Vous êtes sur la page 1sur 9

SPECTROPHOTOMETRY / Turbidimetry and Nephelometry 343

possible as the differentiation is dened unchangeable by the resistor and the capacitor of the differentiating circuit [7]. Analogous to the adjustment of the window width, different modulation amplitudes a [8] can be chosen in optical derivative spectroscopy. According to [12], larger derivative signals and hence better signalto-noise ratio can be gained by increasing a. This was conrmed experimentally; however, a signal increase comes along with a washing out of derivative shapes and minute derivative features are lost (Figure 7). The values for the modulation amplitude given in Figure 7 are typical examples. Thus, the modulation amplitude can be optimized for different applications: if several analytes in mixtures have very similar and overlapping absorption spectra, a small modulation amplitude helps to pronounce minor differences for discrimination. If, however, clearly different absorption spectra are present, a large modulation amplitude is selected for improved signal-to-noise ratio.
See also: Chemometrics and Statistics: Multivariate Calibration Techniques. Optical Spectroscopy: Radiation Sources; Wavelength Selection Devices; Detection Devices. Spectrophotometry: Overview.

Further Reading
Bosch OC, Sanchez RF, and Cano PJ (1995) Recent developments in derivative ultraviolet/visible absorption spectrophotometry. Talanta 42: 11951214. Brown C, Vega-Montoto L, and Wentzell P (2000) Derivative preprocessing and optimal correction for

baseline drift in multivariate calibration. Applied Spectroscopy 54: 10551068. Dixit L and Ram S (1985) Quantitative analysis by derivative electronic spectroscopy. Applied Spectroscopy Review 21: 311418. Fell A (1983) Biomedical applications of derivative spectroscopy. Trends in Analytical Chemistry 2(3): 6366. Hager R and Anderson R (1970) Theory of the derivative spectrometer. Journal of the Optical Society of America 60: 14441449. Hawthorne A and Thorngate J (1978) Improving analysis from second-derivative UV-absorption spectrometry. Applied Optics 17: 724729. Knowles A and Burgess C (1984) Practical Absorption Spectrometry/Ultraviolet Spectrometry. New York: Chapman & Hall. OHaver TC and Begley T (1981) Signal-to-noise in higher order derivative spectrometry. Analytical Chemistry 53: 18761878. Sassenscheid K, Klocke U, Marb C, et al. (1998) Dynamic derivative UV-spectroscopy for combustion monitoring. Proceedings of SPIE 3535: 204214. Sassenscheid K, Klocke U, Marb C, et al. (1998) Enhanced selectivity and sensitivity in UV-analysis of volatile organic compounds. Proceedings of SPIE 3533: 222233. Savitzky A and Golay M (1964) Smoothing and differentiation of data by simplied least squares procedures. Analytical Chemistry 36: 16271639. Steiner J, Termonia Y, and Deltour J (1972) Smoothing and differentiation of data by simplied least square procedure. Analytical Chemistry 44: 19061909. Vogt F, Klocke U, Rebstock K, et al. (1999) Optical UV derivative-spectroscopy for monitoring gaseous emissions. Applied Spectroscopy 53: 13521360. Williams D and Hager R (1970) The derivative spectrometer. Applied Optics 9: 15971605.

Turbidimetry and Nephelometry


D M Lawler, University of Birmingham, Birmingham, UK
& 2005, Elsevier Ltd. All Rights Reserved.

Introduction
Turbidity is an expression of the optical property of a medium, which causes light to be scattered and absorbed rather than transmitted in straight lines through the sample. The International Organization for Standardization (ISO) denes turbidity as the reduction of transparency of a liquid caused by the presence of undissolved matter. It is, therefore, the opposite of clarity. The medium concerned is normally a uid (but may be solid) in which light is

scattered by matter usually small particles suspended in the light path. Measurements of turbidity can be used in many analytical elds to determine the mass concentration of suspended particles in a sample and, for some simple contexts, particle size distributions. The eld is hampered, however, by a lack of standardization in units, measurement devices and calibration techniques. Analytical determinations of concentrations tend to be empirical. Such methodological problems have recently driven a profusion of technical papers. This article reviews turbidity theory, measurement principles, instrumentation systems, and applications, with particular reference to suspended sediment concentrations (SSCs) in natural waters (e.g., rivers, estuaries, and nearshore zones).

344 SPECTROPHOTOMETRY / Turbidimetry and Nephelometry

Definitions and Applications


Measurement Definitions

components turbidity apportionment has advanced with recent technological developments.

Turbidity can be measured using the techniques of turbidimetry or nephelometry (from nephelo cloud (Greek)). Turbidimetry is the measurement of turbidity by quantifying the degree of attenuation of a beam of light of known initial intensity. It is usually applied to media of fairly high turbidity in which the scattering particles are relatively large (e.g., natural waters), for reasons, which will be addressed below. Nephelometry is the measurement of turbidity by the direct evaluation of the degree of light scattering taking place in the medium. It is much more appropriate to media of lower turbidity in which the suspended particles are small. Turbidimetry and nephelometry can offer considerable time-saving advantages over gravimetric methods for the determination of particle concentrations, and are nondestructive techniques.
Typical Applications

Principles and Theory


Light passing through a liquid medium may be scattered and absorbed by inhomogeneities in the light path, especially suspended particles of silt, clay, algae and other plankton, microbes, organic matter, and other ne insoluble particulate substances. Bubbles and density discontinuities can also scatter light. Scattering occurs when a minute particle interacts with incident light by absorbing the light energy and then, as if a point light source itself, reradiating the light energy in all directions. Absorption takes place when light is converted to other energy forms (e.g., heat) within the particle. Scattered light includes that reected from the surface of the particle and that refracted within the particle, possibly after many internal reections. Scattering is often accompanied by absorption. The direct relationship, however, between turbidity data and suspended solids concentrations is weakened by the complex interactions of light energy with suspended particles. This interplay is heavily dependent on many factors, including:
*

Turbidimetry and nephelometry have found many applications in scientific laboratories and in the chemical, pharmaceutical, foodstuffs, and beverage industries. In addition, turbidimetry and nephelometry are well-established procedures wherever ltration processes have to be effected, monitored, and controlled. Within the hydrological sciences, and water supply and wastewater management industries, turbidity values can act as simple and convenient surrogate measures of the concentration of suspended solids, sulfate ions (which are precipitated as BaSO4 in acidic media (HCl) with barium chloride), and other particulate material, and remain one of the most common applications of turbidimetry. Also, atmospheric and space physicists effect nephelometric analyses because of the importance of dust particles to radiation and other processes. In quantitative chemical and biological analysis, applications are common, especially the calculation of absolute molecular weights and dimensions of polymers in solution, as well as particle size determinations of suspended matter. Chemical proles can also be obtained by observing turbidity changes deliberately induced by the addition of specific substances to the solution. Within microbiology, cell and bacteria growth can be monitored through the media turbidity changes such activity causes. In foodstuff manufacturing, turbidimetry is often used to monitor product quality and treatment process efciency, especially in the dairy and brewing industries. Clarity (and hence turbidity) is also a key concern in the petrochemical industries. Determining turbidity

* *

* * *

concentration of scattering particles suspended in the medium; size distribution of the scattering particles; shape, orientation, and surface condition of the scattering particles; refractive index of the scattering particles; refractive index of the suspension medium; wavelength of the light source employed.

Consequently, separate bodies of theory have been developed to describe the many different processes that result. At its simplest level, light intensity is reduced during transmission through a collection of scattering particles in a sample according to an attenuation function of the form:
I I0 etl 1

where I0 is the initial beam intensity, I the beam intensity after passing through a medium of length l, and t the turbidity coefcient of the medium. Equation [1] ignores losses of light through true absorption by suspended particles or reection from the sides of the sample container.
Light-Scattering Theory and the Inuence of Particle Size

Appropriate light-scattering theory is governed by the diameter, D, of the scattering elements in relation

SPECTROPHOTOMETRY / Turbidimetry and Nephelometry 345

Incident beam

Incident beam

Incident beam

(A) Small particles

(B) Large particles

(C) Larger particles

Figure 1 Inuence of particle size on the angular distribution of scattered light: (A) small particles (D o 0.1l); (B) large particles (D B 0.25l); and (C) larger particles (D 4 1l). (From Vanous RD, Larson PE, and Hach CC (1982) The theory and measurement of turbidity and residue. In: Minear RA and Keith LH (eds.) Water Analysis, vol. 1, pp. 163234. New York: Academic Press.)

Light absorbance (i c )

to the wavelength, l, of the light emitted by the measuring instrument. Indeed, theory is often specied in terms of the Mie size parameter a 2pR=l, where R is particle radius. Particle size thus forms an appropriate basis for the subdivision of the theoretical discussion that follows.
Small Particles

1.0
<5 m

0.3
68 m 1218 m 3050 m

0.1

For particles where Do0.05l, Rayleigh scattering theory of 1871, originally developed for gases, is applicable to liquids with low concentrations of suspended particles which do not interact with each other. For such small particles, relatively symmetrical light-scattering distributions are obtained (Figure 1A). If a visible light source (i.e., l 0.40.7 mm) is employed, then it follows that this theory is applicable for particles where DoB0.03 mm. The Rayleigh equation describing the angular distribution of resultant scattering is:
iy =I0 fn0 =n 1g2 NV 2 =l4 r2 1 cos2 y 2

0.03 30 100 300 1000 3000 10 000 Suspended sediment concentration (mg l1)
Figure 2 Inuence of sediment particle diameter on light absorbance by samples of different concentrations. Note how a given concentration effects much greater absorbance at the smaller particle diameters. (Reproduced with permission from Ward PRB and Chikwanha R (1980) Laboratory measurement of sediment turbidity. Proceedings of the American Society of Civil Engineers, Journal of the Hydraulics Division 106: 10411053.)

where iy is the intensity of light scattered at angle y, I0 is the initial light source intensity, n0 is the refractive index of the particles, n is the refractive index of the suspension medium, r is the distance from the particles to the point of measurement, in terms of the number, N, of particles, each of volume V. Rayleighs work thus shows that the intensity of scattered light varies: (1) with the square of the particle volume and thus with the sixth power of the particle radius, assuming spherical shapes; and (2) inversely with the fourth power of the light wavelength used. Rayleigh theory has since been developed to allow relative molecular masses and sizes to be determined.
Large Particles

particles, scattering intensity is less dependent on wavelength.


Very Large Particles

For larger particles, however, where 0.1loDo0.8l, the angular distribution of scattered light becomes asymmetrical. Destructive interference of light scattered in the backward direction leads to a bias in forward-scattered light (Figure 1B). In these contexts, Mie scattering theory of 1908 for larger spheres becomes more appropriate. For such larger

For larger particles still, where D40.8l, the Mie equations are still workable, although for particles larger than B0.4 mm in diameter wide oscillations in scattering patterns emerge. For particles of D41 mm, extreme concentration of scattering in the forward direction emerges (because of mutually destructive backscattering effects), along with secondary peaks in the angular distribution of scattered light (Figure 1C). Theory and practice also demonstrate that the most efcient scattering elements are those of a diameter similar to the light wavelength used. Also, a given mass of small particles causes much greater light attenuation than the same mass of large particles (Figure 2). One complexity is that much classical theory has been developed for identically sized spherical particles conditions that may not be obtained in all laboratory or eld situations. Indeed, many natural

346 SPECTROPHOTOMETRY / Turbidimetry and Nephelometry

waters, like the atmosphere, contain an ensemble of variably sized, irregularly-shaped, and randomly oriented particles, for which theory is still being developed. Furthermore, processes become highly complex when concentrations are so great that multiple scattering occurs (i.e., particles receive light previously scattered from other particles: this normally increases opportunities for light absorption).

5.9

Photocell

Instrumentation
Range of Turbidimetric and Nephelometric Systems

Lamp (A)

Lens Sample cell (vertical view)

Early procedures were based on manual operation of analytical systems and visual turbidity assessment (e.g., the Secchi disk). Several instruments are now available, however, for quantitative turbidity determination in a variety of scientific, industrial, and process management applications. Choice will depend largely on the analytical aims (e.g., mass concentration, particle size distribution, molecular dimensions, or crystal/cell growth), the nature of the scattering elements and suspension medium, and whether eld or laboratory measurement is needed. Turbidimeters also vary in optical geometry, mode of operation, sample handling capabilities, data recording options (e.g., automatic/manual or analogue/ digital) and portability. Ultraviolet-visible spectrophotometry can also be used for turbidimetric measurements by measuring the absorption of light by particles at a xed wavelength or full spectrum (e.g., for kinetic studies of the time decay of species). The discussion that follows focuses on the use of turbidimetric instrumentation to estimate the mass concentration of suspended matter in liquid samples, with particular reference to sediment in natural waters.
Basic Elements of Measurement Systems

Transmitted detector Lamp (B) Filter Lens Sample cell

90 Detector

Forward scatter Transmitted detector

Lamp (C) Filter Lens

Sample cell

Figure 3 Three basic designs of turbidity meter: (A) the nephelometer, which directly measures light scattered (usually at 901 to the beam direction) by suspended particles; (B) the turbidimeter, where the transmitted light is detected, in relation to initial beam intensity; (C) the ratio turbidimeter in which both transmitted and scattered light is detected. (Reproduced with permission from Hach CC, Vanous RD, and Heer JM (1982) Understanding turbidity measurement. Technical Information Series, Booklet No. 11, 1st edn., 11pp. Hach Chemical Co.)

Modern measurement devices use photosensitive cells to quantify scattered and/or transmitted light. Figure 3 illustrates that most laboratory bench instruments usually have ve basic components: a light source of known, constant intensity, and given wavelength characteristics; a lens to collimate the light beam; a sample cell; photosensor(s); and a meter or logger to record the output signals from the photosensor(s). Versions for continuous monitoring of turbidity values (e.g., for online industrial systems or process measurement in environmental sciences) include some kind of ow-through measurement chamber (instead of a static sample cell) and outputs for a datalogger. The two basic measuring instruments are the nephelometer and the turbidimeter.

A nephelometer measures directly the intensity of light scattered by the sample, which is proportional to the amount of matter suspended in the light path, though the inuence of size, shape, and refractive index of the scattering particles is also important. With nephelometers, the sensor is mounted at an angle to the traversing beam (often 901) to record scattered light in one part of the angular distribution (Figure 3A). Some more sophisticated versions can monitor scattering intensity at many different angles: this allows angular summation values to be checked against initial and attenuated signals. Nephelometers usually provide better precision and sensitivity than turbidimeters and are normally used for samples of low turbidity containing small particles. A turbidimeter, sometimes called a transmissometer, absorptiometer, or turbidity meter (the latter term is commonly used for eld instruments in the earth and environmental sciences), measures the

SPECTROPHOTOMETRY / Turbidimetry and Nephelometry 347

intensity of the beam after it has passed through the sample, i.e., it quanties the amount of transmitted light remaining (Figure 3B). Suspended matter in the light path causes scattering and absorption of some light energy, which reduces the incident illumination falling on the photocell. These instruments are more appropriate for relatively turbid samples in which the scattering particles are large in relation to the light wavelength used. This is because a signicant reduction in the intensity of incident light is needed to yield precise results. Some newer instruments, called ratio turbidimeters, incorporate measurement systems for light which is side-scattered (usually at 901), forwardscattered, and transmitted (Figure 3C). The turbidity value is obtained as the ratio of the 901 signal to the sum of forward-scattered and transmitted values. The ratio feature has a number of advantages: it increases the long-term stability of the sensor (by reducing effects of instrumental drift); it compensates for ageing of, and deposits on, the optics; it reduces the inuence of temperature changes in the electronics; it minimizes the need for repeated recalibration; and it limits the effect of sample color on readings. This can be more appropriate for strongly and/or variably colored liquids, or for samples of high turbidity. A four-beam instrument version has emerged recently, which reduces error still further. Recent developments include laser-based turbidimeters, reectometers, or ber-optic systems. The development of the optical backscatter sensor (OBS) has become popular for eld deployment in the hydrological and oceanographic sciences: this instrument monitors water turbidity through the backscattering of pulsed infrared light emitted from the OBS instrument head. Also, remotely sensed turbidity measurement, using satellite or air borne instruments (e.g., the CASI (Compact Airborne Spectrographic Imager) system deployed by the UK Natural Environment Research Council), has recently eased the mapping of turbidity patterns over large spatial scales. There is a strong dependency of scattering efciency on light wavelength (see above). Consequently, for a given detector, light sources of short wavelength are more sensitive to, and therefore more useful for, the detection of small particles. Conversely, longer wavelengths are more appropriate for samples containing large particles (e.g., sediment in many earth or environmental science systems). The sourcedetector relationship can vary widely between instruments, and is cited as the key reason explaining the different readings obtained on the same sample by different devices.

Units of Measurement and Instrument Calibration

The eld is hampered by a nonstandard, ill-dened, and historically changing unit of measurement. The Nephelometric Turbidity Unit (NTU) is the most common unit employed. The precision with which turbidity data should be reported depends on how turbid the sample is, but should be to the nearest 110%, approximately, of the NTU value determined. For example, NTU values for distilled water, tap water, and raw water are 0.08, 0.54, and 3.52, respectively, but much higher values, well above 150 NTU, are common in many hydrological systems. Formazin polymer, developed in 1926, can be used for turbidimeter calibration, and is straightforward to prepare, control, and reproduce. Standard procedures for the production of a stock formazin turbidity suspension of 400 NTU are given in American Public Health Association. Other calibration materials can be used (e.g., Fullers Earth or Hach Gelex xed standards metal oxide particles permanently and statically suspended in silica gel) and may provide suitable alternatives, especially given the health concerns voiced in some quarters over formazin use.

Field Calibration

In natural waters, suspended material may largely consist of particles in the size range of clay (Do2 mm), silt (2oDo63 mm), or even sand (63oDo2000 mm). It may also include organic matter and compounds and microscopic organisms. For eld applications in hydrology or oceanography, analysts should preferably calibrate turbidity readings against known mass concentrations of the suspended sediment typical of that context, and declare the strength of the diagnostic statistics for derived relationships. Such correlations can be weak, reecting temporal changes in suspended load composition (and hence its light-scattering efciency), water color, or bubble presence. Predictive relationships can be strengthened by accounting for such changes (especially in sediment load particle size distribution), which can occur over various timescales (e.g., interannual, seasonal, subseasonal, ood event). It may even be necessary to produce multivariate or separate calibration equations to incorporate the effects of, for example, changing ow levels, sediment source areas, and season on sediment load constitution. For such eld applications, the turbidimeter reading (often in arbitrary units) is converted to estimated SSC using a site-specific calibration curve.

348 SPECTROPHOTOMETRY / Turbidimetry and Nephelometry Turbidity Meters for Continuous Field Operation
400
HF DRT-1000

For unattended automated eld use a range of suitable turbidity meters with additional instrumental features are available. The measurement cell is replaced by a ow-through chamber, which must prevent stray light from reaching the photosensor. Instruments can be boom-mounted directly in the ow, or at the end of pump-lines connecting sampling point to measurement system. Turbidity meters with narrow-band near infrared (NIR) light sources (peak output at 0.86 mm; spectral bandwidth o0.06 mm) are recommended by ISO. Such instruments reduce problems of algal build-up on the optical surfaces, are less affected by color, and are more sensitive to the slightly larger particles typical of sediment transport systems. However, some relaxation of the infrared protocol is tolerable for eld instruments operating in continuous monitoring mode. To limit further the impact of problematic algal growth on the optics, eld turbidimetric systems can be equipped with a pulsed light source, antifouling chemicals or lms, and/or wiper blades for convenient (sometimes automated) cleaning of optical surfaces. Alternatively, a dual-beam (twin-gap) instrument to compensate for these effects can be deployed. Power requirements for eld instruments are important considerations. The use of low-consumption light-emitting diodes (LEDs) in sensors, and photovoltaic sensors which convert incident light directly into electrical energy, reduces power needs to a minimum. Solar panels are useful to trickle-charge instruments and dataloggers. Some instruments are temperature sensitive, largely because LEDs can emit more strongly when warm, and photovoltaic detectors convert photons into electrons more efciently at low temperatures. Given, for example, the annual range of river temperature in the UK is typically B2025 K, some correction procedures may be necessary. Figure 4 also demonstrates that many turbidi meters are relatively insensitive to very ne or very coarse particulate matter. Because most instruments work in the visible or NIR spectrum this means that the most readily detectable particles are those where, B0.2oDo1.8 mm. In standard sedimentological and engineering classications, these are clay-sized particles, and hence the occasionally used term siltmeter for turbidimetric instruments is not entirely appropriate. Turbidity meters need to be interfaced with portable multimeters, dataloggers, or computers for recording and storage of turbidity data. Dataloggers

Instrument response (NTU)

300

200
Hach 2100A

100

5 103 wt. % conc.

0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 Particle diameter (m)
Figure 4 Sensitivity of two turbidity meters to particle diameter (HF H.F. instruments). Test material here is spherical latex particles of very narrow size distribution. (From Vanous RD, Larson PE, and Hash CC (1982) The theory and measurement of turbidity and residue. In: Minear RA and Keith LH (eds.) Water Analysis, vol. 1, pp. 163234. New York: Academic Press.)

equipped with an appropriate input channel and needing minimal power for operation are heavily used today. They provide quasi-continuous digital data on temporal variations in turbidity, which are ideal for computational analysis and the study of turbidity dynamics. Telemetry systems for real-time data acquisition, alarm facilities, and remote downloading capabilities are becoming increasingly common. Many eld scientists, however, despite the range of commercially available turbidity instrumentation, still recognize a need for low cost, rugged, and reliable systems. This is especially so when a network of instruments is required for permanent installation to dene a turbidity eld, for simultaneous manual turbidity measurements by a research team, or for specialist applications such as in subglacial environments. This has led many researchers to custombuild their own instruments.

Hydrological Applications of Turbidimetry


The monitoring of turbidity and SSCs in rivers, estuaries, lakes, reservoirs, nearshore zones, etc. is attracting increasing attention from hydrologists, limnologists, geomorphologists, freshwater ecologists, engineers, oceanographers, glaciologists, water resource managers, and policy makers. Such measurement programs can allow inferences to be made about upstream hydrogeomorphological processes,

SPECTROPHOTOMETRY / Turbidimetry and Nephelometry 349

catchment erosion rates, downstream uvial processes and sedimentation impacts, pollutant and contaminant transfer, and aquatic habitat quality. The recreational value of water bodies can be partly linked to their clarity, as demonstrated in the Lake Tahoe turbidity reports of 200102. Increasingly, there is a legal requirement for environmental impact assessments and water supply managers to consider the possibility of short- or long-term turbidity increases resulting from proposed development schemes. Automated, in-stream, high-frequency turbidity monitoring has become increasingly popular, mainly because the alternative practice of sampling and subsequent laboratory processing of samples is laborious and resource-intensive. Sampling approaches thus inevitably constrain the level of temporal and spatial detail possible, making it difcult to reveal the patterns, dynamics, and processes present. Recent advances in the understanding of the hydrodynamics of ne-sediment transport in river, tidal, and nearshore environments, for example, would have been impossible without very high frequency (e.g., 5 Hz) monitoring of transient turbidity changes. In remote environments, and in the developing world, a likely paucity of suitable sample analysis facilities underlines the need for an automated and direct eld-based method. Three example applications of continuous monitoring are outlined below: river turbidity variation during rainstorms; very short-lived turbidity pulsing in glacial and coastal waters; and the definition of the estuarine turbidity maximum. These illustrate the many advantages of turbidity instrumentation over sampling-based approaches in quantifying and understanding complex temporal and spatial patterns of suspended sediment uxes. They also demonstrate the substantial variations of turbidity and SSCs in natural systems in space and time.
River Turbidity Variation through Individual Storm Events

Concentration (mg I1) Discharge (Is1)

100

75

50
Sediment concentration

25
Discharge

0 12 18.11.1986 12 19.11.1986 12 20.11.1986

Figure 5 Continuous record of stream suspended sediment concentration, in relation to river discharge changes during storm events, derived from calibration of a Partech turbidity record. (From Gippel CJ (1989) The use of turbidity instruments to measure stream water suspended sediment concentration, Monograph Series No. 4, Department of Geography and Oceanography, University College, University of New South Wales, Australian Defence Force Academy, Canberra, 204pp.)

hysteresis effects in relation to the discharge series (hitherto undetected for the system by earlier workers using a conventional sampling program). This led, in turn, to more securely based explanations of ne-sediment delivery processes.
Turbidity Pulsing in Rivers and Nearshore Zones

SSC in streams can change appreciably over seasonal timescales and during high-ow events. Automated river turbidity monitoring is very useful in refining calculations of suspended sediment loads, because it detects the short-lived, but very important order-ofmagnitude changes in SSC that occur in many rivers during complex storm events (e.g., Figure 5). The 1-min datalogging scan interval used in Figure 5 dened all peaks and troughs in turbidity for most events. Following calibration, this permitted the full definition of exhaustion phenomena in SSCs, and

Turbidimetric instrumentation also facilitates the detection of very short-lived pulsing of suspended sediment, which characterizes many systems, especially proglacial meltwater environments. The example in Figure 6, for the Jo a So lheimasandi glacial river kulsa in southern Iceland, shows two substantial sedimentpulsing events detected by a 2-min turbidity scanning program. These were unrelated to river ow variations, and showed that other signicant sediment mobilization processes were present in the system. The alternative approach of ow-triggered automated sampling is unsatisfactory in these situations where many sedimentux perturbations are unrelated to water discharge. One strength of automated turbidity monitoring is that the logging system can also be used to record, on the same time base, data on associated or explanatory variables. For glacial meltwater studies (Figure 6), these typically include energy budget components relevant to glacial ablation, rainfall intensity, river discharge, temperature, and electrical conductivity, and, with Photo-Electronic Erosion Pin (PEEP) sensors, even the erosion events themselves which generate sediment plumes. Such data can strengthen the

350 SPECTROPHOTOMETRY / Turbidimetry and Nephelometry


100 5000 Suspended sediment concentration (mg I1) 20 10 0 10 20 v (cm s1) 8 9 10

80 Discharge (m3 s1)

Discharge

4000

60

3000

40 Concentration 20

2000

1000

0 18:00

20:00

22:00

00:00

0 02:00

GMT (also Icelandic time)


Figure 6 Two-minute scanning of turbidity and river discharge (89 Aug 1988) showing a compound pulsing of suspended sediment a So lheimasandi glacial river in southern Iceland. (Reproduced with kulsa concentration, unrelated to ow variations, in the Jo permission from Lawler DM and Brown RM (1992) A simple and inexpensive turbidity meter for the estimation of suspended sediment concentrations. Hydrological Processes 6: 159168; & John Wiley and Sons Ltd.)

OBS output (mV) 1700 1950 2200

OBS output

U (cm s1) 60 80 100 120

Time (min, from 16:57 h PST)


Figure 7 A high-frequency, 5 Hz, 10-min time series of OBS turbidity alongside river ow components (u, streamwise; v, normal to the bed) for the Fraser River, near Mission, BC, Canada. (Reproduced with permission from Lapointe M (1992) Burst-like sediments suspension events in the sand bed river. Earth Surface Processes and Landforms 17: 253270; & John Wiley and Sons Ltd.)

process-inference capabilities of the whole exercise. Thus, the correlation of high-frequency velocity and OBS turbidity series (Figure 7) allowed a clearer understanding of river sediment transport events to be gained. In coastal zones, very high frequency (5 Hz) monitoring of OBS turbidity, wave height, and currents (Figure 8), facilitated the definition of the critical ow velocities required to mobilize bed sediment. Knowledge of these threshold conditions is important for the stability, engineering, and protection of coastlines and their ecosystems.

Estuarine Turbidity Maxima

The estuarine turbidity maximum is the term given to the clear peak in mean SSC observable in many estuaries around the limit of saline intrusion. Improved explanations of turbidity elds in estuarine systems, including the nature, location, and migration of the turbidity maximum, and the tidal pumping processes responsible, have recently been obtained by supplementing water sampling approaches with detailed automatic turbidity monitoring. Furthermore,

SPECTROPHOTOMETRY / Inorganic Compounds 351


0.2 0 0.2 200 SSC Hs 0 50 100 150 200 250 300 350 400 Time (h) Finish 19.02.96 08:00 100 0

V (m s1)

0.5 0

U V

H S (m)

2 1 0

SSC (mg l1) 18

0.5 3

for any spatial survey, the fact that turbidity values are obtained at the eld sites themselves, rather than in the laboratory subsequently, can allow instant decisions to be made regarding any further environmental sampling (including turbidity) that may be desirable. Figure 9 illustrates the value of repeated estuarine turbidity measurement in revealing the spatial and temporal structure of the turbidity maximum.
See also: Color Measurement. Environmental Analysis. Geochemistry: Sediment. Particle Size Analysis. Sensors: Photometric. Water Analysis: Particle Characterization.

Start 04.02.96 12:00

Figure 8 A high-frequency, 5 Hz record averaged to hourly time series of OBS turbidity (SSC), in relation to signicant wave height, Hs; cross-shore current, U; and longshore current, V, for the North Sea nearshore zone at Holderness, UK. Data from Feb. 1996; water depth 16.8 m. The OBS here has been deployed within BLISS (Boundary Layer Intelligent Sensor System). (From Blewett J and Huntley D (1999) Measurement of suspended sediment transport processes in shallow water off the Holderness coast, UK. Marine Pollution Bulletin 37(37): 134143.)

U (m s1)

Further Reading
American Public Health Association (1998) Turbidity. Standard Methods for the Examination of Water and Wastewater. American Public Health Association, American Water Works Association, and Water Pollution Control Federation. Davies-Colley RJ and Smith DG (2001) Turbidity, suspended sediment, and water clarity: a review. Journal of American Water Resources Association 37(5): 10851101. ISO (International Organisation for Standardization) (1984). Water Quality Determination of Turbidity, 1st edn., 5pp. ISO 7027 1984(E). Mitchell SB, Lawler DM, West JR, and Couperthwaite JS (2003) Use of continuous turbidity sensor in the prediction of ne sediment transport in the turbidity maximum of the Trent Estuary, UK. Estuarine Coastal and Shelf Science 58: 643650. Schuerman DW (ed.) (1980) Light Scattering by Irregularly Shaped Particles, 334pp. New York: Plenum Press. Van de Hulst HC (1981) Light Scattering by Small Particles, 2nd edn., 470pp. New York: Dover Publications. Walling DE (1977) Limitations of the rating curve technique for estimating suspended sediment loads, with particular reference to British rivers. Erosion and Solid Matter Transport in Inland Waters. International Association of Hydrological Sciences Publ. No. 122, pp. 3448.

180 160 140 Turbidity (NTU) 120 100 80 60 40 20 0 0 2 4 6 8 10 12 14 Distance from mouth (km) 16 Top Bottom

Figure 9 Relationship between estuarine turbidity in NTU and distance from the mouth of St. Lucia Estuary, Natal, on a rising tide on 20 March 1981, showing a clear turbidity maximum B812 km from the estuary mouth for both near-bed (bottom) and surface (top) waters. (From Cyrus DP (1988) Turbidity and other physical factors in Natal estuarine systems. Part 1: selected estuaries. Journal of the Limnological Society of southern Africa 14(2): 6071.)

Inorganic Compounds
M A Zezzi-Arruda and R J Poppi, University of Campinas, Campinas, Brazil

Introduction
Spectrophotometry is an excellent alternative for the determination of inorganic compounds. It is characterized by a wide analytical working range,

& 2005, Elsevier Ltd. All Rights Reserved.

Vous aimerez peut-être aussi