Waiting
Login processing...

Trial ends in Request Full Access Tell Your Colleague About Jove

Engineering

Characterizing Far-infrared Laser Emissions and the Measurement of Their Frequencies

Published: December 18, 2015 doi: 10.3791/53399

Summary

We describe the generation of far-infrared radiation using an optically pumped molecular laser along with the measurement of their frequencies with heterodyne techniques. The experimental system and techniques are demonstrated using difluoromethane (CH2F2) as the laser medium whose results include three new laser emissions and eight measured laser frequencies.

Abstract

The generation and subsequent measurement of far-infrared radiation has found numerous applications in high-resolution spectroscopy, radio astronomy, and Terahertz imaging. For about 45 years, the generation of coherent, far-infrared radiation has been accomplished using the optically pumped molecular laser. Once far-infrared laser radiation is detected, the frequencies of these laser emissions are measured using a three-laser heterodyne technique. With this technique, the unknown frequency from the optically pumped molecular laser is mixed with the difference frequency between two stabilized, infrared reference frequencies. These reference frequencies are generated by independent carbon dioxide lasers, each stabilized using the fluorescence signal from an external, low pressure reference cell. The resulting beat between the known and unknown laser frequencies is monitored by a metal-insulator-metal point contact diode detector whose output is observed on a spectrum analyzer. The beat frequency between these laser emissions is subsequently measured and combined with the known reference frequencies to extrapolate the unknown far-infrared laser frequency. The resulting one-sigma fractional uncertainty for laser frequencies measured with this technique is ± 5 parts in 107. Accurately determining the frequency of far-infrared laser emissions is critical as they are often used as a reference for other measurements, as in the high-resolution spectroscopic investigations of free radicals using laser magnetic resonance. As part of this investigation, difluoromethane, CH2F2, was used as the far-infrared laser medium. In all, eight far-infrared laser frequencies were measured for the first time with frequencies ranging from 0.359 to 1.273 THz. Three of these laser emissions were discovered during this investigation and are reported with their optimal operating pressure, polarization with respect to the CO2 pump laser, and strength.

Introduction

The measurement of far-infrared laser frequencies was first performed by Hocker and co-workers in 1967. They measured the frequencies for the 311 and 337 μm emissions from the direct-discharge hydrogen cyanide laser by mixing them with high order harmonics of a microwave signal in a silicon diode1. To measure higher frequencies, a chain of lasers and harmonic mixing devices were used to generate the laser harmonics2. Eventually two stabilized carbon dioxide (CO2) lasers were chosen to synthesize the necessary difference frequencies3,4. Today, far-infrared laser frequencies up to 4 THz can be measured with this technique using only the first harmonic of the difference frequency generated by two stabilized CO2 reference lasers. Higher frequency laser emissions can also be measured using the second harmonic, such as the 9 THz laser emissions from the methanol isotopologues CHD2OH and CH318OH.5,6 Over the years, the accurate measurement of laser frequencies has impacted a number of scientific experiments7,8 and permitted the adoption of a new definition of the meter by the General Conference of Weights and Measures in Paris in 1983.911

Heterodyne techniques, such as those described, have been immensely beneficial in the measurement of far-infrared laser frequencies generated by optically pumped molecular lasers. Since the discovery of the optically pumped molecular laser by Chang and Bridges12, thousands of optically pumped far-infrared laser emissions have been generated with a variety of laser media. For example, difluoromethane (CH2F2) and its isotopologues generate over 250 laser emissions when optically pumped by a CO2 laser. Their wavelengths range from approximately 95.6 to 1714.1 μm.1315 Nearly 75% of these laser emissions have had their frequencies measured while several have been spectroscopically assigned1618.

These lasers, and their accurately measured frequencies, have played a crucial role in the advancement of high-resolution spectroscopy. They provide important information for infrared spectral studies of the laser gases. Often these laser frequencies are used to verify the analysis of infrared and far-infrared spectra because they provide connections among the excited vibrational state levels that are often directly inaccessible from absorption spectra19. They also serve as the primary radiation source for studies investigating transient, short-lived free radicals with the laser magnetic resonance technique20. With this extremely sensitive technique, rotational and ro-vibrational Zeeman spectra in paramagnetic atoms, molecules, and molecular ions can be recorded and analyzed along with the ability to investigate the reaction rates used to create these free radicals.

In this work, an optically pumped molecular laser, shown in Figure 1, has been used to generate far-infrared laser radiation from difluoromethane. This system consists of a continuous wave (cw) CO2 pump laser and a far-infrared laser cavity. A mirror internal to the far-infrared laser cavity redirects the CO2 laser radiation down the polished copper tube, undergoing twenty six reflections before terminating at the end of the cavity, scattering any remaining pump radiation. Therefore the far-infrared laser medium is excited using a transverse pumping geometry. To generate laser action, several variables are adjusted, some simultaneously, and all are subsequently optimized once laser radiation is observed.

In this experiment, far-infrared laser radiation is monitored by a metal-insulator-metal (MIM) point contact diode detector. The MIM diode detector has been used for laser frequency measurements since 1969.2123 In laser frequency measurements, the MIM diode detector is a harmonic mixer between two or more radiation sources incident on the diode. The MIM diode detector consists of a sharpened Tungsten wire contacting an optically polished Nickel base24. The Nickel base has a naturally occurring thin oxide layer that is the insulating layer.

Once a laser emission was detected, its wavelength, polarization, strength, and optimized operating pressure were recorded while its frequency was measured using the three-laser heterodyne technique2527 following the method originally described in Ref. 4. Figure 2 shows the optically pumped molecular laser with two additional cw CO2 reference lasers having independent frequency stabilization systems that utilize the Lamb dip in the 4.3 μm fluorescence signal from an external, low pressure reference cell28. This manuscript outlines the process used to search for far-infrared laser emissions as well as the method for estimating their wavelength and in accurately determining their frequency. Specifics regarding the three-laser heterodyne technique as well as the various components and operating parameters of the system can be found in Supplemental Table A along with references 4, 25–27, 29, and 30.

Subscription Required. Please recommend JoVE to your librarian.

Protocol

1. Planning of Experiments

  1. Conduct a survey of the literature to assess prior work performed using the laser medium of interest, which for this experiment is CH2F2. Identify all known laser emissions along with all information about the lines such as their wavelength and frequency. Several surveys of known laser emissions are available13,3137.
  2. Compile all spectroscopic investigations of the molecule used as the laser medium with a focus on prior Fourier transform34 and optoacoustic studies38,39.

2. Generating Far-Infrared Laser Emissions

  1. Safety Overview.
    1. Develop a standard operating procedure for the lab that includes proper eye protection when working with the CO2 and far-infrared laser systems.
  2. Alignment and Calibration.
    1. Calibrate each CO2 laser using a grating-based spectrum analyzer designed for the CO2 laser according to the manufacturer’s protocol.
    2. Align the end mirrors and the coupling mirror in the far-infrared laser cavity using a He-Ne laser so that their radiation is focused onto the MIM diode detector.
    3. Direct the radiation from the CO2 pump laser into the far-infrared laser cavity through a sodium chloride window at an angle of approximately 72o with respect to the cavity axis.
    4. Direct the radiation from the two CO2 reference lasers to either their respective low-pressure fluorescence reference cell or co-linearly onto the MIM diode detector using beam splitters and additional mirrors.
  3. Detection of far-infrared laser radiation.
    1. Polish the Nickel base every several days using a standard metal polish.
    2. Crimp a 25 μm tungsten wire into a copper post and bend the wire into the configuration shown in Figure 3.
    3. Adjust the length of the wire so that it is between 10 to 20 wavelengths of the radiation being measured.
    4. Electrochemically etch the tip of the wire in a saturated sodium hydroxide (NaOH) solution by applying a voltage (approximately 3.5 to 5 VAC) to the solution.
    5. Re-etch the tip with a low voltage (less than 1 VAC). This roughens the tip of the wire and improves the diode’s performance.
    6. Rinse the wire with distilled water.
    7. Insert the copper post into the MIM diode’s housing once the wire is dry.
    8. Place the wire in contact with the Nickel base using a fine screw and level system. Contacts yielding a resistance across the diode between 100 and 500 Ω are typically used when detecting and measuring far-infrared laser radiation.
  4. Generation of far-infrared laser radiation.
    1. Set the CO2 pump laser on a specific laser emission, e.g., 9P36.
    2. Rotate the micrometer dial on the CO2 pump laser back and forth to achieve maximum intensity on the beam stop.
    3. Adjust the tilt of the CO2 pump laser’s grating to achieve maximum intensity on the beam stop.
    4. Repeat steps 2.4.2 and 2.4.3 until the output power for the CO2 pump laser appears optimized on the beam stop.
    5. Remove the beam stop from the path of the CO2 pump laser.
    6. Turn on and align the optical chopper into the beam path of the CO2 pump laser.
    7. Open the valve on the CH2F2 cylinder to introduce the far-infrared laser medium into the far-infrared laser cavity.
    8. Adjust the metering valve on the inlet line until a pressure of approximately 10 Pa is achieved.
      Note: Only the approximate pressure is necessary since it is used as a way of systematically scanning the far-infrared laser cavity.
    9. Set the position of the output coupler such that its outermost tip is approximately 1 cm from the middle of the laser cavity as indicated by a calibrated scale on the outside of the laser cavity.
      Note: Only the approximate location is necessary since it is used as a way of systematically scanning the far-infrared laser cavity.
    10. Adjust the position of the moveable far-infrared laser mirror in approximately 0.25 mm increments by rotating the calibrated micrometer dial back and forth. Simultaneously tune the frequency of the CO2 pump laser through its gain curve by changing the voltage applied across the CO2 pump laser’s piezoelectric transducer (PZT).
    11. If no signal is observed on the oscilloscope display, repeat step 2.4.10 with the output coupler moved to its next position where the tip is approximately 1.5 cm from the middle of the laser cavity as indicated by a calibrated scale on the outside of the laser cavity.
    12. If no signal is observed on the oscilloscope display, repeat step 2.4.10 with the output coupler moved to its next position where the tip is approximately 2 cm from the middle of the laser cavity as indicated by a calibrated scale on the outside of the laser cavity.
    13. If no signal is observed on the oscilloscope display, repeat steps 2.4.9 through 2.4.12 with a far-infrared laser pressure of approximately 19 Pa as adjusted with the metering valve on the inlet line.
    14. If no signal is observed on the oscilloscope display, repeat steps 2.4.9 through 2.4.12 with a far-infrared laser pressure of approximately 27 Pa as adjusted with the metering valve on the inlet line.
    15. If no signal is observed on the oscilloscope display, insert the beam stop into the path of the CO2 pump laser and close the valve on the CH2F2 cylinder until the far-infrared laser pressure is approximately 0 Pa.
    16. Set the CO2 pump laser to the next laser emission, e.g., 9P34, and optimize the output power using steps 2.4.2 through 2.4.4.
    17. Repeat steps 2.4.5 through 2.4.16 until all emissions generated by the CO2 pump laser are used. When searching for far-infrared laser lines, place a focus on CO2 pump laser emissions whose frequencies overlap with any absorption regions identified in step 1.2.
  5. Characterizing far-infrared laser emissions.
    1. Simultaneously adjust the pressure of the far-infrared laser medium, the voltage applied to the CO2 pump laser’s PZT, and the position of the output coupler until the far-infrared laser emission’s output power is maximized (determined by a maximum peak-to-peak signal from the MIM diode detector as observed on the oscilloscope display, similar to Figure 4).
    2. Turn the micrometer dial clockwise until the far-infrared laser emission is observed on the oscilloscope display. Record the position of the micrometer dial.
    3. Turn the micrometer dial clockwise for an additional 20 modes corresponding to the same far-infrared laser emission. Record the position of the micrometer dial.
    4. Subtract the position of the micrometer dial in steps 2.5.2 and 2.5.3. Divide this difference by 10 to obtain the wavelength of the far-infrared laser emission.
    5. Repeat steps 2.5.2 through 2.5.4 a total of five times and average the wavelength of the far-infrared laser emission. Average laser wavelengths measured by traversing at least 20 adjacent longitudinal modes have a one-sigma uncertainty of ± 0.5 μm.
    6. Measure the polarization of the far-infrared laser radiation, relative to the CO2 pump radiation, using either a gold wire-grid polarizer (394 lines/cm) or a Brewster polarizer.

3. Determining Far-Infrared Laser Frequencies

  1. Identifying the CO2 reference laser emissions.
    1. Calculate the frequency of the far-infrared laser emission based on its measured wavelength.
    2. Identify sets of CO2 reference laser lines whose frequency difference is within several GHz of the calculated frequency for the far-infrared laser emission40. A typical list used for such measurements is shown in Table 1.
  2. Searching for the heterodyne beat signal.
    1. Identify the first set of CO2 reference laser lines and set each CO2 reference laser on their respective laser emission.
    2. Optimize the output power for each CO2 reference laser using steps 2.4.2 through 2.4.4 and the monitor power meter.
      1. Adjust an iris, either internal or external to each reference laser, so that the power from each CO2 reference laser is approximately 100 mW as measured by the monitor power meter shown in Figure 2.
    3. Block the radiation from the CO2 pump laser using a beam stop while unblocking the radiation from the CO2 reference lasers.
    4. Turn on and align the optical chopper into the co-linear beam path of the CO2 reference lasers.
    5. Optimize for maximum peak-to-peak voltage each CO2 reference laser emission on the MIM diode detector using several mirrors, beam splitters, and a 2.54 cm focal length ZnSe plano-convex lens while observing the output on the oscilloscope, similar to Figure 5.
    6. Block the radiation from the CO2 reference lasers using a beam stop while unblocking the radiation from the CO2 pump laser.
    7. Re-optimize the CO2 pump laser and the far-infrared laser, as necessary, so that the far-infrared laser emission has a maximum peak-to-peak voltage as observed on the oscilloscope.
    8. Disconnect the MIM diode detector’s output from the oscilloscope and connect it to an amplifier whose output is observed on a spectrum analyzer.
    9. Unblock the radiation from the CO2 reference lasers.
    10. Remove the optical choppers modulating the CO2 pump and reference lasers.
    11. Set the spectrum analyzer on a 40 MHz span and search for the beat signal in 1.5 GHz increments by manually scanning this frequency range using the spectrum analyzer’s adjustment knob.
    12. If no beat signal is observed, disconnect the MIM diode’s output from the amplifier and connect it to the oscilloscope.
    13. Block the radiation from the CO2 reference lasers and reinsert the optical chopper into the path of the CO2 pump laser.
    14. Repeat steps 3.2.2 through 3.2.13 as necessary until the spectrum analyzer has been used to search for the beat signal between 0 and 12 GHz.
    15. If no beat signal is observed, repeat steps 3.2.2 through 3.2.14 with another set of CO2 reference laser lines until either the beat signal is observed or all possible sets of CO2 reference laser lines are exhausted.
  3. Stabilizing the CO2 reference frequencies.
    1. Apply a voltage between 0 and 900 V to the first CO2 reference laser’s PZT so that the signal from its respective fluorescence reference cell is at the center of the Lamb dip, illustrated in Figure 6 and as viewed on an oscilloscope as in Figure 7.
    2. Activate the feedback voltage applied to the first CO2 reference laser’s PZT using a custom built lock-in/servo amplifier so that it remains locked to the center of the Lamb dip.
    3. Repeat steps 3.3.1 and 3.3.2 for the second CO2 reference laser.
    4. Visually monitor the output of the pre-amp on an oscilloscope, as in Figure 7, to ensure the reference lasers remains locked.
  4. Measurement of the beat frequency.
    1. Center the beat signal on the spectrum analyzer display and adjust its amplitude to maximize its size on the display.
    2. Set the spectrum analyzer to view two simultaneous traces of the beat signal, as in Figure 8, by selecting the Clear Write feature for both Trace 1 and Trace 2. One trace will display the instantaneous signal while the other will record the maximum signal (using a Max Hold feature on the spectrum analyzer for the second trace).
    3. Rotate the micrometer dial on the far-infrared laser cavity back and forth across the gain curve for a given cavity mode.
    4. Use the View feature on the spectrum analyzer to freeze the second (Max Hold) trace once a symmetric pattern is obtained.
    5. Slightly rotate the micrometer dial clockwise to decrease the length of the far-infrared laser cavity. Simultaneously observe the subsequent small shift in the beat frequency on the spectrum analyzer due to this slight increase in the frequency of the far-infrared laser.
    6. Place markers at the full width at half maximum points of the symmetric pattern (Max Hold trace) using the Marker function with the Delta feature on the spectrum analyzer.
    7. Measure the center frequency of the beat signal using the Span Pair feature on the spectrum analyzer.
    8. Repeat steps 3.4.1 through 3.4.7.
    9. Disengage the lock in/servo amplifier for each CO2 reference laser to unlock each laser from its center frequency and re-optimize each CO2 reference laser.
    10. Re-lock the reference lasers using steps 3.3.1 through 3.3.4.
    11. Repeat steps 3.4.1 through 3.4.10 for a total of 6 measurements. Once complete, unlock each CO2 reference laser from its center frequency.
    12. Calculate the revised frequency of the far-infrared laser emission using these beat frequencies to obtain an accurate prediction for the second set of CO2 reference laser lines.
    13. Identify a different set of CO2 reference laser lines whose frequency difference is within several GHz of the calculated frequency for the far-infrared laser emission.
    14. Optimize the next set of CO2 reference laser lines on the MIM diode detector and obtain the beat signal using steps 3.2.2 through 3.2.15 as necessary.
    15. Lock the new set of CO2 reference laser lines using steps 3.3.1 through 3.3.4.
    16. Repeat steps 3.4.1 through 3.4.10 for a total of 6 measurements. Once complete, unlock each CO2 reference laser from its center frequency.
    17. Insert beam stops into the paths of the CO2 pump and reference lasers.
  5. Calculation of the far-infrared laser frequency.
    1. Calculate the unknown far-infrared laser frequency, νFIR, using the measured beat frequency through the relation
                                                         FIR = |νCO2(I)–νCO2(II)| ± |νbeat|                               Eq. 1
      where |νCO2(I)–νCO2(II)| is the magnitude of the difference frequency synthesized by the two CO2 reference lasers and |νbeat| is the magnitude of the beat frequency. The ± sign in Eq. 1 is determined experimentally from step 3.4.5.
    2. Obtain an average frequency and calculate the uncertainty.

Subscription Required. Please recommend JoVE to your librarian.

Representative Results

As mentioned, the frequency reported for a far-infrared laser emission is an average of at least twelve measurements performed with at least two different sets of CO2 reference laser lines. Table 2 outlines the data recorded for the 235.5 μm laser emission when using the 9P04 CO2 pump laser. For this far-infrared laser emission, fourteen individual measurements of the beat frequency were recorded. The first set of measurements were recorded while using the 9R10 and 9P38 CO2 reference laser emissions. For step 3.4.5, as the far-infrared laser frequency was increased slightly, the beat frequency was also observed to increase. This indicates the far-infrared laser frequency was greater than the magnitude of the difference frequency between the 9R10 and 9P38 CO2 reference lasers, |νCO2(I)–νCO2(II)|. Therefore the sign of the beat frequency in Equation 1 was positive for this set of CO2 reference lasers. Conversely, the second set of measurements used the 9R16 and 9P34 CO2 reference laser emissions. When step 3.4.5 was performed, a decrease in the beat frequency was observed while the far-infrared laser frequency was increased slightly. This indicates the far-infrared laser frequency was less than the magnitude of the difference frequency between the 9R16 and 9P34 CO2 reference lasers. Therefore, for this set of CO2 reference lasers the sign of the beat frequency in Equation 1 was negative. As illustrated in Table 2, the calculated far-infrared laser frequency, νFIR, for both situations remained the same to within a ± 0.12 MHz one-sigma standard deviation.

The average far-infrared laser frequencies determined with this experimental technique are listed in Table 3 and are arranged in order of the CO2 pump line. The average laser frequencies are reported with their corresponding wavelength and wavenumber, calculated using 1 cm-1 = 29 979.2458 MHz. All far-infrared laser frequencies were measured under optimal operating conditions. Throughout this investigation, several previously reported frequencies were measured and were found to be in agreement with the published values. The one-sigma fractional uncertainty, Δν, of far-infrared laser frequencies measured with this technique is ± 5 × 107. This uncertainty is derived from the reproducibility of known frequencies with this system, the symmetry and width of the broadened gain curve of the far-infrared laser, and the precision of the measurements4,25,31.

The far-infrared laser emissions discovered during this investigation were observed to have a strength of ‘W’ corresponding to a range in power from 0.001 to 0.01 mW. For comparison, the 118.8 μm line of methanol was observed with this system to be VVS with a power slightly above 10 mW when using the 9P36 CO2 pump having a power of 18 W. Additionally, Table 3 includes the polarization of each new far-infrared laser emission measured relative to its respective CO2 pump laser. In most cases, only one polarization was observed to dominate, either a polarization parallel or perpendicular to the CO2 pump laser. For situations where no dominant polarization was observed, both polarizations have been listed.

In sum, eight far-infrared laser emissions were generated by difluoromethane using an optically pumped molecular laser system having a transverse pumping geometry. This includes the discovery of three far-infrared laser emissions having wavelengths of 235.5, 335.9, and 416.8 μm. Once detected, the three-laser heterodyne technique was used to measure the frequency for each observed far-infrared laser emission. The frequencies for these laser emissions ranged from 0.359 to 1.273 THz and are reported with fractional uncertainties of ± 5 parts in 107.

Figure 1
Figure 1. Schematic diagram of the optically pumped molecular laser system consisting of a carbon dioxide pump laser and a far-infrared laser cavity. The far-infrared laser medium is excited using a transverse pumping geometry. Reprinted with minor modifications from Ref. 15 with kind permission from Springer Science and Business Media. Please click here to view a larger version of this figure.

Figure 2
Figure 2. Schematic diagram of the three-laser heterodyne frequency measurement system. The heterodyne system includes the optically pumped molecular laser utilizing a transverse pumping geometry and two additional carbon dioxide reference lasers. Not shown are the electronic systems used to monitor and stabilize the radiation generated by each laser. © [2015] IEEE. Reprinted, with minor modifications and permission, from Ref. 27. Please click here to view a larger version of this figure.

Figure 3
Figure 3. The Tungsten wire used in the MIM point contact diode detector as viewed through a magnifying lens. The length of the wire is approximately 2 mm. For best spring action, the angles in the bend should be near 90o and all lie in the same plane.

Figure 4
Figure 4. The waveform generated by the 274.8 μm laser emission of optically pumped CH2F2 using the 9P04 CO2 pump laser as viewed on the oscilloscope display. The CO2 pump radiation is modulated by an optical chopper operating at approximately 45 Hz. The resistance of the MIM diode detector is approximately 100 and the signal is approximately 6 μV (peak-to-peak). The oscilloscope display is set on 10 μV/division.

Figure 5
Figure 5. The left and middle photos show the output from each CO2 reference laser, 9R16 and 9P34, respectively. The respective modulated signal on the oscilloscope is approximately 4 mV (peak-to-peak) for about 100 mW of power, as measured by the monitor power meter. The right photo shows the combined signal from both reference lasers to be approximately 7 mV (peak-to-peak) indicating the two reference signals are properly mixing on the MIM diode detector. The resistance of the MIM diode detector is approximately 100 Ω. The oscilloscope display in each photo is set on 1 mV/division. The CO2 radiation is modulated by an optical chopper operating at approximately 70 Hz.

Figure 6
Figure 6. The saturated fluorescence signal in low pressure (6 Pa) CO2 while using the 9R24 CO2 laser emission. This graph is obtained by modulating the CO2 reference laser emission via an external chopper at 52 Hz while the voltage applied to the CO2 reference laser’s PZT is ramped from 0 to approximately 570 V in about 13 min. The lock-in amplifier is set to a 300 msec time constant and a 200 mV sensitivity. Please click here to view a larger version of this figure.

Figure 7
Figure 7. The saturated fluorescence signal in low pressure (6 Pa) CO2 while using the 9R24 CO2 laser emission as viewed on an oscilloscope. The left photo indicates the oscilloscope display when the PZT voltage is away from the center of the Lamb dip, approximately 80 V in this photo. The middle and right photos indicate the oscilloscope display when the PZT voltage is either immediately to the left or right of the center of the Lamb dip, approximately 278 and 295 V respectively in these photos. Please click here to view a larger version of this figure.

Figure 8
Figure 8. The beat signal between the 235.5 μm laser emission of optically pumped CH2F2 using the 9P04 CO2 pump laser and the 9R16 and 9P34 CO2 reference lasers. A span of approximately 25 MHz is typically used. The majority of beat signals are observed within ± 5 GHz. However, there are certain frequency regions within these search parameters that have a low signal-to-noise. Therefore, using a slightly larger search region has sometimes been helpful.

Figure 9
Figure 9. Portion of a typical laser resonator interferogram (or cavity scan) consisting of a set of discrete peaks that correspond to the resonator's modes, separated by regions where no lasing occurs. This scan shows the 511.445 μm laser emission generated by optically pumped CH2F2 using the 9R28 CO2 pump. A decrease in the micrometer position corresponds to a decrease in the length (mirror-to-mirror separation) of the far-infrared laser cavity. The MIM diode detected a 20 μV peak-to-peak maximum signal generated by this far-infrared laser emission. The output from the detector was recorded using a lock-in amplifier, set on a 300 msec time constant and 20 μV sensitivity, interfaced to a computer. Please click here to view a larger version of this figure.

Table 1
Table 1: Sets of CO2 reference lasers whose difference frequency is near the calculated frequency for the 235.5 μm laser emission from optically pumped CH2F2 when excited using the 9P04 CO2 laser emission.

Table 2
Table 2: Measured beat frequencies for the 235.5 μm laser emission from optically pumped CH2F2 when excited using the 9P04 CO2 laser emission. Two sets of CO2 reference lasers are used to generate the known difference frequency (|νCO2(I)–νCO2(II)|).

Table 3
Table 3: New far-infrared laser frequencies from optically pumped CH2F2.

Table 4
Supplemental Table A: Technical details of the experimental system including some relevant commercial components.

Subscription Required. Please recommend JoVE to your librarian.

Discussion

There are several critical steps within the protocol that require some additional discussion. When measuring the far-infrared laser wavelength, as outlined in step 2.5.3, it is important to ensure the same mode of the far-infrared laser emission is being used. Multiple modes of a far-infrared laser wavelength (i.e., TEM00, TEM01, etc.) can be generated within the laser cavity and thus it is important to identify the appropriate adjacent cavity modes being used to measure the wavelength13,29,41. To assist in eliminating higher order modes, irises are included within each laser cavity. When accurately measuring a far-infrared laser frequency, it is imperative the lasers, particularly the CO2 reference lasers, operate in their fundamental (TEM00) mode. Irises are also used to ensure the pattern traced out by the far-infrared laser on the spectrum analyzer is symmetric. For situations where multiple far-infrared laser wavelengths are generated by a particular CO2 pump line, as in the case of 9P04, a set of absorbing filters, calibrated with wavelength, are used to assist in distinguishing far-infrared laser wavelengths. They can also be used to attenuate any scattered CO2 laser radiation exiting the far-infrared laser cavity.

Section 2.4 describes the generation of far-infrared laser radiation. Over numerous investigations, we have found that multiple distinct wavelengths could be generated by the same CO2 pump laser set at slightly different offset frequencies. For example, the 9P04 CO2 pump laser is capable of generating the 289.5 and 724.9 μm wavelengths of CH2F2 at one pump frequency while the remaining wavelengths measured during this investigation were generated using a slightly different frequency from the 9P04 CO2 pump laser. This is accomplished by changing the voltage applied to the PZT that tunes the frequency of the CO2 pump laser through its broadened gain curve (approximately ± 45 MHz from its center frequency in this experiment). Although not specifically addressed in section 2.4, we believe this is a noteworthy feature in searching for far-infrared laser radiation.

For situations where multiple far-infrared laser emissions are generated by the same CO2 pump laser line at the same offset frequency, a laser resonator interferogram (or cavity scan) can be performed to assist in identifying the different far-infrared laser emissions being generated. Figure 9 illustrates a portion of a typical laser resonator interferogram, with the output power plotted as a function of decreasing far-infrared laser cavity length4245.

As outlined in section 3.4, two distinct sets of CO2 reference lasers are used to measure the far-infrared laser frequency. This helps eliminate the uncertainty about whether the beat frequency is above or below the difference frequency generated between the CO2 reference lasers. Along with providing a way to independently verify the far-infrared laser frequency, it has been particularly useful when working with weak beat signals where observing the slight shift in the beat frequency as the far-infrared laser frequency increases can be challenging.

The MIM diode detector is an essential component to this experimental system due to its high speed, sensitivity, and broad spectral coverage23,24. However, there are some limitations to the MIM diode detector that include mechanical instability, susceptibility to electromagnetic disturbances, poor reproducibility, and a limit to the maximum power it is capable of detecting while maintaining its sensitivity. While measuring far-infrared laser frequencies, the sensitivity of the MIM diode detector was found to decrease rapidly over time if the power from each CO2 reference laser exceeded 150 mW.

Beyond the MIM diode detector, the main limitation to the present technique is the stability of the far-infrared laser4,31,46. A limitation in the experimental system’s current configuration is the inability to measure the offset frequency of the CO2 pump laser. As mentioned, the offset frequency is defined as the difference between the frequency used by the CO2 pump laser to generate the far-infrared laser emission and the CO2 pump laser’s center frequency. Thus it represents the difference between the absorption frequency of the far-infrared laser medium and the center frequency of the CO2 pump laser. Typically, the offset frequency is readily measured using any CO2 laser radiation that is inadvertently scattered out of the far-infrared laser cavity. In our current configuration however, very little CO2 laser radiation is available for such a measurement. Other methods of measuring the offset frequency could be incorporated into future iterations of the project. This includes using additional beam splitters and mirrors to couple a portion of the pump radiation to the MIM diode detector. The measurement of an offset frequency is beneficial when assigning spectroscopic transitions to the far-infrared laser emission25,34.

Far-infrared laser frequencies have also been measured by heterodyning two optically pumped far-infrared lasers and a microwave source on a MIM diode detector whereby the frequency of one of the two far-infrared lasers is known and is used as the reference frequency47. The use of far-infrared frequencies with greater accuracy is possible using other techniques, such as with THz frequency-comb synthesis similar to those discussed in Refs. 48-54. Measuring laser frequencies expands the role of optically pumped molecular lasers in THz applications from THz imaging55, its role as a source of THz radiation for high-resolution spectroscopy13,20, and in assisting with the analysis of the complex spectra associated with its lasing medium19,34,37.

Subscription Required. Please recommend JoVE to your librarian.

Disclosures

Certain commercial equipment is identified in this paper to foster understanding. Such identification does not imply recommendation or endorsement by the authors, nor does it imply that the equipment identified is necessarily the best available for the purpose.

Acknowledgments

This work was supported in part by the Washington Space Grant Consortium under Award NNX10AK64H.

Materials

Name Company Catalog Number Comments
Vacuum pump Leybold Trivac D4A HE-175 oil; Quantity = 3
Vacuum pump Leybold Trivac D8B or D16B Fomblin Fluid; Quantity = 1 of each
Vacuum pump Leybold Trivac D25B HE-175 oil; Quantity = 1
Optical chopper with controller Stanford Research Systems SR540
Lock-in amplifier Stanford Research Systems SR830
Spectrum analyzer Agilent E4407B ESA-E Series, 9 kHz to 26.5 GHz Spectrum Analyzer
Amplifier  Miteq AFS-44 Provides amplification of signals between 2 and 18 GHz. The amplifier is powered by a Hewlett Packard triple output DC power supply, model E3630A.
Amplifier  Avantek AWL-1200B Provides amplification of signals less than 1.2 GHz.
Power supply Hewlett Packard E3630A Low voltage DC power supply for amplifier.
Power supply Glassman KL Series High voltage power supply for the CO2 lasers; Quantity = 2; negative polarity
Power supply Fluke 412B High voltage power supply used with the NIST Asymmetric HV Amp
Detector Judson Infrared Inc J10D For fluorescence cell; Quantity = 2
CO2 laser spectrum analyzer Optical Engineering  16-A Currently sold by Macken Instruments Inc.
Thermal imaging plates with UV light Optical Engineering  Primarily used for aligning the CO2 reference lasers. Currently sold by Macken Instruments Inc.
Resistors Ohmite  L225J100K 100 kW, 225 W. Between 4 to 6 resistors are used in each ballast system. Each CO2 laser has its own ballast system. Fans are used to cool the resistors.
HV relay, SPDT CII Technologies H-17 Quantity = 3; one for each CO2 laser
Amplifier  Princeton Applied Research PAR 113 Used with fluorescence cell; Quantity = 2
Oscilloscope Tektronix 2235A Similar models are also used; Quantity = 2
Oscilloscope/Differential amplifier Tektronix 7903 oscilloscope with 7A22 differential amplifier
Power meter with sensor Coherent 200 For use below 10 W.  This is the power meter shown in Figure 2.
Power meter with sensor Scientech, Inc Vector S310 For use below 30 W
Multimeter Fluke 73III Similar models are also used; Quantity = 3
Data acquisition National Instruments NI cDAQ 9174 chassis with NI 9223 input module Uses LabVIEW software
Simichrome polish Happich GmbH Polish for the Nickel base used in the MIM diode detector. Although the Nickel base can be used immediately after polishing, a 12 hour lead time is typically recommended.
Pressure gauge Wallace and Tiernan 61C-1D-0050 Series 300; for CO2 laser; Quantity = 3
Pressure gauge with controller Granville Phillips Series 375 For far-infrared laser
Zirconium Oxide felt Zircar Zirconia ZYF felt Used as a beam stop
Zirconium Oxide board Zircar Zirconia ZYZ-3 board Used as a beam stop; Quantity = 4
Teflon sheet Scientific Commodities, Inc BB96312-1248 1/32 inch thick; used for the far-infrared laser output window
Polypropylene C-Line sheet protectors 61003 used for the far-infrared laser output window
Vacuum grease Apiezon
Power supply Kepco NTC 2000 PZT power supply
PZT tube Morgan Advanced Materials 1 inch length, 1 inch outer diameter, 0.062 inch thickness, reverse polarity (positive voltage on outside); Quantity = 3
ZnSe (AR coated) II-VI Inc CO2 laser window (Quantity = 3), lens, and beam splitter (Quantity 3)
NaCl window Edmond Optics Quantity = 1
CaF window Edmond Optics Quantity = 2
Laser mirrors and gratings Hyperfine, Inc Gold-coated; includes positioning mirrors
Glass laser tubes and reference cells Allen Scientific Glass
MIM diode detector Custom Microwave, Inc
Other Other materials include magnetic bases, base plates, base clamps, XYZ translation stage, etc.

DOWNLOAD MATERIALS LIST

References

  1. Hocker, L. O., Javan, A., Ramachandra Rao, D., Frenkel, L., Sullivan, T. Absolute frequency measurement and spectroscopy of gas laser transitions in the far infrared. Appl. Phys. Lett. 10 (5), 147-149 (1967).
  2. Wells, J. S., Evenson, K. M., Day, G. W., Halford, D. Role of infrared frequency synthesis in metrology. Proc. IEEE. 60 (5), 621-623 (1972).
  3. Whitford, B. G., Siemsen, K. J., Riccius, H. D., Baird, K. A. New frequency measurements and techniques in the 30-THz region. IEEE Trans. Instrum. Meas. 23 (4), 535-539 (1974).
  4. Petersen, F. R., et al. Far infrared frequency synthesis with stabilized CO2 lasers: Accurate measurements of the water vapor and methyl alcohol laser frequencies. IEEE J. Quantum Elect. 11 (10), 838-843 (1975).
  5. Uranga, C., Connell, C., Borstad, G. M., Zink, L. R., Jackson, M. Discovery and frequency measurement of short-wavelength far-infrared laser emissions from optically pumped 13CD3OH and CHD2OH. Appl. Phys. B. 88 (4), 503-505 (2007).
  6. Jackson, M., Milne, J. A., Zink, L. R. Measurement of optically pumped CH318OH laser frequencies between 3 and 9 THz. IEEE J. Quantum Elect. 47 (3), 386-389 (2011).
  7. Evenson, K. M., et al. Optically pumped FIR lasers: Frequency and power measurements and laser magnetic resonance spectroscopy. IEEE J. Quantum Elect. 13 (6), 442-444 (1977).
  8. Evenson, K. M., Jennings, D. A., Petersen, F. R. Tunable far-infrared spectroscopy. Appl. Phys. Lett. 44 (6), 576-577 (1984).
  9. Evenson, K. M., et al. Speed of light from direct frequency and wavelength measurements of the methane-stabilized laser. Phys. Rev. Lett. 29 (19), 1346-1349 (1972).
  10. BIPM. Resolution 1. Comptes Rendus des Séances de la 17e Conférence Générale des Poids et Mesures, Sevres, France, , 97-98 (1983).
  11. Giacomo, P. News from the BIPM. Metrol. 20 (1), 25-30 (1984).
  12. Chang, T. Y., Bridges, T. J. Laser action at 452, 496 and 541 µm in optically pumped CH3F. Opt. Commun. 1 (9), 423-426 (1970).
  13. Douglas, N. G. Millimetre and Submillimetre Wavelength Lasers: A Handbook of CW Measurements. Walter, H. 61, Springer Series in Optical Sciences. Springer-Verlag. (1989).
  14. Zerbetto, S. C., Vasconcellos, E. C. C., Zink, L. R., Evenson, K. M. 12CH2F2 and 13CH2F2 far-infrared lasers: New lines and frequency measurements. Int. J. Infrared Millim. Waves. 18 (12), 2301-2306 (1997).
  15. Jackson, M., Alves, H., Holman, R., Minton, R., Zink, L. R. New cw optically pumped far-infrared laser emissions generated with a transverse or ‘zig-zag’ pumping geometry. J. Infrared, Millim., Terahertz Waves. 35 (3), 282-287 (2014).
  16. Danielewicz, E. J. The optically pumped difluoromethane far-infrared laser. Reviews of Infrared and Millimeter Waves. Button, K. J., Inguscio, M., Strumia, F. 2, Plenum. 223-250 (1983).
  17. Deroche, J. C., Benichou, E. K., Guelachvili, G., Demaison, J. Assignments of submillimeter emissions in difluoromethane pumped by 12C18O2 and 12C18O2 lasers. Int. J. Infrared Millim. Waves. 7 (10), 1653-1675 (1986).
  18. Jackson, M., Zink, L. R., McCarthy, M. C., Perez, L., Brown, J. M. The far-infrared and microwave spectra of the CH radical in the v = 1 level of the X2Π. J. Mol. Spectrosc. 247 (2), 128-139 (2008).
  19. Zhao, S., Lees, R. M. CH318OH: Assignment of FIR laser lines optically pumped in the in-plane CH3-rocking band. J. Mol. Spectrosc. 168 (1), 67-81 (1994).
  20. Evenson, K. M., Saykally, R. J., Jennings, D. A., Curl, R. F., Brown, J. M. Far infrared laser magnetic resonance. Chemical and Biochemical Applications of Lasers. 5, Academic Press. 95-138 (1980).
  21. Hocker, L. O., Sokoloff, D. R., Daneu, V., Szoke, A., Javan, A. Frequency mixing in the infrared and far-infrared using a metal-to-metal point contact diode. Appl. Phys. Lett. 12 (12), 401-402 (1968).
  22. Daneu, V., Sokoloff, D., Sanchez, A., Javan, A. Extension of laser harmonic-frequency mixing techniques into the 9 μ region with an infrared metal-metal point-contact diode. Appl. Phys. Lett. 15 (12), 398-400 (1969).
  23. Jennings, D. A., Evenson, K. M., Knight, D. J. E. Optical Frequency Measurements. Proc. IEEE. 74 (1), 168-179 (1986).
  24. Zink, L. R. Highly accurate molecular constants for CO, HF, HCl, OH, NaH, MgH, and O2: Rotational transition frequencies measured with tunable far infrared radiation [thesis]. , University of Colorado. (1986).
  25. Xu, L. -H., et al. Methanol and the optically pumped far-infrared laser. IEEE J. Quantum Elect. 32 (3), 392-399 (1996).
  26. Jackson, M., Zink, L. R., Garrod, T. J., Petersen, S., Stokes, A., Theisen, M. The generation and frequency measurement of short-wavelength far-infrared laser emissions. IEEE J. Quantum Elect. 41 (12), 1528-1532 (2005).
  27. Jackson, M., Smith, M., Gerke, C., Barajas, J. M. Measurement of far-infrared laser frequencies from methanol isotopologues. IEEE J. Quantum Elect. 51 (4), 1500105 (2015).
  28. Freed, C., Javan, A. Standing-wave saturation resonances in the CO2 10.6 μ transitions observed in a low-pressure room-temperature absorber gas. Appl. Phys. Lett. 17 (2), 53-56 (1970).
  29. DeShano, B., Olivier, K., Cain, B., Zink, L. R., Jackson, M. Using guide wavelengths to assess far-infrared laser emissions. J. Infrared, Millim., Terahertz Waves. 36 (1), 13-30 (2015).
  30. Jackson, M., Nichols, A. J., Womack, D. R., Zink, L. R. First laser action observed from optically pumped CH317OH. IEEE J. Quantum Elect. 48 (3), 303-306 (2012).
  31. Inguscio, M., Moruzzi, G., Evenson, K. M., Jennings, D. A. A review of frequency measurements of optically pumped lasers from 0.1 to 8 THz. J. Appl. Phys. 60 (12), R161-R191 (1986).
  32. Pereira, D., et al. A review of optically pumped far-infrared laser lines from methanol isotopes. Int. J. Infrared Millim. Waves. 15 (1), 1-44 (1994).
  33. Zerbetto, S. C., Vasconcellos, E. C. C. Far infrared laser lines produced by methanol and its isotopic species: A review. Int. J. Infrared Millim. Waves. 15 (5), 889-933 (1994).
  34. Moruzzi, G., Winnewisser, B. P., Winnewisser, M., Mukhopadhyay, I., Strumia, F. Microwave, Infrared and Laser Transitions of Methanol: Atlas of Assigned Lines from 0 to 1258 cm-1. , CRC Press. FL. (1995).
  35. Handbook of Laser Wavelengths. Weber, M. J. , CRC Press. FL. (1999).
  36. De Michele, A., et al. FIR laser lines from CH3OD: A review. Int. J. Infrared Millim. Waves. 25 (5), 725-734 (2004).
  37. De Michele, A., Carelli, G., Moruzzi, G., Moretti, A. Hydrazine far-infrared laser lines and assignments: a review. J. Opt. Soc. Am. B. 22 (7), 1461-1470 (2005).
  38. Moraes, J. C. S., et al. Experimental investigation of 13CD3OH infrared transitions by means of optoacoustic spectroscopy. Int. J. Infrared Millim. Waves. 13 (11), 1801-1823 (1992).
  39. Viscovini, R. C., Scalabrin, A., Pereira, D. Infrared optoacoustic spectroscopy of 13CD3OD around the 10R and 10P CO2 laser lines. Int. J. Infrared Millim. Waves. 17 (11), 1821-1838 (1996).
  40. Maki, A. G., Chou, C. C., Evenson, K. M., Zink, L. R., Shy, J. T. Improved molecular constants and frequencies for the CO2 laser from new high-J regular and hot-band frequency measurements. J. Mol. Spectrosc. 167 (1), 211-224 (1994).
  41. Douglas, N. G., Krug, P. A. CW laser action in ethyl chloride. IEEE J. Quantum Elect. 18 (10), 1409-1410 (1982).
  42. Schwaller, P., Steffen, H., Moser, J. F., Kneubühl, F. K. Interferometry of resonator modes in submillimeter wave lasers. Appl. Opt. 6 (5), 827-829 (1967).
  43. Steffen, H., Kneubühl, F. K. Resonator interferometry of pulsed submillimeter-wave lasers. IEEE J. Quantum Elect. 4 (12), 992-1008 (1968).
  44. Whitbourn, L. B., Macfarlane, J. C., Stimson, P. A., James, B. W., Falconer, I. S. An experimental study of a cw optically pumped far infrared formic acid vapour laser. Infrared Phys. 28 (1), 7-20 (1988).
  45. Belland, P., Véron, D., Whitbourn, L. B. Mode study, beam characteristics and output power of a cw 337 μm HCN waveguide laser. J. Phys. D: Appl. Phys. 8 (18), 2113-2122 (1975).
  46. Inguscio, M., Ioli, N., Moretti, A., Strumia, F., D'Amato, F. Heterodyne of optically pumped FIR molecular lasers and direct frequency measurement of new lines. Appl. Phys. B. 40 (3), 165-169 (1986).
  47. Carelli, G., et al. CH318OH: FIR laser line frequency measurements and assignments. Infrared Phys. Technol. 35 (6), 743-755 (1994).
  48. Pearson, J. C., Müller, H. S. P., Pickett, H. M., Cohen, E. A., Drouin, B. J. Introduction to submillimeter, millimeter and microwave spectral line catalog. J. Quant. Spectrosc. Radiat. Transf. 111 (11), 1614-1616 (2010).
  49. Ehasz, E. J., Goyette, T. M., Giles, R. H., Nixon, W. E. High-resolution frequency measurements of far-infrared laser lines. IEEE J. Quantum Elect. 46 (4), 474-477 (2010).
  50. Pearson, J. C., Drouin, B. J., Yu, S., Gupta, H. Microwave spectroscopy of methanol between 2.48 and 2.77 THz. J. Opt. Soc. Am. B. 28 (10), 2549-2577 (2011).
  51. Consolino, L., et al. Phase-locking to a free-space terahertz comb for metrological-grade terahertz lasers. Nat. Commun. 3, Article No. 1040 (2012).
  52. Bartalini, S., et al. Frequency-comb-assisted terahertz quantum cascade laser spectroscopy. Phys. Rev. X. 4 (2), 021006 (2014).
  53. Finneran, I. A., Good, J. T., Holland, D. B., Carroll, P. B., Allodi, M. A., Blake, G. A. Decade-spanning high-precision terahertz frequency comb. Phys. Rev. Lett. 114 (16), Article No. 163902 (2015).
  54. De Natale, P., et al. Quantum cascade laser THz metrology. Proc. SPIE.. 9370 Quantum Sensing and Nanophotonic Devices XII, , 93701D (2015).
  55. Dickinson, J. C., Goyette, T. M., Waldman, J. High resolution imaging using 325 GHz and 1.5 THz transceivers. 15th International Symposium on Space Terahertz Technology Proceedings, , 373-380 (2004).
  56. Vasconcellos, E. C. C., Zerbetto, S. C., Holecek, J. C., Evenson, K. M. Short-wavelength far-infrared laser cavity yielding new lines in methanol. Opt. Lett. 20 (12), 1392-1393 (1995).

Tags

Far-infrared Laser Emissions Measurement Frequencies Optically Pumped Molecular Laser Three-laser Heterodyne Technique Carbon Dioxide Lasers Reference Frequencies Beat Frequency Metal-insulator-metal Point Contact Diode Detector Spectrum Analyzer Uncertainty Reference
Characterizing Far-infrared Laser Emissions and the Measurement of Their Frequencies
Play Video
PDF DOI DOWNLOAD MATERIALS LIST

Cite this Article

Jackson, M., Zink, L. R.More

Jackson, M., Zink, L. R. Characterizing Far-infrared Laser Emissions and the Measurement of Their Frequencies. J. Vis. Exp. (106), e53399, doi:10.3791/53399 (2015).

Less
Copy Citation Download Citation Reprints and Permissions
View Video

Get cutting-edge science videos from JoVE sent straight to your inbox every month.

Waiting X
Simple Hit Counter