driven by external current modulation: a 2-kHz
linear current ramp, yielding a wavelength of approximately 2 cm -1, summed with a 500-kHz
sinusoidal current modulation generating a=0.047 cm -1. The second harmonic component of the
detector signal is measured with a lock-in amplifier.
Figure 11 shows the measured temperature and its fast Fourier transformation (FFT) power
spectrum for a laser LOS near the flame tip with an airflow rate of 820 standard liters per minute
(SLM) and propane flow rate of 39.6 SLM (𝜑 = 1.1). The dominant oscillation mode (232 Hz) and its
harmonic (464 Hz) can be seen from the FFT spectra. This confirms the interpretation of the
temperature data: thermoacoustic instability is the coupling of unsteady heat release to acoustic
oscillations.

Figure 11 – Measured signals and FFT power spectra [6]

A phase-delay feedback control strategy was used for the demonstration of the TDL sensor
implemented into active control. The experimental setup is shown in Figure 12. The 2 kHz real-time
sensor output is the first time delayed, and this delayed signal is filtered and amplified to drive the
loudspeakers to modulate the intake airflow.
In Figure 12, the power spectra of the TDL sensor are also shown. Without control, a 388-Hz
oscillation is seen in the spectrum, whereas with control, the thermoacoustic instability is successfully
suppressed by the phase-delay feedback control. These results indicate that the TDL sensor can
accurately identify the frequency, phase, and amplitude of the flame instability and provide the
appropriate feedback for the active control system. For the LBO indication and control, the same
approach can be used to measure an increase in low-frequency fluctuations to detect the proximity to
LBO.

13
Figure 12 – Experimental setup and measurements of active phase-delay
feedback control [6]

Measurements of temperatures at high pressure (CO 2 measurements)

Many of the current combustion devices (e.g. gas turbines) operate at high pressures and future
engines will be operating at even higher pressures. New diagnostic methods are required for
quantitative and time-resolved measurements of reactants, intermediate species, products, and
temperature of various fuels and fuel blends at elevated pressures. Extending sensor capability to high-
pressure environments is complicated by the broadening and overlap of discrete spectral features at
high gas density.
High-pressure CO2 measurements were carried out to understand the challenges of TDL sensor
design for high-pressure applications [8]. This work is aimed at developing a TDL sensor for high-
pressure measurement of CO2 concentration and temperature near 2.7 μm using 1f-normalized WMS
with second harmonic detection. The two transitions, 3633.08 and 3645.56 cm −1, were used for
detection. The absorbance for these two transitions is simulated for 1% CO 2 at 1000 K for pressures
of 1 and 10 atm, and these results are plotted in Figure 12. Since the WMS-2f signal is sensitive to the
curvature of the line shape and approximates the second derivative of the transition, it is desired to
select lines that have sufficient curvature at high pressures. The transition near 3633.08 cm −1
(E”= 316.77 cm−1) is relatively well isolated and retains its curvature at 10 atm, but the transition near
3645.56 cm−1 (E”=1936.09 cm−1) mixes with the nearby transitions and has a little curvature at
elevated pressures, producing a relatively small WMS-2f signal and, thus, a low SNR at elevated
pressures. Therefore, for this high-pressure work, another nearby transition at 3645.20 cm −1
(E”= 994.19 cm−1) is chosen, also shown in Figure 13, that has more curvature at high pressures, can
14
be accessed by the same laser, and has a larger line strength at the temperatures of interest. Accurate
measurements of fundamental spectroscopic parameters, including line strength, line broadening, and
pressure shift are also essential in the development of a new sensor for combustion applications.

Figure 13 – Simulated absorbance for the R(28) CO2 transition near 3633.08 cm−1, the R(50) CO2
transition near 3645.20 cm−1, and the P(70) transition near 3645.56 cm −1. T=1000 K, L=14.13 cm,
and XCO2=1% in Air [8]

With the knowledge of the wavelength shift due to modulation, WMS-2f/1f spectra are
measured for the two CO2 transitions and compared with Voigt simulations at 5 and 10 atm in
Figure 14. The agreement between the measurements and simulations is excellent for the entire 2f line
shape. These measurements also validate the accuracy of this new WMS-2f/1f sensor for
measurements of CO2 concentration and temperature at high pressures.

Figure 14 – 1f-normalized WMS-2f magnitude for the two CO2 transitions. f = 100 kHz,
T = 296 K, 2% CO2 in air, 3633.08 cm−1 transition, L = 6 cm, a = 0.201 cm−1; 3645:20 cm−1
transition: L = 100 cm, a = 0.115 cm−1. (left) P = 5 atm, (right) P = 10 atm [8]

These experiments are carried out in a stainless steel shock tube with a diameter of ∼15 cm,
where argon is the main bath gas and the CO2 mole fraction ranges from a few hundred parts per
million to a few percent. The temperature range was 800–1200 K, with pressures of 10 atm. Shock
15
tubes can provide well-defined temperatures and pressures for kinetic investigations that cover broad
engineering and scientific interest regimes. The shock tube-driven section is 8.54 m in length,
separated from the driver section by a polycarbonate diaphragm. A simple schematic and the cross-
section of the tube are shown in Figure 15, along with the optical and electronics setup. The light from
each laser is collimated and transmitted through optical windows on the shock tube side wall. The
optical arrangement is based on the assumption that the gas properties across the shock tube are
uniform, as expected for heating by a planar shock. Measurements are carried out at 2 cm from the
end wall. The two tunable DFB diode lasers were used. These lasers are modulated by a sinusoidal
injection current at 100 kHz, with a modulation depth of 0.201 and 0.115 cm −1, respectively.

Figure 15 – Experimental setup for the high-pressure shock-tube measurements [8]

A sample measurement is shown in Figure 16 which plots the time history of the gas
temperature measured by the WMS-2f/1f sensor, and the transducer-measured pressure trace. The
average measured temperature over the time interval of 0-1 ms is 820.4 K ± 8.6 K, which is in excellent
agreement with the value calculated from the ideal shock equations (T5 = 826 K). Similar tests are
performed at different temperatures and a pressure range of P5 ∼ 8–12 atm. The results are also plotted
in Figure 16, where the top panel compares the sensor-measured temperature (over 0-1 ms time
interval) with that calculated from ideal shock relations (T5), and the bottom panel compares the
sensor-measured CO2 mole fraction with known mixture values. The measured and calculated
temperatures are in excellent agreement (∼1%) over the tested temperature range of 800–1200 K and
the measured mole fraction agrees with the mixture values within 1.5%. These results validate the

16
sensor accuracy for temperature and CO2 concentration measurements at high pressures and high
temperatures.

Figure 16 – Measurements of 2% CO2 in the air, L = 14.13 cm. (left) Measured temperature and
pressure trace for a shock arriving at t = 0, reflected-shock conditions T5 = 826 K, P5 = 8.7 atm.
(right) Temperature and CO2 concentration measurements by the fixed-wavelength high-pressure
WMS-2f/1f sensor, P ∼ 8–12 atm [8]

Measurements of temperatures at high pressure (H2O measurements)

Another paper also presents the laser diagnostic technique for temperature measurement at
high pressure [9]. The design and validation of a two-color wavelength-modulation spectroscopy
sensor for measurements of temperature and H2O at temperatures and pressures up to 3000 K and
50 bars were performed. The sensor used two TDLs near 2.474 and 2.482 nm that were fiber-coupled
in free space and frequency multiplexed to enable measurements along a single LOS. Furthermore,
the sensor operates in the fundamental vibration bands of H 2O near 2.5 μm to achieve 5 to 10 times
larger signals than comparable near-infrared sensors. The measured spectra of H 2O are shown in
Figure 17. First-harmonic-normalized WMS with second harmonic detection (WMS-2f/1f) was used
for three primary reasons: (1) the WMS-2f/1f signal is immune to emission and non-absorbing
transmission losses that vary at frequencies much less than the modulation frequency and/or outside
the passband (9 kHz here) centered at the first and second harmonics of each laser’s modulation
frequency, (2) WMS-2f/1f is a differential absorbance technique that does not require knowledge of
the absolute absorbance, and (3) WMS-2f/1f is insensitive to non-Lorentzian effects that can
compromise other absorption methods at high pressures.

17
Figure 17 – Simulated high-pressure absorbance (top) and WMS-2f /1f H2O spectra (bottom) [9]

The Stanford high-pressure shock tube (HPST) was used to validate the sensor’s performance
at high pressures and temperatures. The tube has a total length of 8.4 m, an inner diameter of 5 cm,
and initially consists of a driver and driven section that are separated by a diaphragm. During
operation, the driven section is filled with the test gas (H2O–N2 here), and the driver section is filled
with helium until the diaphragm ruptures. After the diaphragm bursts, a shock wave travels down the
tube thereby setting the test gas into motion and raising its temperature and pressure. Once the shock
wave reaches the end wall, the wave reflects and stagnates the test gas, further raising its temperature
and pressure. The test setup is shown in Figure 18. Measurements of temperature and H 2O were
acquired at two locations within the shock tube using two different TDL sensors. Two fiber-coupled
TDLs near 1,392 and 1,469 nm were used to measure the temperature and H 2O behind the incident
shock. These lasers were multiplexed onto a single polarization-maintaining fiber and the light was
collimated and pitched across the shock tube approximately 92 cm from the end wall. Two narrow-
linewidth TDLs near 2.5 μm were used to measure temperature and H 2O behind the reflected shock.

18
Figure 18 – Experimental setup used in shock tube experiments [9]

To validate this sensor’s accuracy over a broad range of temperatures and pressures relevant
to combustion environments, experiments were performed at temperatures and pressures from 1,000
to 2,700 K and 8 to 50 bar. Figure 19 summarizes the accuracy of the temperature (left) and H 2O
(right) sensing performed. On average, the sensor recovered the known steady-state temperature and
H2O mole fraction within 3.2 and 2.6 % root mean square (RMS) of known values, respectively (i.e.,
within measurement precision). For 75 % of the experiments, the sensor recovered the temperature
and H2O to within less than 2 % of known values. These results suggest that the accuracy of this sensor
is independent of temperature, pressure, and the center wavelength of each laser (of those studied) and
indicate that the spectroscopic model used here is highly accurate.

Figure 19 – Accuracy of temperature (left) and H2O mole fraction (right) sensing [9]

19
Temperature 2D Imaging of reactive flows

Another aim of the recent research in gas turbine measurements is to implement laser
diagnostic techniques to indicate the temperature and species concentration of reactive flows in Jet
Engines. For 2D imaging of the concentration of chemical species, the well-known established
techniques are PLIF, Rayleigh scattering, and HT. Comparison of HT to PLIF and Rayleigh scattering
motivates the HT techniques in reactive flow implementation. For example, the quantitative
interpretation of PLIF measurements requires independent information on temperature and local
quenching rates, which can be difficult or even impossible to obtain in practical reactive flows. The
Rayleigh signals depend on local gas composition, which can make the signal indecipherable in
reactive flows. Furthermore, Reyleigh signals are relatively weak because of their non-resonant and
elastic nature. Lastly, the laser equipment involved in PLIF and Rayleigh scattering are typically not
fiber-coupled, requiring their implementation to be near the target test rig. Such requirements often
pose significant challenges in practice because of the harsh environment created by combustion and
propulsion systems.
The application was designed to demonstrate the HT technique in the exhaust plane of a
practical aero-propulsion engine (General Electric J85) [10]. This setup simultaneously measures the
2D temperature distribution and water vapor concentration (H2O). It showed several unique
advantages of the HT, such as robustness and ease of implementation in practical systems. This
diagnostic system operates near 1350 nm to monitor H 2O absorption features.
The experimental setup was applied to the exhaust stream of a J85 engine, the setup is shown
in Figure 20. The laser signals were transmitted through optical fibers. A 4x32 multiplexer was used
to combine and split the three laser signals into 32 independent outputs. The laser source chosen for
this application was a TDM combination of three FDML lasers. The tomographic frame was built to
hold laser beams in position to create the 15 x 15 grid pattern with 36.3 mm beam spacing.

20
Figure 20 – Experimental setup for reactive flow conditions 2D imaging [10]

Each of the 3 FDMLs can be independently adjusted, allowing choosing a broader range (≈150
cm-1) of absorption to monitor the spectra shape in case of high pressure (e.g., 30 bar). In the current
research, the pressure is equal to atmospheric, so the narrow wavelength sweep (≈10 cm-1) was
chosen. The test setup can measure the temperature which spans 300-2300 K. Using the optic fibers
and multiplexer for 32 outputs implements the Hyperspectral Tomography technique which measures
the 2D image of temperature distribution and H2O concentration Figure 21

Figure 21 – Studied H2O transitions (left) and measured temperature field (right) [10]

21
4 Summary

The observed papers have documented several experimental implementations of laser


absorption techniques in gas turbine environments. The main problems the LAS is facing are the
nonuniformity of gas conditions, broadened spectra at high pressures, and harsh environments in gas
turbines. Despite these complications limiting the broad LAS exploits in the modern jet engines field,
the last experiments showed the possibility and potential of implementation of LAS techniques in the
future. The modern approaches allow to provide accurate measurements of equivalence ratio,
combustion temperature, 2D temperature imaging, and concentrations of emissions.
The WMS techniques can measure equivalence ratio and temperature in conditions close to
the modern gas turbine combustor’s conditions, based on CH4, H2O, and CO2 transitions. It was
shown the ability of steady-state temperature measuring at 50 bar and 2500 K. Also, the HT technique
expands the ability of 2D temperature measuring in reactive flows. Based on the presented results,
sensing combustion instabilities and temperature, exhaust emissions, and their concentration map will
help to increase the controllability of the engines, reduce fuel consumption, reduce noise and pollution,
and increase reliability.
The achievements discussed above make the LAS techniques implementation crucial for
future engine design. For future investigation of these methods, the experiments have to be conducted
using the real engine under a range of different operational conditions. The sensors which will be
possible to install in the real jet engines have to demonstrate a high level of robustness, repeatability,
and opportunity to measure a wide range of parameters. Possibly this technology might be resettable
for different species detection. All these aspects makes the work in this field attractive for engaging
world scientific resources into it.

22
References

[1] Christopher S. Goldenstein, R. Mitchell Spearrin, Jay. B. Jeffries, Ronald K. Hanson, Infrared
laser-absorption sensing for combustion gases, Progress in Energy and Combustion Science,
Volume 60, 2017, Pages 132-176.
[2] C. Goldenstein, I. Schultz, J. Jeffries, and R. Hanson, Two-color absorption spectroscopy
strategy for measuring the column density and path average temperature of the absorbing species
in nonuniform gases, Appl. Opt. 52, 2013, pages 7950-7962.
[3] Hejie Li, Shawn D. Wehe, Keith R. McManus, Real-time equivalence ratio measurements in gas
turbine combustors with a near-infrared diode laser sensor, Proceedings of the Combustion
Institute, Volume 33, Issue 1, 2011, Pages 717-724.
[4] G. Rieker, J. Jeffries, and R. Hanson, Calibration-free wavelength-modulation spectroscopy for
gas temperature and concentration measurements in harsh environments, Appl. Opt. 48, 2009,
pages 5546-5560.
[5] L. Ma and W. Cai, Numerical investigation of hyperspectral tomography for simultaneous
temperature and concentration imaging, Appl. Opt. 47, 2008, pages 3751-3759.
[6] Hejie Li, Xin Zhou, Jay B. Jeffries, and Ronald K. Hanson, Sensing and Control of Combustion
Instabilities in Swirl-Stabilized Combustors Using Diode-Laser Absorption, AIAA
Journal 2007 45:2, pages 390-398.
[7] Hejie Li, Xin Zhou, Jay B. Jeffries, Ronald K. Hanson, Active control of lean blowout in a swirl-
stabilized combustor using a tunable diode laser, Proceedings of the Combustion Institute,
Volume 31, Issue 2, 2007, Pages 3215-3223.
[8] Farooq A, Jeffries JB, Hanson RK, Measurements of CO(2) concentration and temperature at
high pressures using 1f-normalized wavelength modulation spectroscopy with second harmonic
detection near 2.7 microm, Appl Opt. 48(35), 2009, pages 6740-6753.
[9] Goldenstein, C.S., Spearrin, R.M., Jeffries, J.B., Wavelength-modulation spectroscopy near
2.5 μm for H2O and temperature in high-pressure and -temperature gases. Appl. Phys. B 116,
2015, pages 705–716.
[10] L. Ma, X. Li, S. Sanders, A. Caswell, S. Roy, D. Plemmons, and J. Gord, 50-kHz-rate 2D
imaging of temperature and H2O concentration at the exhaust plane of a J85 engine using
hyperspectral tomography, Opt. Express 21, 2013, pages 1152-1162.

23

You might also like