Physiological monitoring: Gasses

Published on 27/02/2015 by admin

Filed under Anesthesiology

Last modified 27/02/2015

Print this page

rate 1 star rate 2 star rate 3 star rate 4 star rate 5 star
Your rating: none, Average: 0 (0 votes)

This article have been viewed 1846 times

Chapter 15 Physiological monitoring

Gasses

Gas analysis during anaesthesia requires continuous monitoring of respired gasses and at times, intermittent sampling of blood gasses. The different techniques described in this chapter utilize various physical or chemical properties of the gas molecules, to detect and quantify the gas. As with all clinical measurement techniques, it is important to understand the principles on which the gas analyzers are based, so that their applications and limitations are recognized.1

Respiratory gas sampling

In a clinical setting, respiratory gas sampling and analysis is important for a number of reasons: ensuring that the gas supply to the patient contains adequate oxygen and volatile anaesthetic agent and ensuring adequacy of ventilation by capnography, which also gives some information about the circulation.2

A number of factors may affect or complicate gas sampling. Common to all methods is the delay in the sample reaching the analyzer and the response time of the analyzer itself. Also, not all analyzers return the sample to the breathing system. This is advantageous when the gas analyzer alters the integrity of the gas molecule. However, if low fresh gas flows are being used in a circle breathing system, the non-return of gas samples being removed at 150 ml min−1 significantly increases the fresh gas requirement.

Following a step change in the gas concentration, delay in response time of the analyzer is due to two factors. The first is the delay time or transit time: the time it takes for the sample to get from the patient’s airway to the gas analyzer. The second is the response time or rise time of the analyzer. The response time is usually considered to be the time taken for an analyzer to respond to within 90–95% of an actual step change in gas concentration. A step change can be produced in one of three ways: by moving a gas sampling tube rapidly into and out of a gas stream; by bursting a small balloon within a sampling volume containing a gas sample; or by switching a shutter to a gas sample volume using a solenoid valve. Fig. 15.1 shows how these are related.

Most modern analyzers use side stream sampling, where the sampling tube takes the gas sample to the analyzer. It is important that only the recommended sample tubing be used for the analyzer concerned, because various types of tubing may absorb some of the gas mixture as well as water vapour to different extents; calibration of the analyzer will have allowed for this only with the recommended tubing. Gas analyzers sample gas at a rate of between 50 and 200 ml min−1. If the sampling rate is higher than this, or if the tubing is too long or too wide, the sampling waveform will be distorted, thereby reducing accuracy. The delay time depends to great extent on the sampling rate and on the length of the sampling tube, which should be as short as possible.

In trying to sample gasses at the end of expiration, it is important to sample as close to the patient’s trachea as possible, particularly where the tidal volume is small and the respiratory rate is high, as when anaesthetizing infants. Some systems have a sampling catheter that extends down the endotracheal tube, but this adds to airway resistance where that tube is of small diameter. Most systems, however, have a sampling port attached to the breathing system adjacent to the artificial airway. It is important that the software within the analyzer can detect minima and maxima in the respiratory waveform, and associate these with inspiratory and expiratory gas concentration values appropriately. It is still possible, however, for a gas sample, taken, for example, from the patient end of a coaxial Mapleson D breathing system, a type of T-piece, to give erroneously low end tidal readings, due to confusion between inspiratory and expiratory gas flows.

Some analyzers have used mainstream sampling instead, where the analyzer itself sits astride the artificial airway outside the mouth. This is a bulky addition to the airway but it eliminates transit time, and is reported to be more useful in detecting sleep apnoea than sidestream analyzers.3 There is also less of a problem dealing with water vapour in the sample, and the sample is not degraded by the analyzer. An example is the Hewlett Packard infrared CO2 analyzer, shown in Fig. 15.2. The sensor fits onto a sampling chamber inserted into the breathing system. The miniaturized unit contains a motor with filter wheel, an infrared (IR) source and detector, and IR transparent glass for the optical pathways. A second fully encapsulated optical window on the airway adapter provides a reference for calibration.

Gas concentration monitoring

Refractometry

Refractometry is a technique that detects the difference in refractive indices between two media, one of which contains the gas sample of interest (e.g. air and anaesthetic vapour), the other which is of otherwise identical constitution (e.g. air only). The refractive index of a medium is a measure of the ratio of velocity of light in a vacuum to the velocity of light in that medium. The molecules of the medium delay the transit of the light wave through it; the delay depends on the number of molecules and hence on the density of the medium; thus refractive index for a gas medium depends on its concentration, pressure and temperature. When a light beam passes through parallel slits whose width is of the same order of magnitude as the wavelength of the light, interference patterns are produced by light waves arriving in phase (bright fringe) and 180° out of phase (dark fringe) with each other. When two such sets of fringes are formed from light passing through gas samples with differential velocities, they are displaced relative to each other and the displacement between them can be measured as a difference in refractive indices.

The refractometer is included, not because it is frequently used clinically as a gas analyzer, but because it is an accurate standard of gas analysis against which others are compared. All gasses and volatile agents can be quantified using this method, since all possess the physical property of refractive index. Laboratory refractometers are used by vaporizer manufacturers to calibrate vaporizers. A diagram of the portable version of the refractometer (the Riken refractometer) is shown in Fig. 15.3. Light from a small light source is collimated into a beam, which is split into two parallel beams by prisms. One beam is passed through the reference chamber containing, say, air, and the other is passed through the measuring chamber containing air and another gas/vapour species, for example sevoflurane, whose concentration is to be measured. The beams are reflected back through the chambers to enhance the effect. Each beam is passed through a slit, producing interference fringes. As discussed above, the effective path lengths of the two beams differ and two sets of fringes are produced. These can be realigned by a knob controlling a vernier scale and the amount of realignment is taken as a measure of refractive index change from the reference sample, and thus of concentration of the gas under test.

Infrared absorption spectroscopy

The interatomic bond between dissimilar atoms of a molecule absorbs radiation in the infrared (IR) range. Thus, molecules such as nitrous oxide, carbon dioxide, water vapour and the volatile agents absorb IR light, but oxygen, nitrogen and helium do not. Polyatomic molecules of different species absorb IR radiation across a range of IR wavelengths and there is frequently overlap between species. However, different polyatomic species absorb maximally at characteristic wavelengths within the IR bandwidth, which means that it is possible to identify the gas molecule as well as quantify the gas concentration. The amount of absorption (A) of any radiation by any substance is governed by the Beer-Lambert Law, which links A to the intensity of the incident radiation Ii, the non-absorbed, transmitted radiation It, the extinction coefficient ε, the path length L and the concentration C of the absorbing substance, where:

image

Since CO2 analysis is important in anaesthesia, such devices are available in most anaesthetic locations in modern operating rooms. Fig. 15.4 shows the principle of the Luft or non-dispersive type of IR gas analyzer, which uses a differential technique to minimize error. An IR light source emits radiation at between 1 and 15 µm. A ‘chopper’ wheel, with small windows containing optical filters at the rim, rotates in front of the light source at 25–100 Hz to prevent the gas sample from overheating; additionally the filters allow the passage of IR light of a narrower bandwidth than the source itself, to match the wavelength of maximum absorption of the gas under study. This window might contain a 4.26 µm filter for CO2, a 3.3 µm filter for halothane, isoflurane and enflurane, and a 4.5 µm filter for N2O. The light passes through this to two chambers, a reference chamber and a sample chamber; the sample chamber contains the same background gas as the reference chamber, but additionally contains the gas under analysis. The light, which is not absorbed (the transmitted light) in each sample cell, is passed to a pair of air-filled detector chambers, separated by a diaphragm. Because different amounts of IR light are absorbed in the reference and sample chambers, different amounts of light are, therefore, transmitted to the detector chambers, heating the air in them differentially. Because the chopper wheel produces an oscillating IR source, the changing air pressure in the two detector chambers causes the diaphragm separating them to oscillate. The diaphragm usually forms half of a capacitor and is, therefore, a component of electrical circuitry, which amplifies and processes the signal, to produce an output signal, which is proportional to the gas concentration.

There are a number of sources of error in IR spectroscopy. Although CO2, N2O and CO absorb IR light maximally at 4.3, 4.5 and 4.7 µm respectively, there is considerable overlap in the absorption spectra of these gasses, particularly in the 3–5 µm range. There is also the phenomenon of collision broadening, where the presence of one gas may broaden the IR absorption spectrum of another. This effect is not only caused by IR absorbing agents themselves (such as CO2 and N2O) on the absorption spectra of other IR absorbing agents, but by other non-IR absorbing carrier gasses, such as helium, argon or hydrogen. It has been reported, for example, that in a gas mixture containing 79% helium in oxygen, an IR analyzer under-reads CO2 values.4 Desflurane,5 cyclopropane,6 acetone and alcohol7 all produce errors in IR spectroscopy of respiratory gasses. There are electronic correction factors in the analyzers to allow for the presence of N2O in the gas mix containing CO2. Since water is a strong absorber of IR across the bandwidths of interest, it must be eliminated in the sampling process, in order to avoid some inaccuracy. If one wavelength is used for all volatile agent detection, then that agent must be selected manually, or the device will be inaccurate, if either the wrong agent is selected or if there is a mixture of agents present. Some analyzers use the 10–13 µm bandwidth to detect volatile agents, where there is less chance for interference between absorption spectra of the volatile agent and other gasses.

Infrared spectroscopy detects numbers of molecules in a gas sample and thus constitutes a partial pressure analyzer. Therefore, error can be introduced if the pressure of the gas sample changes or if there is ambient pressure change, against which the device is calibrated. If the gas sample pressure changes due to back pressure from a ventilator, or if there is significantly low pressure in a sampling tube, the partial pressure of the gas being analyzed will change, without there being a real change in the fractional concentration of the gas. Similarly, if the device is calibrated at sea level and subsequently used at altitude, there will be an error in calculating gas concentration. Therefore, calibration should be carried out frequently and under the conditions of intended use, or the device should be configured to display partial pressure rather than concentration.

The 90–95% response time of IR analyzers to a step change, including transit time and rise time, should be under 150 ms. Water vapour blocking the sampling tube sometimes causes the response time to increase. To deal with water vapour, most devices have either a water trap or use sample tubing, which absorbs water vapour. IR capnographs are accurate to about 0.1% in a range of CO2 up to 10%.

A variant of the IR analyzer described above utilizes a combination of IR and photo-acoustic spectroscopy. The gasses under test are drawn through a small measurement chamber that is irradiated by an IR source. This IR source is first ‘chopped’ by a rapidly rotating wheel that has three concentric bands of holes that cause the IR energy passing through the holes to pulsate at three different sampling rates. Each part of the now divided beam passes through a separate narrow band optical filter allowing through only that wavelength that corresponds to the maximum absorption of the gas/vapour to be tested (i.e. CO2, N2O and volatile agent). The components of the IR light, now differentiated both by sampling frequency and wavelength, are absorbed by the gas molecules, which are caused to vibrate at different frequencies, leading to a set of audible, fluctuating pressure waves. The families of audible pressure waves are detectable by microphone. Fig. 15.5 shows the Bruel and Kjaer model of the photo-acoustic IR spectroscope. Note that, since oxygen does not absorb IR light, this device has a separate sampling channel that uses magneto-acoustic spectroscopy to measure oxygen (see the section on Paramagnetic oxygen analyzers).

The signal from all gaseous components is taken from the microphone and filtered electronically into the four components of O2, CO2, N2O and volatile agent. Separation and, therefore, gas identification occurs by audio frequency identification and quantification of the gas occurs by audio-amplitude. Note that the device does not distinguish between different volatile agents. The advantages of photo-acoustic IR spectroscopy over conventional IR spectroscopy include stability, zero drift, reduced need for calibration over prolonged periods and fast response time.

A more recent design of infrared absorption spectroscopy does away with the moving parts of the chopper wheel. Light from a broad-spectrum infrared source passes through a cuvette containing sampled gas. The Dräger ILCA (Fig. 15.6) then uses a prism as a beam splitter. On the four sides of the multichannel detection chamber are appropriate infrared filters and sensors for the maximal absorption wavelengths of each gas. By using two such modules in series the gas bench can quantify CO2, N2O and the individual volatile agents in the gas mixture.

Mass spectrometry

Buy Membership for Anesthesiology Category to continue reading. Learn more here