Soil Organic Matter Evaluation

Rosell R.A., J.C. Gasparoni, J.A. Galantini. 2001. Soil organic matter evaluation. In Assessment Methods for Soil Carbon (Ed. R. Lal, J.M. Kimble, R.F. Follett y B.A. Stewart). Serie Advances in Soil Science, Chapter 21, pp. 311-322



METHODOLOGICAL ASPECTS OF SOIL ORGANIC MATTER EVALUATION
Rosell R.A., J.C. Gasparoni and J.A. Galantini
CIC (PBA), CERZOS (UNS-CONICET) Dpto. Agronomía -   UNS
8000 BAHIA BLANCA - ARGENTINA
juangalantini@gmail.com  or  juan.galantini@cyt.cic.gba.gob.ar 

Introduction

Soil organic matter (SOM) is a complex mixture of animal and plant residues, fresh and at all stages of decomposition, living and decaying microbial tissue or heterotrophic biomass, and the relatively resistant humic substances. It has a series of beneficial properties for soils, even at low percentage levels. Organic matter (OM) is one of the most complex, dynamic and reactive soil components. The various organic fractions have different properties which affect the soil behavior and fertility as a whole. Unfortunately, it is difficult to separate and quantify those fractions in order to be able to correlate soil constituents and plant productivity.
Carbon is the main element present in SOM, comprising from 48 to 60% of the total weight. Organic carbon determination is often used as the basis for SOM estimation by multiplying the OC value by a conversion factor. Many researchers use the so-called Van Bremmelen factor of 1.724 (see below). Actual evidence indicates that the factors most appropriate for converting OC to OM are 1.9 to 2.5 for surface and subsoil, respectively (Broadbent, 1953; Nelson and Sommers, 1982).
Direct determination of SOM by either dry and/or wet oxidation procedures includes several errors and, therefore, is not convenient. Several conversion factors which produce uncertainty must be utilized. Rasmussen and Collins (1991) have indicated that the conversion factor to transform % OC to % SOM varies between 1.4 to 3.3. For that reason it is suggested to use the “% of OC”, which is determined, instead of “% of SOM”, which must be estimated.
The wet oxidation procedure is based on the oxidation of OC in soil with a dichromate-sulfuric acid mixture, either without or with external heat in order to accelerate and complete the oxidation. It is known that these methods do not produce complete oxidation of the OC, requiring therefore a correction (oxidation) factor for calculating the results. In both groups of procedures the excess dichromate (Cr2O72-) is back titrated with ferrous ammonium sulfate by using one of the three redox indicators (diphenylamine, o-phenanthroline, or N-phenylanthralinic acid). The chemical reactions are :
2 Cr2O72-
+
3 Cº
+
16 H+

4 Cr3-
+
3 CO2
+
8 H2O
  Cr2O72-
+
6 Fe2+
+
14 H+

2 Cr3-
+
6 Fe3+
+
7 H2O

Soil laboratories normally measure SOM  by the wet combustion method of Walkley and Black (Allison, 1965), assuming that the procedure measures easily oxidizing OC. A correction (oxidation) factor (1/0.76=1.33) is used to multiply the partially-oxidized OC content to convert this value to total organic carbon (TOC). Another factor (1.724) converts OC to OM. This conversion factor of 1.724 is the result of attributing 58% to the content of OC in the OM (Tabatabai, 1996). In the originally proposed method of Walkley and Black, the authors recovered an average of 76% of the OC in 20 soil samples, but recovery values ranged from 60 to 86%. Other authors, reported by Shulte and Hopkins (1996), recovered 49 to 96% (average 78%) of the OC in 96 Wisconsin soils by this method.
Native, virgin soils decrease their carbon content and negatively change their physical, chemical and biological properties upon continuos cultivation.
Prairie soils also lose SOM rapidly after they are first cultivated (Miglierina et al., 1988). The loss is usually exponential, declining sharply during the first 10 to 20 years. Later losses continue declining until an equilibrium is reached after 50 to  60 years (Jenny, 1941).
Dalal and Mayer (1986) have found that after cultivation OC bound to sand-sized particles declined rapidly; that associated with silt-sized particles increased from 48 to 61% (Sikora et al., 1996). The OC differential behavior when it is associated with or is present in several soil particle size fractions is a clear indication of the importance of physically-fractionating SOC.
It has been recognized that the interaction of SOM fractions with clay minerals is an essential factor in the behavior of the organic fractions (Christensen, 1991). Soil aggregate fractionation is based on the principle that the fractions of SOM are associated and/or bound to inorganic particles of different sizes, which differ in structure and function, playing important and differential roles in SOM turnover. Size fractionation is mainly applied to whole soil samples. Soil samples are usually dispersed in water and then separated by wet sieving. The procedure is considered to be very efficient because of the solution and mechanical effects of water. In sandy soils dry sieving is commonly used.
In general, most of the humified SOM in agricultural soils is found associated with the silt- and clay-sized particle fractions, which are called the fine fraction. The clay fraction can contain more than 50 % of SOM (Tiessen and Stewart, 1983 ; Bonde et al., 1992). Others authors (Whatson and Parsons, 1974 ; Elustondo et al., 1990) have reported that the silt-size fraction appeared to have the highest proportion of SOM.
Recently, Galantini and Rosell (1997) studied the evolution of SOC, N, P, and S in two granulometric fractions of an Entic Haplustoll under several crop sequences after 10 years of cultivation. They used wet sieving and obtained the denominated fine fraction (FF, < 100 µm particle size) and coarse fraction (CF, 100 - 2000 µm particle size). They determined that the level of some SOC fractions associated with the FF was maintained after cultivation. On the other hand, mineralization of OM was more intensive in the coarse fraction, thus showing its effects on the changes and dynamics of SOC and plant nutrients. Knowledge of the SOC distribution, mainly the composition of the CF, may therefore predict the OC and nutrient turnover and the present crop-soil fertility status.
The concept behind physical fractionation of soils emphasizes the role of soil minerals in SOM stabilization and turnover. The physical fractionation techniques are considered chemically less destructive, and the results obtained from physical soil fractions are considered to be related more directly to the structure and function of SOM in situ.
The different SOM fractions and their transformation rates require a knowledge of their composition and functions. In this paper, SOM definitions, extraction or sampling, fractionation (physical, biological, and chemical), and determination methodologies as well as their agronomic implications are discussed.

Definitions

Different SOM definitions have been used by numerous authors throughout the years. Historically, chemical concepts and procedures have primarily influenced the methodology employed for SOM research, mainly the knowledge of the chemical structure and properties such as elemental composition, cation exchange capacity (CEC), total and/or partial acidity, nutrient availability and/or sequestration, etc. in the humified fractions. In general, chemical characterization of OM and agronomic responses have not been mutually correlated.
A present concept of SOM components or composition can be defined as follows (whole soil < 2 mm) :
· organic residues ( 53µm or 100-2000 µm size), relatively fresh residues (Magdoff, 1996), decaying plant and animal tissues or debris and their partial decomposition products, including compounds of different biochemical origin ( amino acids, carbohydrates, fats, waxes, organic acids, etc.) present in decaying tissues. This compartment is also defined as the organic matter of the coarse fraction (CF) of the soil (Andriullo etal., 1990). Cambardella and Elliot (1992) have defined this fraction as particulate organic matter (POM). (fine or coarse POM, according to the authors: Galantini et al., 2014, 2016; Duval et al., 2013, 2016, 2018)
· soil biomass (microbial size), the SOM present as live microbial tissues, which carry out important functions related to residues decomposition, nutrient cycling, pest controls, etc. and also promoting porosity channels (Magdoff, 1996).
· the OM compartments of the soil biomass and the HS associated to the less than 53 µm (or 100 µm) size is also called the fine fraction (FF) of the SOM.
· humic substances (< 53 µm size), “ a series of relatively high-molecular-weight, brown to black colored substances formed by secondary synthesis reactions” (Stevenson, 1982). According to their molecular size and their solubility in alkali and acid aqueous solutions, well decomposed humic substances (HS) include :
·  humin, the alkali insoluble fraction of the HS, formed possibly by the condensation of humic and fulvic acids and residual lignin components.
·  humic acids, the dark-colored organic material insoluble in acid media but soluble in alkali solutions and specific organic solvents
·  fulvic acid fraction, the yellow-reddish (depending on pH), water soluble, colored material which remains in solution after precipitation of the humic acids by acidification
The agronomic or functional aspects of SOM and its fractions have been important in situations related to soil fertility and plant productivity. For over a century, SOM has been separated by chemical procedures (water or alkali - acidic extractions) to be used more in pedogenic studies than in agricultural applications. However, Feller (1993) considers that:
·                 humic substances have low turnover rate (Anderson and Paul, 1984; Duxbury et al., 1989) and are, therefore, not necessarily implicated in the short-term processes which occur in agronomic situations;
·                 the behavior of these chemical compartments in relation to major soil processes such as aggregation, mineralization and surface properties are not yet fully understood.
In the last 30 years, SOM characterization based on physical (particle size or density) separation (Feller, 1979; Tiessen and Stewart, 1983; Elliott and Cambardella, 1991; Christensen, 1992) has been intensified. Much evidence has accumulated to demonstrate that fractionation of soils according to particle size provides an useful tool for the study of SOM dynamics and distribution. A key point in these separations is to achieve adequate soil dispersion to prevent inclusion of microaggregates of smaller size particles in the silt and sand size separates. Chemical dispersants used regularly in the past may be too drastic, causing denaturalization of part of the SOM. Three different compartments of SOM have been separated from tropical soils (Feller et al., 1991):
·       plant debris (>20 µm size), with C/N >15;
·       organo-clay (0-2 µm size), amorphous organic material closely associated to clays, with C/N <10;
·       organo-silt (2-20 µm size), with properties intermediate between the former two and a C/N ratio ranging from 10 to 15.

FUNCTIONAL ORGANIC MATTER FRACTIONS

Normally, SOM is used as an indicator of soil quality. The most labile compartments (light carbon or bioactive C) of SOM are thought to be participating in supplying nutrients (N, P, S and minor elements) for plant growth. The most labile compartments are rapidly depleted as a result of cultivation or soil movement. The intermediate labile compartments, which have a turnover rate of 10 to 20 years, must be periodically replenished to maintain an adequate level of SOM, fertility and sustainability.

Particulate SOM

Actual research indicates that the POM properties are consistent with the characteristics of the intermediate labile SOM compartments. In native grassland soils, POM can account for up to 48 % of the total SOC and 32 % of the total soil N (Greenland and Ford, 1964).
Numerous authors have found that POM (or POC) shows promise as a short-term or early warning indicator of long-term changes in soil quality. The POC has been shown to be much more responsive than SOC to changes in agricultural management and, as such, has been proposed as an indicator of soil quality (Gregorich and Ellert, 1993).
Others authors (Sikora et al., 1996) have indicated that the POC pool consist of partially decomposed fractions of plant residue and has a density of < 1.85 g cm-3 and a C/N ratio larger than 20. Light fraction (LF) SOM, as defined originally by Greenland and Ford (1964), has a similar density (< 2.0 g cm-3) and an approximate C/N ratio of 25.
For the propose of this paper, LF SOM is considered analogous to POM.

Soil Microbial Biomass

Live soil biomass is responsive for the mineralization of plant residue or POM and the genesis of more stable humic substances and the production of plant nutrients. For that reason soil biomass is essential to the dynamics of SOM quality. Besides microbial biomass, respiration, enzymes and active OC (carbohydrates, peptides, etc.) are factors associated with total OC and may function as indicators of soil quality. However, Campbell et al. (1991) suggested that an increase in biomass N does not necessarily imply an improvement of soil quality. On the other hand, Chien et al. (1964) reported that respiratory capacity increased with improvement of soil fertility. Sikora et al. (1996) consider that “data suggest that microbial biomass, CO2 evolution, and specific respiration changes are much larger than SOM changes and that these compartments may allow more rapid evaluation of treatment effects” on soil fertility and plant production.
Presently, the more used biomass indicators are : Biomass C (kg ha-1) ; Biomass C / total OC ratio (%) ; CO2-C evolution (kg ha-1 d-1) ; potentially mineralizable N (kg ha-1 y-1) ; and CO2-C / Biomass C (d-1). Enzyme production and evolution have been studied, but results are not yet conclusive.
The biomass indicators must have a common feature : they require a good, fresh sample, with no excess of humidity and temperature, and incubation processes under controlled conditions when required.

Physical separation and characterization of SOM

Many authors (Turchenek and Oades, 1979 ; Anderson et al., 1981 ; Tiessen and Stewart, 1983 ; Christensen, 1985 ; Balesdent et al., 1988 ; Christensen, 1992 ; Cambardella and Elliot, 1993) have employed soil physical fractionation according to particle-size or density to study the behavior of SOM in these fractions.
The procedures have been useful to determine differences in the structural and dynamics properties of SOM. Ultrasonic disruption is commonly used to obtain a good level of dispersion of soil structure. Different ultrasonic levels may generate changes in the distribution of the SOM in particle size fractions. “For example, increasing the intensity of sonication can result in the distribution of more organic material into finer soil fractions compared with a lower intensity of sonication energy” (Elliot and Cambardella, 1991).
As indicated previously, Andriulo et al. (1990) have separed two main particle-size fractions : the fine (FF) and the coarse (CF) fractions, wich have important agronomic significance.
Mineral particles in size fractions alterate estimations of OC turnover time within the fractions. The more complex densitometric fractionation allows one to physically separate OC found within an specific size class from the more dense mineral components.
Cambardella and Elliot (1993) developed a dry and wet sieving procedure to isolate and characterize SOM fractions that were originally occluded in the soil aggregates and evaluated their effect on the turnover of the OM in several cultivation systems.
Nutrient contents were determined for discrete size/density OM fractions isolated from within the macroaggregate structure. Eighteen percent of the total SOC and 25 % of the total N in a no-till soil was associated with fine-silt size isolated from inside macroaggregates (enriched labile fraction, ELF) and particles having a density of 2.07 - 2.21 g cm-3. Sodium matatungstate was used to obtain different solution/suspension densities (1.85, 2.07 and 2.22 g cm-3). The amount of OC and N sequestered in the ELF decreased as the intensity of tillage increased, when comparing no-till and conventional tillage.
The procedure proposed by Cambardella and Elliot (1993) has improved the quantitative determination of SOM fractions, which in turn increases understanding of SOM dynamics in cultivated grassland systems.
Aggregate disruption was done by suspending 10 g dried aggregate soil in 60 mL of water at room temperature and immersing an ultrasonic probe to a depth of 3 - 5 mm into the soil suspension. The disrupted soil suspension was passed through a series of sieves in order to obtain four size fractions : 1) > 250 µm ; 2) 53 - 250 µm ; 3) 20 - 53 µm ; and 4) < 20 µm, which were dried over-night at 50º C.
The specific rate of mineralization (µg net mineral N / µg total N in the fraction) from microaggregate-derived ELF was not different for the various tillage treatments (bare fallow, stubble mulch, and no-till) but was greater than for intact macroaggregates.

Sampling

The objective of soil sampling is to obtain reliable information through samples collected from a large soil body called, for statistical purposes, a “population”. The sample may or may not be representative of the population, depending on how the sample is selected and collected.
Sampling is aimed at obtaining the true value of certain properties such as organic matter, nitrogen (N), phosphorous (P) and other elements or property values and/or contents. Also the purpose of sampling is to estimate these parameters with an accuracy that will meet our needs at the lowest possible cost.
Sampling errors may be larger than those obtained from different laboratories for the same determination. Therefore, the precision of field sampling is a major limiting factor in a soil testing program.
Information from the sample is of interest only insofar as it yields information about the population, but the information may or may not be representative, depending on how the sample was selected.
Unless the soil sample can be assured of a true field representation, then differences in laboratory methods, fertilizer strategy and computer adaptations can only be of academic interest.
Important aspects to be taken into account in sampling are:
1.    accuracy and precision;
2.    sample areas that are representative of the field;
3.    effect of field size on accuracy;
4.    when, how deep, and how often to sample ;
5.    in-site variability of soil properties ;
6.    soil bulk density at the time the sample is taken ;

The failure to adjust for bulk density when a sample is taken is perhaps the greatest errors in determining soil C content in the field.
In order to obtain a truly representative sample it should be considered the  uniformity of the area, topography, texture, cropping patterns, etc., taking into account horizontal, vertical, and temporal variations. Soil map units derived from changes in topography, underlying geology, and dominant vegetation type can be used for horizontal subdivisions. Soil horizons are excellent subdivisions of vertical change.
The deposition of dung and urine on grazed fields increase the variation associated with soil sampling (Friesen y Blair, 1984). It also can exist a biological activity (large animals and/or earthworms) which will disrupt the soil properties (Subler and Kirsch, 1998).
In general, the best sampling design is one that provides the maximum precision (smaller error) at a given cost or that provides a specified precision (acceptable error) at lowest cost.
Effort has sometime been directed toward the use of worker judgment in searching the most “typical” site or a representative view of the population. However, this results is samples with biased errors. If a small sample is to be taken, a “judgment” sample may have a lower error than a random one. As sample size increases, however, the error with a random sampling become smaller whereas that with a “judgment” sampling becomes larger.
Stratified composite random sampling combines composite sampling with sampling by sub-areas having each of them similar topographic characteristics.
Rigney (1956) concluded that errors in the zigzag and stratified random sampling methods were consistently greater than those of the random distribution. Because sampling by the zigzag method is relatively easier, it has been used by many soil testing programs.

Simple random sample

Simple random sampling is defined as a sample obtained in such a way that each possible combination of n units to be taken has an equal chance of being selected.
When more than one sampling units are included, estimation of mean (y) and variance (s)  are given by:
Sampling and analytical work are expensive. But cost can be reduced by taking only as many samples as are needed for a given level of precision. We can calculated the size of the sample needed (n) by (Crépin and Johnson, 1993):
where:
t, is a number taken from a “t” table for a chosen level of probability and the degree of freedom (n-1, arbitrarily fixed the first time and modified by reiteration);
s2, is the variance of SOM levels known from other studies or estimated by s2=(R/4)2, where R is the estimated range likely to be found in sampling;
D, is the variability in mean estimation of SOM we are willing to accept.

Sample management

Prevention of contamination

Great care must be taken to prevent sample contamination during sampling collection (dirty containers, sampling tools contaminated with previous samples, etc.) and drying processes (dusty places).

Mixing

A three - four L sample should be thoroughly mixed for subsample homogenization and reduce the volume down to 0.5 - 1 L of representative sample.

Drying and sieving

Wet samples should be dried at temperatures not greater than 35-50°C. Higher temperatures can alter nutrient solubility from organic and mineral soil fractions.
Soil aggregates are crumbled in a porcelain or agate mortar and passed through a 2 mm sieve. Homogenization of fine soil for further analysis is made with a mechanical grinding mill.
Dust from sample processing is a nuisance and a important contaminant. A mask should be worn to avoid breathing the dust.

Splitting and storing

The field-collected material must be thoroughly mixed and a sub-sample for analysis and storage should be taken. Clearly identified samples should be kept in a cool, ventilated, and dry storeroom.

ESTIMATION OF SOM

Soil OM has been defined as “the organic fraction of the soil exclusive of undecayed plant  and animal residues” (SSSA, 1997) and has been used synonymously with “humus”. However, for laboratory analysis, the SOM generally includes only those organic materials that accompany soil particles through a 2-mm sieve (Nelson and Sommers, 1982).
It is difficult to quantitatively estimate the amount of SOM. Procedures used in the past involve determination of change of weight of the soil sample resulting from destruction or descomposition of organic compounds by hydrogen peroxide (H2O2) treatment or by ignition at high temperature. Both techniques are subject to error.
Numerous methods are available for the determination of OC and the estimation of OM, all of then having inherent associated problems. For that reason, the method used should be most applicable to the soils being analyzed and the required accuracy of the results. All current methods are included in one of the following principles (See Table 1 for a classification and comments on the different methodologies) :
I.                   Wet oxidation of OC
a.                  using acid dichromate solution
1.                  followed by back trititation of the remaining dichromate or photometric determination of Cr3+;
2.                  followed by collection and determination of evolved CO2;
b.                  by H2O2 followed by measurement of weight change;
II.                Dry oxidation of OC in a furnace
a.                  followed by measurement of weight change (weight-loss-on-ignition, WLOI)
b.                  followed by collection and determination of evolved CO2;
Note : for high temperature combustion techniques carbonate-C will also be included in the total C recovered and, therefore, must be subtracted to obtain organic C.

I.a.1. Titrimetric dichromate redox methods

The potassium dichromate procedures are widely used in soil OC determination because of their rapidity and simplicity compared with others wet or dry combustion methods. Advantage of acid oxidation is the removal of carbonate C.
Dichromate redox methods potentially suffer from a number of interferences and low recoveries. In the Walkley and Black (1934) method, OC is oxidized solely by the action of dichromate, heated by the exothermic mixing of aqueous dichromate and concentrate sulfuric acid (»120°C). As indicated in the Introduction, a factor must be used due to the incomplete oxidation reached with this temperature (Walkley & Black, 1934). This factor is not valid for different soils (Gillman et al., 1986), different horizons of a soil profile or organic fractions (Galantini et al., 1994).
Some authors proposed supplementary heating on a hot plate or block digestor (Mebuis procedure) to improve OC oxidation (Mebius, 1960; Schollemberger, 1945; Heans, 1984). In some cases two blanks are recommended due to the thermal instability of dichromate acid.
Several errors can affect the quantification of OC by this procedure:
·                 a source of positive errors is the presence of Cl- (producing CrO2Cl2) and Fe2+ or Mn2+ (producing Cr3+ and Fe3+ or Mn3+)
·                 the higher oxides of Mn (largely MnO2) compete with dichromate for oxidizable substances when heated in an acid medium resulting in a negative error (Tabatabai, 1996).
·                 the acid dichromate digestion solution decompose at temperature above 150°C, limiting the heating procedure (Charles and Simmons, 1986).
Despite the difficulties and inaccuracies, the dichromate redox method is widely used because it requires a minimum of equipment, can be adaptable to handle large number of samples, is suitable for comparative work on similar soils and it is not costly.

I.a.2. Wet oxidation of OC with collection and determination of evolved CO2

The collection and determination of evolved CO2 can eliminate many of the interferences of the titrimetric dichromate redox procedure and can be employed at a higher digestion temperature (Allison, 1960). The evolved CO2 may be trapped in solid absorbents (followed by measurement of weight changes) or in a alkali solution (followed by back titration). An adaptation described by Snyder and Trafymov (1984) allows the processing of large number of samples at a time.

I.b. Wet oxidation with hydrogen peroxide

The H2O2 method proposed by Robinson (1927) produces the incomplete oxidation of OM varying the extent of oxidation among soils (Gallardo et al., 1987). Additional errors are introduced both in filtering and drying the oxidized residue and the filtrate at 110 °C. The method is, therefore, unsatisfactory as a means of determining total SOM.

II.a. Dry combustion followed by measurement of weight changes

The dry combustion is followed by the measurement of changes in weight-loss-on-ignition (WLOI). This is a  alternative method (non Cr contaminant) for the dichromic acid digestion, which is widely used in soil testing laboratories. The method assume that SOC is oxidized completely within a narrow temperature range at which losses (water of hydration for example) from minerals is negligible. Unfortunately, this is not the case, so temperature selection is somewhat arbitrary but, at the same time, critical to minimize errors (Schulte and Hopkins, 1996). Temperature higher than 500°C can result in errors from loss of CO2 from carbonates, structural water from clay minerals, oxidation of Fe2+, and decomposition of hydrated salts (Schulte and Hopkins, 1996). Temperatures below 500°C should eliminate many of these errors but may result in incomplete SOM oxidation (Gallardo et al., 1987).
Hygroscopic water can be removed by heating 24 h at 105°C, but more temperature is required to remove water from gypsum or MgSO4.7H2O.
Some authors have reported a complete SOM oxidation at temperatures between 430º C and 500º C (Davis, 1974 ; Giovanni et al., 1975 ), with no loss of carbonates. Others studies showed that part of the soil humic compounds resist 600°C (Gallardo et al., 1987).
Differential thermal analysis showed an endothermic peak around 250°C (due to gibbsite interference), a first exothermic peak around 350°C (oxidation of humic acid lateral chains) and a second peak around 450°C  (oxidation of humic acid central nucleous).
Dupuis (1971) found that the presence of sodium salts thermically stabilized the SOM, while aluminum slows down (humic-aluminum complexes) or accelerates (excess of aluminum) combustion.
Gallardo et al. (1987) concluded that part of the SOM may be thermostable in presence of certain inorganic compounds such as the transitional elements.
A review of papers comparing WLOI method is presented by Schulte and Hopkins (1996). Significant regression equations were found in all cases and r2 higher than 0.90 in most of them. Differences in slopes for regressions may be attributed to differences in heating time and temperature, and/or differences in the nature of the clay and SOM fractions.
WLOI may be reasonably accurate and economical for estimating SOM if precautions such as hygroscopicity and hydrated salt content are taken into account.

II.b. Dry oxidation of OC with collection and determination of evolved CO2

The use of a stream of O2 in the tube furnace at temperature over 900°C provides for near complete combustion of OC. The addition of MnO2 and CuO as catalysts convert any evolved CO to CO2 (Tiessen et al., 1981).
The most convenient combustion methods use automated instruments, which heat the sample by induction and determine the evolved CO2 by infrared absorption. Many of the commercially available units are equipped for simultaneous C, H, N, and/or S analysis.
In soils with significant levels of carbonates, however, organic carbon can be overestimated if carbonates are not removed before analysis or account for by analysis.

Soil thickness and bulk density as errors in an analytical result

Chemical analysis indicate the concentration of elements in soils, but thickness and bulk density of the soil layer in the field must be considered to estimate quantities of element per unit of area. For management-induced changes in SOM sequestration to be well assessed, the masses of soil being compared must be equivalent (Ellert and Bettany, 1995).
Sampling activities require quality assurance principles such as the development and use of standardized sampling procedures, training and documentation of sampling personal, and creation of traceable and defensible information through the use of labels and chains of custody (Kesta and Bartz, 1996)

Comments

SOC determination and SOM estimation can be obtained by using wet and dry combustion procedures. The wet combustion titrimetric method is rapid, simply and rather accurate. The WLOI oxidation determination is also rapid and accurate provided the procedure errors are evaluated and applied .
The main sources of errors in the SOM estimation are :
· erratic field sampling
·  the use of correction factors : oxidation of OC (1/0.76 = 1.33) and conversion of OC to OM (1.724) coefficients
· inherent titration and volumetric analytical procedure
· presence of inorganic C (carbonates)
·  the need to express results on a volumetric basis by using the soil bulk density taken at sampling


No comments:

Post a Comment