Anda di halaman 1dari 31

A.I.J.M. van Dijk (2002) Water and Sediment Dynamics in Bench-terraced Agricultural Steeplands in West Java, Indonesia.

PhD Thesis, Vrije Universiteit Amsterdam

Chapter 6 Water budget of inter-cropped maize and cassava on bench terraces

Abstract: A combined measuring and modelling approach was employed to establish the annual water budget of a rain-fed cropping system with maize (Zea mays L.) and cassava (Manihot esculenta Crantz) on a bench-terraced hillside in upland West Java. Latent heat fluxes were calculated from micro-meteorological measurements made during five wet season months by the temperature-variance Bowen-ratio energy balance (TVAR-BREB) method. A simple radiation energy partitioning model was used to estimate the contribution of soil evaporation via the Priestley-Taylor model. The Penman-Monteith model with calibrated values for surface conductivity (12.7 mm s-1), roughness length (0.03 times crop height) and displacement length (0.64 times crop height) was used to estimate annual crop transpiration and soil evaporation, allowing for the effect of reduced soil water availability during the dry season. An adapted version of the analytical model of rainfall interception was used to estimate interception losses as a function of changing vegetation density. The consistency of modelled ET rates was investigated with the aid of a soil-vegetation-atmosphere (SVAT) model which satisfactorily simulated observed soil water contents. The results suggest an annual water use of 1228 mm, divided between 590-662 mm (48-54%) crop transpiration, 368-440 mm (30-36%) soil evaporation and 198 mm (16%) rainfall interception losses.

Parts of this chapter are submitted to Agricultural and Forest Meteorology with L.A. Bruijnzeel and J. Schellekens as co-authors.

6.1. Introduction The rapid conversion of forest to other land uses (notably annual cropping and pasture) in the humid tropics has raised concerns about the potentially negative impacts on water resources (Hamilton and King, 1983; Pereira, 1989; Jepma, 1995). Conversion of forest to shorter vegetation types has been observed to affect both amounts and timing of streamflow, reflecting both the decrease in infiltration opportunities that usually accompanies deforestation (Lal, 1987) and differences in vegetation water use (Bruijnzeel, 1989, 2002). There is a considerable body of information on evapotranspiration (ET) for tropical forest (Bruijnzeel, 1990; Gash et al., 1996;

73

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA Schellekens et al., 2000; Roberts et al., 2002) but a distinct lack with regard to ET from rain-fed upland crops (Klinge et al., 2001). In this chapter we report measured and modelled water fluxes in a rain-fed agricultural cropping system with inter-cropped maize (Zea mays L.) and cassava (Manihot esculenta Crantz) in upland West Java, Indonesia. Originating from tropical South America, this cropping system is also widespread in Southeast Asia and Central Africa (e.g. Moreno and Hart, 1979), in particular on less fertile soils. The micrometeorology of mono-cropped maize has been studied rather extensively in temperate and, to a lesser extent, (seasonally) dry tropical climates (e.g. Uchijima, 1976; Steduto and Hsiao, 1998a, 1998b; Rochette et al., 1991), but studies in the humid tropics are scarce if not entirely absent. No studies of total water use by cassava, whether mono- or inter-cropped, are known to the authors although rainfall interception by tree cassava has been investigated in Thailand (Witthawatchutikul and Tangtham, 1986), Cameroon (Waterloo et al., 1997) and Amazonia (Sommer, 2000). Past work on rain-fed crop water use in the tropics has mainly used the soil water depletion technique (e.g. Ringoet, 1952; Gunston and Batchelor, 1983) but inferences are limited to drier periods because of the difficulty of separating drainage and soil water uptake (Ward and Robinson, 1990). Possibilities for adequate evaluation of the drainage component have improved considerably with the advent of soil hydrological recording equipment, however (cf. Klinge et al., 2001). In the current study a combination of micro-meteorological and soil hydrological measurements and soil-water-atmosphere (SVAT) modelling is employed to establish the water budget for an entire year of cropping of the maize-cassava crop mixture. If lateral transfer of water through the soil is neglected, the water budget of a column of soil can be written as: P = D + Q + ET + S = D + Q + Ei + Ev + Es + S [6.1]

where P is rainfall, D drainage, Q surface runoff and S the change in water contents of the block of soil, while total ET can be divided further into interception loss (Ei), crop transpiration (Ev) and soil evaporation (Es) (with Et=Ev+Es). In the current study rainfall and surface runoff were measured whereas S was estimated from measurements of soil water contents and suction. Furthermore, micro-meteorological measurements were made during the first five (wet-season) months of the cropping cycle (November 1998 to April 1999) to determine ET. Latent heat fluxes were determined by the temperature variance Bowen ratio energy balance (TVAR-BREB) method and these were in turn used to parameterise the Penman-Monteith model for the rest of the year, accounting for changes in vegetation density. Although this approach in principle should yield the sum of the three ET components, there is considerable evidence to suggest that micro-meteorological methods may lead to an under-estimation of wet canopy evaporation (Ei), particularly under the maritime tropical conditions prevailing in the study area (Chapter 5; Schellekens et al., 2000). For this reason, only Et during intervals without rainfall was determined via micro-meteorological methods, whereas evaporation during rainfall was equated to interception losses (Ei). This approach will introduce some errors of its own, as any crop transpiration and evaporation from the wet soil (or its mulch cover) during storms is not accounted for. This component will initially be assumed to be insignificant, however. Interception losses were estimated using the analytical rainfall interception

74

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

600 500 P, E 0 (mm mo )


-1

400 300 200 100 0 N D J F M A M J J A S O N

Fig. 6.1. Monthly rainfall (P, closed circles) and Penman open water evaporation (E0, open circles) during the measuring period (17 November 1998-16 November 1999), along with average monthly rainfall for 1994-2001 (dashed line).

model adapted for use in a vegetation with variable leaf area index (Chapter 4) and employed successfully to simulate locally observed throughfall and stemflow amounts (Chapter 5). Furthermore, simple models were used to simulate the effect of reduced soil water availability on water use during the dry season and to obtain an estimate of the relative contributions of soil evaporation and crop transpiration to Et. Finally, the consistency between simulated ET amounts and observed soil moisture contents was investigated with a one-dimensional soil-vegetation-atmosphere model (VAMPS; Schellekens, 2000).

6.2. Materials and methods 6.2.1. Site characteristics, cultural practices and vegetation characteristics All work was carried out in the 125 ha Cikumutuk catchment, of which ca. 62% was covered with rain-fed mixed cropping system at the time of research. The area is situated 40 km East of Bandung, near the town of Malangbong (703S, 10804W). Oxisols have developed in the Quarternary volcanic tuffs and underlying agglomerates. Slopes are up to 15 and mostly bench-terraced. Further details on geology, soils and land use may be found in Chapter 3. The area experiences a humid tropical climate with a drier season (average monthly rainfall less than 60 mm) extending from July until September. Mean annual rainfall is about 2650 mm and Penman open water evaporation (E0) was estimated to be ca. 1400 mm for the modelled hydrological year (17 November 1998-16 November 1999). Average 24-hours minimum and maximum temperatures during the wet season were 20.1EC and 27.8EC, respectively and minimum and maximum humidity values 56% and 96%. Monthly rainfall and E0 during the 1998/99 measuring period are shown in Fig.
75

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA 6.1. Monthly rainfall totals were generally similar to longer-term average values, although February was notably drier whereas December, January and March were slightly wetter. Micro-meteorological and soil hydrological measurements were made on a 0.4 ha terraced field situated on the lower part of the south-west facing slope of the catchment, at an altitude of 590 m a.s.l., close to the site where simultaneous measurements of throughfall and stemflow were made (indicated by code EF in Fig. 3.1). The original, now terraced slope had a gradient of 19% at this point. The crop stand was uniform within the field; surrounding fields were also terraced and cropped with cassava or cassava-maize mixtures. A 6 m mast was erected in the middle of the field, ca. 40 m away from the south-western (downslope) end of the field, this being the prevailing direction of wind during daytime. Paths and gullies aligned with scattered low shrubs and a few severely lopped trees constituted the borders of the fields. Vertical steps between adjacent bench terraces within the field typically measured 0.5-1.0 m. The cultural practice during the measuring season can be considered typical for rainfed agriculture in the study area and has been described in some detail in Chapter 5. Maize (cultivar Pioneer P5) was sown on 17 November 1998, some weeks after the rainy season had started. Sowing rows spaced 60-70 cm apart on the terrace bed were orientated perpendicular to the terrace toe drain and resulted in a planting density of 4.3 plants m-2. Upland rice was also sown but failed to germinate. Cassava stem cuttings were planted two weeks later in rows alternating with the maize at a density of 1.8 plants m-2. The maize flowered after about 55 days after planting (DAP) and the cobs started to develop a week later. Maize senescence occurred after 82 DAP and the cobs were harvested on 126 DAP, at the end of March 1999. Cassava started to form tubers after about three months but these were not harvested before the end of the simulation period in November 1999. The manner in which leaf area index (LAI) and fractional canopy cover were determined has been discussed in Chapter 5. In short, the development of LAI was estimated indirectly from observations of plant height, leaf biomass and specific leaf area as collected at different stages during the cropping cycle. Canopy cover (c) was related to LAI via the Beer-Lambert equation for light interception by a vegetation canopy:

c = 1 e LAI

[6.2]

where is a dimensionless extinction coefficient determined from vertical photographs of the canopy taken from the top of the micro-meteorological mast on a regular basis between November 1998 and May 1999 (N=10). The photographs were scanned and analysed using image processing software to derive canopy cover fractions, which were then related to estimated LAI values using Eq. [6.2] to give a value of =0.75 for nonpenetrating types of radiation (e.g. PAR, by approximation; Chapter 5). Seasonal changes in crop height, LAI and canopy cover are illustrated in Fig. 6.2. Maximum crop height, total LAI and canopy cover were reached just before maize senescence (82 DAP) and determined at 2.20 m, 3.7 (of which 70% maize leaves) and 0.94, respectively. Although the cassava gradually grew taller after the maize harvest, the LAI remained more or less constant at ca. 1.1 because the lower leaves were shed as new ones emerged at the top of the plants.

76

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

Crop height, h v (m)

3
(a)

2 1 0
0 20 40 60 80 100 120 140 160

DAP

Leaf Area Index, LAI

4
(b)

total LAI
maize

3 2 1 0
0 20 40 60 DAP
1.0
(c)

cassava

80

100

120

140

160

Canopy cover, c

0.8 0.6 0.4 0.2 0.0


0 20 40 60 80 100 120 140 160

DAP

Fig. 6.2a-c. Development of (a) crop height (hv), (b) leaf area index (L; circles indicate estimates based on plant height measurements, lines are fitted functions) and (c) canopy cover (c; open circles indicate canopy cover as derived from overhead photographs, whereas the drawn line indicates values calculated from LAI using Eq. [6.2]) during the period of micro-meteorological measurements. DAP = days after planting on 17 November 1998.

6.2.2. Micro-meteorological methods The Bowen Ratio Energy Balance method (BREB; Angus and Watts, 1984) can be used to determine the latent heat flux (8E, W m-2), which is linked to ET (=E, mm h-1) via the latent heat of vaporisation 8 (2.44A106 J kg-1). The method solves the surface energy balance, written as:

77

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

E = RN H G S

[6.3]

where RN is net radiation, H sensible heat flux, G soil heat flux and S heat stored in the vegetation. Net radiation was not measured during the full observation period, but estimated where necessary from net incoming short-wave radiation using a regression equation. Preliminary calculations (cf. Brutsaert, 1982) suggested that changes in heat storage in the crop biomass and in the air within the vegetation (S) were both less than 2 W m-2 K-1 because of the limited vegetation density and height. Therefore, these fluxes were assumed negligible. The Temperature Variance Bowen Ratio method (TVBR; Vugts et al., 1993) may be used to evaluate the magnitude of 8E and H in a horizontal homogeneous atmospheric surface layer based on Monin-Obukhov similarity theory. When rapid fluctuations in air temperature ( T) and humidity ( q) are measured simultaneously, the Bowen ratio ($=H/8E) can be obtained with the TVBR method. Fluctuations in wet-bulb temperatures ( Tw) measured in the present study can be used instead of air humidity, as long as there is near-perfect linear correlation between T and Tw (Vugts et al., 1993). The corresponding equation reads:

c p T H = = T E q

{ [(s + )

T w T ] 2 ( s + ) T T w TTw 1 2

)}

1 2

[6.4]

where ( is the psychrometric constant (66.8 Pa K-1), s the slope of the temperaturevapour pressure curve at wet-bulb temperature Tw (Pa K-1) and TTw the coefficient of correlation between dry- and wet-bulb temperature fluctuations. The conditional sign in Eq. [6.4] is opposite to that of TTw. Compared with the conventional temperature variance method (Tillman, 1972) the TVBR approach suffers less from errors associated with frequency losses as both fluctuation terms are affected (Van der Molen, 2002). Having calculated the Bowen ratio, the latent heat flux can be found from:

E =

1 A 1+

[6.5]

where A (W m-2) is the available energy (equal to RN-G, cf. Eq. [6.3]). Half-hourly evaporation rates determined with the TVBR method were used to evaluate the magnitude of surface conductance (gc, m s-1) by inverse application of the PenmanMonteith big leaf model (Monteith, 1965). The original equations reads:

E =

sA + c p (es ea )g a s + ( ga gc ) 1

[6.6]

where (kg m-3) is the specific density of air, cp (J kg-1 K-1) the specific heat of air at constant temperature, (es-ea) the deficit (Pa) between saturated and actual vapour pressures, respectively, and ga (m s-1) the aerodynamic conductance. The latter can be estimated from wind speed measured under neutral atmospheric conditions (Brutsaert and Stricker, 1979; Monteith and Unsworth, 1990):

78

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

Reference Lemon (1960) Cowan (1968), Tanner and Pelton (1960) * This study
a

d 1.04hv0.88

z0 0.062hv1.08

ga (3.4) ga (6.0) mm s-1 mm s-1 28 36 15 20 27 14

Avg. Diff. 22% 20% 9%

r2 0.53 0.63 0.87

0.64 to 0.67 hv 0.10 to 0.13 hv 0.64 hv 0.03 hv

Lower values were used in calculating ga.

Table 6.1. Agreement between values of aerodynamic conductance (ga) of a mixed maizecassava crop as calculated from wind speed measurements at 3.4 and at 6.0 m, respectively, using Eq. [6.7] and expressions for d and z0 proposed in the literature (for maize only) as well as optimised d/hv and z0/hv ratios.

zd zd z ln +ln 0 g a = k u z ln z z0 z 0 0 h
2

[6.7]

where k is the Von Krmn constant (0.41), z (m) the height above the surface at which wind speed uz (m s-1) is measured, d (m) the zero-plane displacement length and z0 and z0h (m) the roughness lengths for momentum and heat transport, respectively. The parameters d, z0 and z0h vary as a function of vegetation height (hv, m) and structure (Thom, 1975) and a variety of equations has been proposed for maize (Table 6.1). Measured crop heights and wind speed at heights of 3.4 and 6.0 m were used to test the cited expressions of z0/hv and d/hv. To limit the number of possible combinations, the ratio z0h/z0 was set equal to 0.25. No stability corrections were applied. Only wind speeds in excess of 1 m s-1 measured under near-neutral conditions were used, as evidenced by small absolute Richardson numbers (-0.01<Ri<0.01). The degree of agreement between ga values calculated with Eq. [6.7] from wind speed measurements at the two heights was used to optimise z0/hv and d/hv. This approach will yield results that are similar to those obtained by the graphical procedures of Thom (1975). For daytime half-hour intervals, surface conductance (gc) was calculated using the inverse Penman-Monteith equation (Eq. [6.6]) with measured values for the other variables. The possibility to use the Penman-Monteith equation for the computation of daily evaporation using crop characteristics and average daytime (instead of half-hourly) values of climatological parameters was also explored. Daytime Et was estimated using daytime average values of the climatological variables (A, T, RH and uz) measured in the field, together with crop height (hv) and appropriate functions describing ga and gc as a function of the day after planting (DAP) and any other relevant variables. Penman open water evaporation (E0, mm d-1) was calculated for comparison following Penman (1956) and using measurements made at a climatological station near the experimental site (see Section 6.2.3). Strictly speaking, the Penman-Monteith equation was developed for use in a vegetation with full canopy cover and its use in sparse vegetation leads to some uncertainty in the interpretation of derived values of ga and gc (cf. Kim and Verma, 1991). A number of Penman-Monteith type models are available to separate the

79

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA contribution of soil and vegetation to total Et (e.g. Shuttleworth and Wallace, 1985; Dolman, 1993). However, neither micro-meteorological measurements within the crop canopy nor independent estimates of one or both Et components, which would have helped parameterise and validate such a model, were made in the present study. A first estimate of soil evaporation (Es) was obtained, however, by assuming that soil evaporation always proceeded at a potential rate and as such can be estimated with the Priestley-Taylor equation (Priestley and Taylor, 1972). This assumption seems reasonable for wet topsoil; it is shown further on that this was the case for most of the measurement period. Soil evaporation can therefore be estimated as (cf. Kabat et al., 1997):

Es =

s (RN ,s G ) s +

[6.8]

where RN,s (W m-2) is the net radiation energy for the soil. To estimate RN,s two further assumptions had to be made, viz.: (i) net radiation was partitioned among the vegetation and soil in the same manner as incoming short-wave radiation, and (ii) the latter could be approximated by considering the fractional coverage and albedo of the soil and the vegetation. A simple model was fitted to the available time series of overall albedo ( ) and ground cover (c):

= c v + (1 c ) s

[6.9]

where c was calculated from the LAI time series using Eq. [6.2] and the albedo values for vegetation ( v) and soil ( s) were optimised by least squares methods. Different values of s were used to represent the bare soil before the maize harvest, and for the soil with partial mulch cover thereafter. Subsequently, RN,s was estimated as:
R N ,s = (1 c )(1 s )R N

[6.10]

It is acknowledged that this approach simplifies the more complex transfer of radiation through a canopy as well as contrasts in the behaviour of short- and long-wave radiation (cf. Ross, 1981; Sinoquet and Bonhomme, 1992), but it arguably yields a fair approximation nonetheless. Values of Es computed by Eq. [6.8] were subtracted from total Et to yield the remaining transpiration component (Ev). 6.2.3. Micro-meteorological instrumentation Micro-meteorological measurements were made from 27 November 1998 (10 DAP) until 24 April 1999 (158 DAP). This period covers most of the cropping cycle including the emergence, growth, senescence (setting in around 82 DAP) and final harvesting of the maize (on 23 March 1999, 126 DAP) plus a subsequent phase with only cassava and 53% soil cover provided by maize harvest residues. Due to instrumental problems, records were not available for the periods 13-25 DAP and 106-116 DAP. Incoming short-wave radiation (RS9, W m-2) was measured with a pyranometer (Skye Instruments SP-1110) and in combination with outgoing short-wave radiation (RS8, W m-2) with an albedometer

80

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA (Kipp & Zonen CM-7). Net radiation (RN, W m-2) was measured with a net radiometer (Radiation and Energy Balance Systems Q7) from 83 DAP onwards. All instruments were mounted at a height of 2.9 m. Soil heat flux (G, W m-2) was determined with a flux plate (Middleton & Co) installed 2 cm below the surface. Air temperature (T, EC) and humidity (RH, %) at 2.9 m and 5.5 m heights were measured by probes (Campbell Scientific CS500) placed inside a non-aspirated gill-type radiation shield. Wind direction was recorded with a potentiometer windvane (Vector Instruments W200P) and wind speed (u, m s-1) at heights z=3.4 m and 6.0 m with three-cup anemometers (Vector Instruments A101M/L). The standard deviations of wet- and dry bulb temperatures ( T and Tw, respectively, EC) were determined with fast-response thermocouples (Tillman, 1972) made of chromel-constantane wire with a diameter of 0.2 mm at the welding point. Three sets of dry and wet thermocouples were used: the first two at 2.6 m and 5.4 m in the mast and the third for replication and back-up at 2.4 m above the ground at a distance of 10 m from the mast, respectively. All sensors were connected to a data logger (Campbell 21X). Thermocouple temperatures were recorded at 0.5 Hz and pre-processed at 5-minute intervals to minimise the effect of trends on half-hourly standard deviations. Half-hourly average temperatures were also stored for quality control. The other instruments were sampled every 30 seconds and the data stored as half-hourly averages and standard deviations. Rainfall was measured using a tipping bucket-logger system (Vrije Universiteit Amsterdam) positioned above the crops, which registered rainfall in increments of 0.07 mm. The data were calibrated and resampled into 30-minute intervals using methods described in Chapter 11. During the entire period daily measurements of standard meteorological variables were also made routinely at a climatological station located a few hundred metres away from the site, on the southern ridge of the experimental catchment (code CS in Fig. 3.1).

6.2.4. Soil physical and hydrological measurements Surface runoff was measured between 15 December 1998 and 11 May 1999 from sections of two bench terrace beds close to the experimental site (Chapter 12). Total runoff amounts were 75 mm (4.2% of 1775 mm rainfall) and 1 mm (less than 0.1%) distributed over 27 and 4 runoff events, respectively. Measurements from two entire bench terrace units nearby resulted in 115-138 mm of surface runoff (6.8-8.1% of 1697 mm rainfall) for the period 2 December 1998-11 April 1999. The different results obtained for terrace beds and terrace units is caused by the lower infiltration capacity of the terrace riser and toe drain sections (cf. Chapter 12). Field surface saturated conductivity (Ksat in m d-1) was measured using a custombuilt double-ring infiltrometer while topsoil unsaturated conductivities (Kunsat in m d-1) at soil water suction values of 3, 6 and 12 hPa were measured with a tension infiltrometer (Watson and Luxmoore, 1986; Jarvis and Messing, 1995). Furthermore, 52 soil cores (100 cm3) were sampled down to depths of 170 cm to determine dry bulk density (DBD, kg m-3) and saturated (2sat, %) and actual water contents (2, %). Ksat was measured on 29 of these cores in a laboratory permeameter in constant- or falling-head mode (Kessler and Oosterbaan, 1973). The same cores were subsequently used to determine soil moisture retention characteristics on a sand box (0<pF<2.0) whereas a pressure membrane extractor (Soil Moisture Equipment Co.) was used for values above pF=3 (Stakman, 1973).
81

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA (Semi-) continuous measurements of soil water contents (2) were made using two time-domain reflectometry probes (TDR, Campbell Scientific CS615) having 30 cm rods inserted at 0-30 and 10-40 cm depth. Soil moisture suction (F, hPa) was measured with four recording tensiometers (Eijkelkamp Agrisearch Equipment T4) installed at 15, 40, 60 and 80 cm depth. The TDR probes and the tensiometers were connected to a data logger (Campbell Scientific 21X) recording half-hourly average values from 9 December 1998 to 7 March and from 27 March to 18 April 1999. Additional soil water suction data were obtained with thirteen septum-type tensiometers installed at depths between 10 and 80 cm, which were read daily using a custom-built pressure-transducer reading device. Finally, after all micro-meteorological and hydrological instruments had been taken from the field on 24 April 1999, every seven to twelve days the soil was sampled at four locations at and near the experimental site to determine average gravimetric soil moisture content. At each location a five metre long section of terrace bed was selected and twelve equally spaced boreholes were augered to a depth of 120 cm. The collected soil was bulked and thoroughly mixed, after which a sub-sample was taken to the field laboratory and oven-dried to obtain an indication of average gravimetric soil water content. This was done on 25 occasions between 18 April and 31 October 1999, in the early morning to minimise evaporation losses.

6.2.5. Modelling water fluxes The soil water module of the soil-vegetation-atmosphere (SVAT) model VAMPS (Schellekens, 2000) that was used in this study shows distinct similarities to SWATR (Feddes et al., 1978). It has more built-in flexibility, however, and allows user-defined relationships between soil moisture suction, moisture content and hydraulic conductivity as continuous functions or in tabulated form. The model simulates one-dimensional flow in unsaturated soil by solving the Richards equation for a given number of soil layers with a defined depth increment and time-step. To relate volumetric water contents (2) to soil moisture suction (F, absolute value) the approach of Van Genuchten (1980) was used:

= res +

[1 + (F ) ]
n

sat

res 1 1 n

[6.11]

where 2sat (%) is volumetric soil water contents at pF=0 (F=1 hPa), 2res the residual water contents (often equalled to 2 at pF=5.5) and (>0, hPa-1) and n (>1) are curve fitting parameters. Eq. [6.11] was fitted separately to measured 2-F data pairs for three soil layers (0-30, 30-60 cm and 60-170 cm) using the Levenberg-Marquardt method. The 2-F data mainly represented laboratory measurements on core samples, but whenever core samples were taken nearby and at approximately the same depth as a tensiometer the measured suction and field water contents were also included. In addition, simultaneous field measurements of 2 and F were available for the top layer. Unsaturated hydraulic conductivities K(F) (m d-1) were approximated with the Van Genuchten-Mualem model (Mualem, 1976; Van Genuchten, 1980):

K (F ) = K 0 S e 1 1 S e
L

[ (

n (n 1)

)]

[6.12a]

82

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA where

Se =

(F ) res sat res

[6.12b]

and K0 (m d-1) is a fitting parameter that is normally similar, but not necessarily equal, to saturated hydraulic conductivity Ksat while L is a fitting parameter that can attain both negative and positive values (Kosugi, 1999; Schaap et al., 2001). The values of K0 and L for the respective soil horizons were optimised by VAMPS during a calibration period (18 January-3 March 1999) for which half-hourly measurements of P, 8E and topsoil 2 and F were available. The half-hourly measurements of net precipitation (rainfall minus interception loss) required by VAMPS were not available, but were assumed to constitute a constant fraction of gross rainfall for individual storms. The magnitude of this fraction was estimated using total rainfall as input for the event-based interception model presented in Chapter 4. Errors arising from the failure to describe any initial canopy wetting losses are thought to have been very minor because of the small canopy storage capacity of the crops (Chapter 5). Although the interception model was calibrated and tested in a similar vegetation close to the experimental site, slight differences in vegetation density existed. Therefore, LAI values determined on-site were used in the model (Fig. 6.2). During the calibration of VAMPS only soil water contents of the top 30 cm of soil were simulated, using 15 sub-layers of 2 cm each. At the bottom of the column free drainage was assumed to occur, which was supported by the observation that suction values never exceeded pF=1.9 (F=80 hPa) during the calibration period and were usually less than pF=1.5 (F=30 hPa). Roots were assumed to extend over the entire 30 cm, although lesser rooting depths had negligible effects on model results for this period. After calibration VAMPS was run at a daily time step for a whole year, starting 17 November 1998. The soil profile was divided into 60 layers of 2 cm each and free drainage was permitted for the lowermost soil layer. Appropriate Van Genuchten parameters were used for each of the three distinguished soil horizons. No optimised values of K0 and L were available for the subsoil layers. Instead, it was assumed that L had the same value for the entire soil, whereas K0 values were estimated from the K0 of the topsoil, assuming that the relative magnitudes of K0 and Ksat for the three layers were equal. Daily amounts of throughfall were modelled from daily rainfall totals using the interception model referred to earlier. Evapotranspiration (Et) was calculated with the Penman-Monteith model using daytime averages for climatic variables. Data from the standard climate station were used for periods when in situ micro-meteorological measurements were not available due to technical problems and for the period after 24 April 1999. Comparison of data from the site and from the climate station indicated very close agreement. Values of gc for the post-measurement period were assumed equal to those for the preceding period, although it was anticipated that gc would decrease during the dry season. More sophisticated ways of modelling changes in gc were not considered feasible, because (i) the albedo and soil heat flux were likely to have changed as the soil dried out and (ii) only gravimetric soil water content measurements were available for comparing model simulations. Refuge was therefore taken to the simple Et reduction model of SWATR (Feddes et al., 1978), assuming that soil moisture limits Et beyond a

83

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

0.30 0.25 0.20

"

0.15 0.10 0.05 0.00 0 20 40 60 80 100 120 140 160 180

DAP

Fig. 6.3. Change in surface albedo ( ) during the wet season cropping cycle (17 November 1998-24 April 1999) and the fitted radiation partitioning model (Eq. [6.9]).

threshold suction value (Flim) and then decreases linearly to zero at suction value Fend according to:
F F * E t (F ) = max end E t , 0 Fend Flim

for F>Flim

[6.13]

where Et(F) is the actual Et at soil water suction F and Et* the rate when soil moisture is not limiting. Values were Flim=1000 (pF=3) and Fend=16,000 hPa (pF=4.2; Szeicz and Long, 1969; Feddes et al., 1978; Russell, 1980). Because the actual (development of) rooting depth was not known, the model was subsequently run for rooting depths of 30, 60, 90 and 120 cm. Amounts of crop transpiration (Ev) and soil evaporation (Es) after 24 April 1999 were estimated from modelled Et by assuming that their relative magnitude remained the same as estimated for the initial period after maize harvesting (cf. Section 6.2.3). Because this probably over-estimates the importance of soil evaporation for the drier conditions during the dry season, an additional estimate was obtained by assuming zero soil evaporation for the period during which topsoil water suction was simulated to be below field capacity (pF=2.0).

6.3. Results 6.3.1. Energy balance components during the wet season During the course of the cropping season there was a distinct trend in albedo (Fig. 6.3), which increased from 0.07 for bare, moist soil at the start of the season to 0.19 at the height of maize development on 82 DAP (94% cover). As the maize crop senesced,

84

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

0.40 0.35 0.30 0.25


initial growth near canopy closure maize senescence cassava only

"

0.20 0.15 0.10 0.05 0.00 6 9 12


Time (hrs)

15

18

Fig. 6.4. Average diurnal pattern of albedo ( ) for the mixed maize and cassava cropping system during four consecutive stages of development.

the albedo decreased and stabilised around 0.17, but after the harvesting of the maize the albedo increased sharply again, to an average value of 0.22 representing the remaining cassava crop with soil surface mulching. The simple radiation partitioning model (Eq. [6.9]) fitted the measured albedo values well, explaining 93% of the observed variance (Fig.6.3). The corresponding values of soil albedo were 0.07 and 0.28 for bare and mulched soil, respectively, whereas a value of 0.18 was determined for the crops. Diurnal patterns of albedo are shown in Fig. 6.4 for four consecutive periods corresponding to the initial growth stage, almost closed canopy, senescence of maize and remaining cassava crop, respectively. The albedo stayed rather uniform throughout the day at the start of the cropping season but became very variable for the solitary cassava phase, with intermediate patterns during the other two periods. Net radiation was measured only from 83 DAP onwards. Daytime net radiation on a half-hourly basis (W m-2) was closely related to net short-wave radiation (Rs,n=Rs,9-Rs,8) according to:
R N = 0.92 Rs ,n 24

(r2=1.00, N=3071)

[6.14]

For night-time conditions an average upward net radiation of -21.5 ("17.8) W m-2 was found. This value and Eq. [6.14] were used to estimate half-hourly values of net radiation whenever direct measurements were lacking. For daily totals (values still expressed in W m-2), the relationship between the two types of radiation was given by:
R N = 0.78Rs ,n 14

(r2=0.98, N=122)

[6.15]

Average net radiation was 10.4 ("3.1) MJ m-2 d-1 or 66% of 15.8 ("4.0) MJ m-2 d-1 incoming radiation. The lowest and highest measured daily net radiation totals were 3.6

85

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

600
(a) RN G

500
R N ,G, E, H (W m )
-2

400 300 200 100 0 -100 6 9 12 Time (hrs) 15

8 E
H

18

1.0
(b)

0.8 0.6

initial growth maize senescence overall mean

near canopy closure cassava only

0.4 0.2 0.0 6 9 12


Time (hrs)

15

18

Fig. 6.5a-b. Average daytime patterns of (a) net radiation (RN), soil (G), latent (8E) and sensible (H) heat fluxes during the period of micro-meteorological measurements and (b) of the Bowen ratio ($=H/8E) during four consecutive stages of crop development.

and 16.2 MJ m-2, corresponding to incoming radiation amounts of 7.9 and 25.1 MJ m-2, respectively. The micro-meteorological measurements produced almost 6000 half-hourly values of energy balance components covering 121 full days out of 148. The average daily course of energy budget components (Fig. 6.5a) shows a maximum RN of 550 W m-2 around noon after which radiation is slightly suppressed by increasing cloudiness during the remainder of the day. Soil heat flux was -0.09 ("0.11) MJ m-2 d-1 (i.e., upwards) on average or 0.8% of average daily net radiation, with a fairly constant outbound flux of -7.5 W m-2 at night time and a downward flux of 2-3% of net radiation during day-time, reaching a maximum of 18 W m-2 just before noon (Fig. 6.5a). Average

86

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

1.00 0.95 0.90


Et / E0

maize harvest

0.85 0.80 0.75 0.70 0 20 40 60


DAP

80

100

120

140

160

Fig. 6.6. Seasonal course of the ratio Et/E0 during the growth cycle of the mixed maizecassava cropping system (17 November 1998 - 24 April 1999). Maize was harvested on 126 DAP.

latent heat flux increased from zero in the early morning to an average 370 W m-2 (or 0.54 mm h-1) at 11:00 a.m., whereas a maximum value of 635 W m-2 (or 0.94 mm h-1) was measured at the beginning of the cropping season (12 DAP). Just after noon the sensible heat flux (H) reached its average maximum value of 170 W m-2 with an absolute observed maximum of 352 W m-2. The average diurnal course of the Bowen ratio for the four consecutive growth stages is shown in Fig. 6.5b. The Bowen ratio generally decreased to around 0.35 in the early morning, followed by a gradual increase until noon (stabilising around 0.49 in the early afternoon) and a rapid decrease after 15:00 h. Although the general pattern remained the same throughout the cropping season, slightly lower average Bowen ratios were found for the well-developed canopy, whereas higher values occurred during senescence (Fig. 6.5b). The Bowen ratio for daily H and 8E totals was 0.33"0.09. Daily evapotranspiration (Et) was 1.1-5.4 mm with an average of 3.2"0.9 mm d-1 for the entire wet season period. By comparison, Penman E0 was 3.8 "1.0 mm d-1 (1.2-6.4 mm d-1). The ratio of daily Et over Penman E0 may be compared with the crop water use coefficient kc (sensu Doorenbos and Pruitt, 1977) and is shown for all days in Fig. 6.6. Values around 0.95 were found at the beginning of the season when bare soil evaporation dominated, decreasing gradually to around 0.77 when the canopy was densest and remaining more or less constant afterwards. For the entire 148 day wet season period an Et/E0 ratio of 0.84 was found. Evaporation rates during rainfall were also calculated with the BREB-TVBR method for comparison with wet canopy evaporation rates used in the rainfall interception (Chapter 5). This resulted in an average wet evaporation rate of 0.080 mm h-1 for half-hourly daytime intervals with rainfall. Calculations of evapotranspiration during dry and wet intervals using the BREB-TVBR method would suggest the latter to contribute 2.4% to total ET. The simple radiation partitioning approach suggested soil evaporation to contribute ca. 46% to Et over the entire wet-season measurement period. Not unexpectedly, the pattern of the ratio of soil Es over Et closely followed that of gap fraction (the reciprocal

87

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

0.030 0.025 0.020


g a (m s -1 )

0.015 0.010 0.005 0.000 0:00 3:00 6:00 9:00 12:00 15:00 18:00 21:00 0:00

Time (hrs)

Fig. 6.7. Average diurnal course of aerodynamic conductance (ga) during the wet-season (November-April) measuring period. Broken lines indicate one standard deviation from the mean.

of canopy cover). As a result, the relative importance of Es decreased from 100% of Et at the start of the season to ca. 7% at the height of canopy development, increasing afterwards to an average 39% for the stage with cassava only.

6.3.2. Aerodynamic and surface conductance during the wet season The height of the crop mixture increased from 0.11 m at the start of the measurement period (27 November 1998) to 2.20 m at around 80 DAP while the height of the cassava crop increased slightly after the maize harvest, from 1.45 m on 126 DAP to 1.55 m at the end of the period (Fig. 6.2a). Wind speeds were generally low during the NovemberApril period, with daytime averages of 1.1"0.4 m s-1 and 0.6"0.2 m s-1 at night. A total of 40 data pairs of wind speed measured at two heights was available to test different expressions to estimate displacement length (d) and roughness length (z0) from crop height (cf. Eq. [6.7]). The average wind speeds corresponding to this sub-set were 1.7"0.6 m s-1 and 2.1"0.7 m s-1 at 3.4 and 6.0 m height, respectively. The resulting values for d, z0 and, ultimately, ga using Eq. [6.7] are summarised in Table 6.1. The best agreement between predictions based on wind speed measurements made at either height was reached when a relatively low value of z0/hv=0.03 was used. Varying the value of d/hv within reasonable limits hardly affected predicted ga values and good agreement was obtained with d/hv=0.64. In combination with the prevailing low wind speeds, this resulted in low values for aerodynamic conductance: ca. 5 mm s-1 for night-time conditions, increasing to an average 12.9"7.1 mm s-1 at noon and decreasing again during the afternoon (Fig. 6.7). Daytime average values of ga varied between 3.5 mm s-1 at the start of the cropping period to 21.6 mm s-1 when the crop was at its highest in February (85-100 DAP), also because wind speeds were greater at this time (Fig. 6.9a).

88

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

60
initial growth

50 g c (mm s -1 ) 40 30 20 10 0 6 9 12 Time (hrs)

maximum canopy maize senescence solitary cassava

15

18

Fig. 6.8. Average daytime course of surface conductance (gc) for the wet season (NovemberApril) measuring period (solid and dashed lines indicate average and standard deviations, respectively) and average daytime trends during four consecutive periods.

The overall average wet-season (November-April) aerodynamic conductance was 9.4"4.7 mm s-1. Bulk surface conductance (gc) was determined for dry canopy conditions using the inverse form of the Penman-Monteith equation in combination with calculated values of ga and 8E. The resulting pattern in daytime gc values for four consecutive growth phases is shown in Fig. 6.8. There appeared to be no strong contrasts between periods: gc generally decreased from about 30 mm s-1 in the early morning to 12 mm s-1 around noon time, after which it remained reasonably constant for a few hours and finally decreased after 15:00 h to around 6 mm s-1. The greatest variation occurred in the early morning, probably because the canopy was sometimes wet from a preceding storm or dew, both of which occurred frequently. In Fig. 6.9, the seasonal trends in values of ga and gc around noon time (11:00-13:00) are compared. The limited variation in gc over time is striking (Fig. 6.9b). To test to what extent this slight variation is explained by variations in LAI of the two crops a simple linear function was fitted to the data (cf. Kelliher et al., 1995). The resulting equation reads: g c = 12.6 + 1.6 LAI maize 2.2 LAI cassava (r2=0.27, N=122) [6.16]

Eq. [6.16] clearly explains little of the observed variability. Moreover, most explained variation can be attributed to the lower surface conductance values after maize harvesting (10"3 mm s-1 versus 13"3 mm s-1, respectively). Correlations between noon-time gc and meteorological variables were also poor. It was decided therefore to use a single average value for gc for the entire wet-season period (Fig. 6.9b). A value of 12.7 mm s-1 produced the best agreement between measured and modelled half-hourly latent heat fluxes (r2=0.975, N=1781).

89

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

40

(a)
g a (mm s -1 )
30 20

ga ga u

4 3 2

10 0 0 20 40 60 DAP 80 100 120 140

1 0 160

40

(b)
30

Eq. [6.16]

g c (mm s )

average

-1

20 10 0 0 20 40 60 80 100 120 140 160

DAP

Fig. 6.9a-b. Trends in (a) aerodynamic conductance (ga) and windspeed (u), and (b) surface conductance (gc) of mixed crops during the wet season (November-April) measuring period. Average gc and the relationship given by Eq. [6.16] plotted for comparison in lower panel.

6 Penman-Monteith E t (mm d -1 ) 5 4 3 2 1 0
0 1 2 3 4
-1

Measured E t (mm d )

Fig. 6.10. Agreement between daily Et values calculated via the BREB-TVAR method and values predicted by the Penman-Monteith model using daytime averages of meteorological variables as inputs for the period November-April.

90

u (m s )

-1

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

Variables RN s esat-ea MJ m-2 -1 Pa K Pa d-1 Day-time average / value used coeff. of var. / range 11.0 28% 196 4% 621 31% uz ms
-1

Parameters hv m 1.51 39% 0.64 0.5 0.8 0% -0% 0.03 0.01 0.15 2% -6% 0.25 0.11.0 1% -2% d/hv z0/hv z0h/z0 gc mm s-1 12.7 530 -19% 10%

1.15 30%

Change corresponding to: - St.dev. / low end -25% -1% 1% -3% 3% 2% 1% 1% -1% + St.dev. / high end 25%

Table 6.2. Percentage changes in daily Et as predicted by the Penman-Monteith model after changing average daytime variables by one standard deviation and after using low and high values for selected parameters based on literature values, respectively.

Using the Penman-Monteith model with daytime average values of climatic variables, plant height and a constant gc value produced good estimates of Et (r2=0.973, N=122). Only a very small (0.3%) systematic error was introduced by using average daytime instead of half-hourly values: observed and modelled average daily Et were 3.17 and 3.18 mm d-1, respectively. The relative fractions of Et driven by radiation and vapour demand in the Penman-Monteith equation were also almost identical for half-hourly and daytime input data, with the vapour-demand term contributing 10.4 and 10.9%, respectively. The sensitivity of the Penman-Monteith model to changes in each variable and parameter was investigated by calculating the change in Et after a representative perturbation of the input variable (i.e., one standard deviation for measured variables and published or estimated ranges for the other parameters). The results are summarised in Table 6.2. Net radiation is clearly the most important parameter, stressing the minor importance of vapour-demand driven Et. The same conclusion is reached when considering computed decoupling factors (5 sensu McNaughton and Jarvis, 1983): average daytime values were in excess of 0.6 on all days and often more than 0.8. This indicates a low degree of coupling between the canopy and the atmosphere, which is readily explained by the high air humidity and low wind speeds at the study site.

6.3.3. Soil physical characteristics and soil-vegetation-atmosphere modelling The various soil physical characteristics are summarised in Table 6.3. Average dry bulk density was 924 kg m-3 for the topsoil (0-30 cm) and increased with depth to 1010 kg m-3 below 60 cm. This corresponded with a decrease of porosity from 64 (8)% to 58 (5)%. The percentage of larger pores decreased even more with depth as evident from the change in drainable pore space (defined as the difference in soil moisture contents between saturation and pF=2), from 23 (8)% in the top soil to 11 (7)% below 60 cm. Despite large variations in saturated hydraulic conductivity for individual cores there was a trend for Ksat to decrease with depth, from 7.88.5 m d-1 (N=14) in the top 30 cm of soil, to 2.43.3 m d-1 (N=11) below 60 cm. At 15.2 m d-1 (N=33), topsoil Ksat values

91

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

Depth range Soil physical characteristics 0-30 cm 30-60 cm 60-170 cm Dry bulk density kg m-3 924"90 (21) 971"117 (12) 1010"78 (3) Porosity % 62"10 (12) 60"5 (3) 64"8 (21) a Drainable pore space % 14"8 (8) 11"7 (9) 23"8 (10) 12"2 (6) 14"2 (4) 14"1 (7) Residual moisture content, 2res % m d-1 Ksat (cores) 7.8"8.5 (14) 3.8"6.6 (13) 2.4"3.3 (11) van Genuchten-Mualem parameters Number of data-points 102 b 52 70 % 0.66 0.62 0.63 2sat % 0.08 0.08 0.08 2res c cm-1 0.130 0.039 0.043 n 1.19 1.17 1.18 d K0 m d-1 16.5 5.0d 8.0 e L 5.3 5.3e 5.3 a calculated as the difference between 2 (1 cm) and 2 (100 cm); b excluding TDR-tensiometer data-pairs
e

(N=662); c assumed values; d K0 estimates based on Ksat measurements (see text for explanation); assumed equal to value for 0-30 cm layer.

Table 6.3. Physical characteristics and optimised Van Genuchten-Mualem model parameters for the three soil layers distinguished in soil hydrological modelling. Values between brackets indicate the number of samples.

determined in the field with the double-ring infiltrometer were considerably higher than values measured with a laboratory permeameter. Spatially averaged maximum infiltration rates derived from runoff measurements (cf. Yu, 1999; Chapter 12) were distinctly lower than either of the previous estimates at 2.7 m d-1 (N=18; Table 6.4). All of the cited rates are high enough to accommodate immediate infiltration of most rainfall and prevent saturation of the topsoil for prolonged periods. This corresponds to the low runoff volumes measured from the terrace beds (0.1-4.2%, cf. Section 6.2.4). As expected, unsaturated topsoil conductivities were considerably lower than saturated values: at a suction value of F=3 hPa (pF=0.5) K had decreased to ca. 0.48 m d-1 (or 20 mm h-1) whereas at F=12 hPa (pF=1.1) an average K of 0.076 m d-1 (or 3.2 mm h-1) was found (Table 6.4). Van Genuchten curves were fitted to measured F-2 data-pairs; the resulting parameters are listed in Table 6.3 while the curves are shown in Fig. 6.11. Because optimisation resulted in a negative value for residual moisture content (2res), a fixed value of 8% was used instead, being intermediate between air-dry and oven-dry water contents. In general, the Van Genuchten curves and the field measurements corresponded reasonably well despite the variation between samples (Fig. 6.11). The measured and modelled calibration soil moisture series are compared in Fig. 6.12. The VAMPS model simulated soil moisture content with good accuracy, resulting in a model efficiency (sensu Nash and Sutcliffe (1970)) of 0.85. An almost perfect fit (model efficiency 0.98) was obtained by optimising the Van Genuchten model parameters2sat, 2res, and n to recorded 2-F data pairs (cf. Fig. 6.11a), but the resulting partial curves were unfit to model soil moisture for the rest of the year, because (i) observed soil moisture contents during the calibration period were never less than 2=0.45 and (ii) the double curvature of the retention curve suggested by the combined

92

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

0.7 0.6 0.5 0.4


TDR pF box Field samples

0.3 0.2 0.1

(a) 0-30 cm
0.0 0 1 2 3 4 5 6

pF
0.7 0.6 0.5 0.4
pF box Field samples

0.3 0.2 0.1

(b) 30-60 cm
0.0 0 1 2 3 4 5 6

pF
0.7
pF box

0.6 0.5 0.4

0.3 0.2 0.1

(c) >60 cm
0.0 0 1 2 3 4 5 6

pF

Fig. 6.11a-c. Van Genuchten curves fitted to measured soil water retention (F-2 ) data-pairs for the three distinguished soil horizons.

93

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

0.70 0.65

(0-30 cm)

0.60 0.55 0.50 0.45 0.40 1/16/99

1/26/99

2/5/99

2/15/99

2/25/99

3/7/99

Fig. 6.12. Measured (dots) and modelled (line) volumetric soil water contents (2) between 0 and 30 cm during the calibration period (18 January-3 March 1999). Saturated water content is indicated by the dashed line.

tensiometer-TDR data (cf. Fig. 6.11a). The more realistic parameters eventually used (given in Table 6.3) still produced good agreement between measured and modelled soil water contents values (Fig. 6.12). At 16.5 m d-1 and 5.3, the optimised values of K0 and L were high and their combination resulted in a K(F)-F curve that dropped off very rapidly as soil moisture suction increased (Fig. 6.13). For the soil water contents range observed during the calibration period (2=0.45-0.64) unsaturated conductivity values of 1.9A10-4 to 1.0 m d-1 (8.0A10-3 to 43 mm h-1) were produced, which is still well below saturated hydraulic conductivities measured in the field or on soil cores (Table 6.4). The calibrated model was used to simulate soil water contents for an entire year, including the period after micro-meteorological measurements had ceased (24 April-16 November 1999). A rooting depth of 60 cm produced the best match between observed and modelled soil water contents patterns (Fig. 6.14). The model simulated water suction in the top 10 cm of soil to fall below pF=2.0 between 21 July and 10 October 1999, while soil moisture suction over the entire 60 cm root zone became less than the Et limiting value of pF=3 on 6 August 1999. Subsequently, soil moisture was gradually depleted and finally approached an averaged soil water contents of 2=0.36 for the full 120 cm profile. Soil moisture status and Et were simulated to increase rapidly again after the first substantial storm on 10 October 1999 (Fig. 6.14). Predicted monthly amounts of Es, Ev, Ei, Et and total ET are compared with rainfall (P), drainage (D) and open water E0 in Table 6.5 and Fig. 6.15. For the entire year a total ET of 1228 mm was modelled, divided between 198 mm (16%) of interception losses and 1030 mm (84%) crop transpiration and soil evaporation. Soil evaporation was estimated to be 440 and 368 mm (36% and 30% of ET), depending on whether a normal fraction of soil evaporation or zero soil evaporation were assumed for the period with dry topsoil (21 July-10 October 1999), respectively. Corresponding crop transpiration amounts were 590 mm (48% of ET) and 662 mm (54%). Over the whole year ET was 12% less than Penman open water evaporation (E0), but water use exceeded E0 at the height of canopy
94

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

100 10 K(F) (m d -1 ) 1 0.1 0.01 0.001 0.0001 1 10 F (cm) 100

Fig. 6.13. Relationship between soil water suction (F, absolute value) and hydraulic conductivity [K(F)] for the topsoil, as predicted by the Van Genuchten-Mualem model with optimised values of K0=16.5 m d-1 and L=5.3 (cf. Table 6.3). Average and range of hydraulic conductivity at four suction values, as measured by double ring and tension infiltrometry, are shown for comparison.

Top soil hydraulic properties Ksat (cores) Ksat (double ring) Im (runoff plots) Kunsat (3 cm) Kunsat (6 cm) Kunsat (12 cm)

Average (m d-1) 7.8 15.2 2.7 0.48 0.23 0.076

Range (m d-1) 0.2-15.3 2.8-24.5 0.4-10.6 0.07-1.5 0.05-0.60 0.04-0.18

N observations 12 33 18 7 7 7

Table 6.4. Summary of measurements of saturated (Ksat) and unsaturated (Kunsat) hydraulic conductivity and average maximum infiltration rate (Im, cf. Chapter 12) of topsoil from bench terraces near the experimental site.

development (January-March), also because of substantial interception losses at that time (estimated at 30% of total ET for these months; Table 6.5). After maize harvesting the ratio of ET/E0 decreased because of reductions in both Et (diminishing soil water availability) and interception (less rainfall during the dry season).

95

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

0.7

0.6 (0-120 cm)

30 cm

0.5

0.4
measured (average) measured (range) modelled (30, 90 cm) modelled (60 cm)

0.3

90 cm

0.2 A M J J A S O N

Fig. 6.14. Modelled water content (2) in the top 120 cm of soil for the period April to November 1999, compared with average values and standard deviations based on gravimetric measurements from four locations at and near the experimental sites.

6.4. Discussion 6.4.1. Energy balance components The presently found reflection coefficient of the bare soil surface (0.07) is at the low end of published values but in agreement with the value of 0.08 found for wet volcanic soil on Hawaii (Ekern, 1965, in Ross, 1981) and 0.09 for wet alluvial soil in Nigeria (Oguntoyinbo, 1970). Presumably, the low value is related to high soil water content (cf. Fig. 6.7). Oguntoyinbo (1970) observed a typical decrease of ca. 0.06 (for example, from 0.15 to 0.09) in the albedoes of alluvial to ferruginous soils in Nigeria when going from dry to wet surface conditions. Albedo values reported for maize in temperate and tropical environments are 0.15-0.21, depending on crop stage and density (Uchijima, 1976; Ross, 1981; Oguntoyinbo, 1970), whereas the last author reported an albedo of 0.17 for unnamed mixed crops in Nigeria. The presently found values for the maize-cassava mixture (estimated at 0.18 for a closed canopy) are within the cited range. Reported reflection coefficients for cassava include a value of 0.19 measured by Oguntoyinbo (1970) in Nigeria and 0.18 for a mature stand in Eastern Amazonia measured over two days by Giambelluca (1996; with 0.17 for bare soil) and also for stands of variable density in the Ivory Coast (compared to 0.12 and 0.23 for wet and dry soil, respectively; Montny, 1986). The simple radiation partitioning model of Eq. [6.9] assumed the cassava crop to have an albedo identical to that of the maize-cassava mixture and therefore the high overall albedo found (0.19-0.27; Fig. 6.3) was attributed to the high reflection coefficient of the mulched soil (estimated at 0.28). Albedoes in excess of 0.20 have been reported for grassland experiencing water stress and for senescing cereals (reviewed in Ross, 1981). Much of the high variability in albedo values observed for the stage with mulched cassava (Fig. 6.3) could be explained by incoming radiation:

96

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA

500 400 mm mo -1 300 200 100 0 N* D J F M A M J J A S O N*


P Eo D Total ET

Fig. 6.15. Modelled monthly water budget components rainfall (P), drainage (D) and total ET for mixed crops, along with Penman open water evaporation E0 (* only part of November included both at start and end of model period).

Water budget components


LAI Nova Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nova 0.0 0.3 2.3 3.5 1.8 1.2 1.2 1.2 1.2 1.2 1.2 1.2 1.2 P mm 211 394 434 237 441 299 158 72 36 2 17 220 168 ET mm 56 121 142 116 136 108 98 89 95 90 40 83 55

ET components
Ev mm 0 24 72 76 63 53 52 50 56-69 54-90 24-39 40-49 26 Es mm 56 87 23 8 34 34 34 33 24-36 0-35 0-16 18-26 17 Ei mm 0 9 47 33 38 21 12 6 3 0 1 16 12 Et mm 56 112 95 83 98 87 86 83 92 103 53 70 43 E0 mm 58 127 114 105 122 106 99 95 108 130 149 130 54 Ei/P ET/E0 %

)S
mm -20 -3 22 -14 33 -62 -17 -27 -63 -96 4 153 97

Db mm 175 276 269 134 272 253 77 11 4 8 -27 -16 16

0.1 2.2 10.8 13.9 8.7 7.2 7.3 7.7 7.7 16.6 7.1 7.2 7.4
7.4

0.96 0.95 1.25 1.11 1.11 1.02 0.99 0.94 0.88 0.69 0.27 0.64 1.01
0.88

Year 1.3 2688 1228 8 1452 590-662 368-440 198 a together representing one whole month. b negative drainage indicates simulated upward soil water movement.

1030 1398

Table 6.5. Monthly and annual amounts of modelled (or measured, in the case of Et for November-March and part of April) water budget components as well as the percentage of rainfall intercepted and ET/E0 ratios for the mixed crops between 17 November 1998 and 1999.

97

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA a negative power relationship between the two variables explained 86% of the variance in daily albedo values and 82% for noon-time (11:00-13:00) half-hourly values. This suggests the observed variability to be mainly related to differences in albedo for diffuse and direct radiation. In any case, the overall reflection coefficients for the well-developed maize-cassava and solitary cassava are well above the 0.12-0.13 normally reported for tropical rain forest (Oguntoyinbo, 1970; Pinker et al., 1980; Shuttleworth, 1988). The coefficient of the relationship between daily totals of net radiation and incoming short-wave radiation (Eq. [6.15]) attained a value of 0.78 which is towards the low end of the (wide) ranges published for bare soil (0.71-0.97), grassland (0.74-0.99) and various crops (0.73-1.10) as reviewed by Ross (1981). It is similar to values for maize (0.78-0.87; Stanhill et al., 1968, in Ross, 1981), for various short tropical vegetation types in Ivory Coast (0.77-0.91; Montny, 1986) and for a fire-climax grassland in Fiji (0.81; Waterloo et al., 2002). During the November-April growing season, overall net radiation constituted 66% of incoming radiation, but was only 54% at the height of canopy development. The former value is more than the 53% and 56% measured for bare soil and mature cassava, respectively, in Eastern Brazil and more comparable to values reported for irrigated soil and harvested paddy rice fields (63-64%) in Thailand (Giambelluca et al., 1999). An important reason for the high RN/Rs,9 fraction found in the present study is the low albedo associated with wet soil. In line with their lower albedo, relative amounts of net radiation for tropical forests are somewhat higher than for agricultural crops (58-71%; Pinker et al. 1980; Shuttleworth et al., 1984). At 0.8% of average daily net radiation (from -7.5 W m-2 at night to 18 W m-2 around noon), the magnitude of soil heat fluxes was comparable to values measured by Waterloo et al. (2002) in tall grassland during the wet season in Fiji (<1% of RN) but lower than the 210% of net radiation measured in a grazed (short) Amazonian pasture (Wright et al., 1992; Bastable et al., 1993). For the entire wet season measuring period (27 November-24 April; 148 days), covering growth, ripening and harvesting of the maize and the presence of cassava only thereafter, an average Et rate of 3.1"0.8 mm d-1 was obtained, which included an estimated 2.0"1.0 mm d-1 (64%) of crop water uptake and 1.1"0.9 mm d-1 (36%) of soil evaporation. This represented 74% of net radiation and a seasonal average daily Bowen ratio of 0.33"0.09. The latter is intermediate between the average Bowen ratios reported by Montny (1986) for well-watered cassava (0.26), rice (0.21) and short grass (0.14) in the Ivory Coast on the one hand, and equally well-supplied tall grassland in Fiji (0.63; Waterloo et al., 2002) and grazed pasture in Amazonia (0.43-0.55; Wright et al., 1992; Grace et al., 1998) on the other.

6.4.2. Aerodynamic and surface conductances Based on wind speed measurements at two heights above the crops an optimum ratio of roughness length over vegetation height (z0/hv) of 0.03 was obtained. This value is less than usually reported (cf. Table 6.1). Part of the difference may relate to differences in crop density (cf. Shaw and Pereira, 1982) while also a tendency for the displacement length d to decrease and for z0 to increase with wind speed has been observed in flexible crops like maize (Lemon, 1960; Denmead and McIlroy, 1971; Steduto and Hsiao, 1998a). For example, for a 3 m high maize stand, the latter authors found a z0 of only 0.08 m (z0/hv.0.03) at wind speeds below 1 m s-1, increasing to over 0.26 m (z0/hv.0.09)

98

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA at wind speeds in excess of 4 m s-1. Given the generally low wind speeds prevailing in the study area (1.7 m s-1 at 3.4 m height for the 40 data pairs used), the resulting value of 0.03 for z0/hv indeed seems plausible. Variations in z0 may also result from the use of different d/hv and z0h/z0 ratios and this explains some of the additional differences between studies, although calculated ga values appeared less sensitive to these two parameters in the present case (Table 6.1). Surface conductance (gc) values derived using Eq. [6.6] were remarkably constant, varying between ca.10 and 14 mm s-1, for solitary cassava and mixed crops at the time of maximum canopy cover, respectively (Fig. 6.9b). The overall average value of 12.7 mm s-1 and the observed range in daily values (7-28 mm s-1) agrees well with published data for maize crops in North America. For example, McGinn and King (1990) reported an average gc of 12.0 mm s-1 (6-30 mm s-1) in Ontario, whereas a range of 1.9-10.3 mm s-1 was found by Lemon et al (1971). Steduto and Hsiao (1998b) derived daytime averages of ca. 5-35 mm s-1 for a well-watered maize crop in California. The latter authors ascribed the higher values within the range to the high density of the full-grown crop (maximum LAI of 6-9) compared with other studies. By contrast, the presently studied mixed crop had an LAI that was distinctly lower (<4, Fig. 6.2b) than typically observed for heavily fertilised temperate maize fields. Based on a graph in Steduto and Hsiao (1998b), gc values of 8-14 mm s-1 would be expected for LAI values below 4, which agrees well with the measured range. The observed diurnal pattern in gc closely resembled measurements above a summer maize field in Ottawa (Rochette et al., 1991), where gc decreased from an approximate 30 mm s-1 in early morning to 5-10 mm s-1 around noon and a further slight decrease thereafter (Fig. 6.8). Similar patterns were reported for well-watered maize in California by Held (1991) and Steduto and Hsiao (1998a). The importance of soil moisture status to the value of gc for maize fields is well documented (Lemon et al., 1971; Rochette et al., 1991; Steduto and Hsiao,1998b). Rochette et al. (1991) found daytime averages in excess of 20 mm s-1 for a day following some rainy days, decreasing to less than 5 mm s-1 after a dry period. Surface conductance has also been observed to decrease during maize senescence (McGinn and King, 1990; Rochette et al., 1991; Steduto and Hsiao, 1998), e.g. from around 12.5 mm s-1 to around 8 mm s-1 in the first mentioned study. In a review of field studies Kelliher et al. (1995) concluded that the value of gc is often more conservative than may be expected on the basis of leaf area alone, particularly when modest transpiration by a sparse canopy is compensated by evaporation from wellwatered soil. The structure of the physically-based Shuttleworth-Wallace model reflects this contention. Consistency in gc values was also found in the present study (Fig. 6.9b): gc was already ca. 10 mm s-1 for an almost bare soil and increased little more as the canopy developed. In fact, the greatest contrast in gc values was observed between the pre- and post-harvesting phases (Fig. 6.9b), which may well be related to the presence of harvest residue on the surface increasing the resistance to vapour transfer between the soil and the atmosphere. Under the prevailing humid tropical conditions, the Penman-Monteith model could be used with daytime averages of climatic variables and surface conductance without introducing significant or systematic errors. Both the estimates based on original (halfhourly) data and the daily calculations suggest that Et is strongly driven by radiation, accounting for ca. 90% of total Et.

99

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA 6.4.3. Soil hydrology and soil-vegetation-atmosphere modelling The VAMPS model simulated soil moisture contents during a 44-day calibration period with considerable accuracy (Fig. 6.12). The unrealistic Van Genuchten model parameter values obtained by fitting the model to 2-F data-pairs measured in the field stress the limitations of such field retention curves, in that their use is confined to the range in soil moisture conditions actually measured. The double curvature in the field measurements (Fig. 6.11a) suggests a dual-porosity type soil, in agreement with the structure of the topsoil which consisted of ca. 1-mm sized clay and silt aggregates. Dual porosity can be simulated (e.g. Van Genuchten et al., 1999; van Dijk, 1996), but this increases both the number of parameters to be optimised and the uncertainty of unsaturated hydraulic conductivity values. The optimised soil hydrological parameters (Table 6.3) can be compared with values predicted via texture-based pedotransfer functions, for example those of Rawls et al. (1993). Using both texture (ca. 70% clay, 30% silt, no sand) and bulk density (924 kg m-3) of the topsoil, these functions suggested values of n=1.16 (versus a fitted value of 1.19) and =0.026 (versus 0.13) hPa-1, although it should be noted that Rawls et al. did not recommend the use of their pedotransfer functions for soils with clay fractions greater than 60% and sand fractions smaller than 5%. The ROSETTA model (Schaap et al., 2000) predicted model parameter ranges of 0.015-0.073 and 1.13-1.34 for and n, respectively. Apparently the value of n (which is related to the distribution of pore sizes) is predicted reasonably well by the various transfer functions, but (which is related to the inverse of air entry suction) is much too low compared with optimised values. At 0.039-0.043, values determined for the soil below 30 cm did fall within the predicted range, however. Giving values for K0 and L of 0.8-4.5 m d-1 and -4.5 to -0.8, respectively, the ROSETTA model also predicted much lower hydraulic conductivity values than were determined during calibration in the present study (K0 =16.5 m d-1 and L=5.3; Table 6.3). Such discrepancies can be explained by the loose, well-aggregated structure of the topsoil. The relatively large voids between individual aggregates resulted in a very high saturated hydraulic conductivity but also in a rapid decline in soil water content and conductivity as soil water suction increased. Field observations and model results demonstrated that soil water content was consistently high during periods of frequent rainfall, but rarely close to saturation, even for short periods (Fig. 6.12). Saturated hydraulic conductivity values measured in the field or in the laboratory therefore have limited application when predicting the response of the soil to natural rainfall. At the same time, the optimised K(F)-F function produced estimates of unsaturated hydraulic conductivity that compared reasonably well with measurements made with the tension infiltrometer (Fig. 6.13). The VAMPS model was combined with the interception model presented in Chapters 4 and 5 to estimate ET and its components over a full years growth cycle. Despite possible errors in the estimated dry season Et (e.g. related to changes in albedo and soil heat flux as the soil dried out, or in the assumptions about the limiting effect of soil moisture on Et), the predicted changes in soil water contents during the solitary cassava phase (April-November) were in reasonable agreement with measured gravimetric soil water contents if a rooting depth of 60 cm was assumed for the dry season (Fig. 6.14). Observations by Sommer (2000) in eastern Amazonia suggested that the roots of tree cassava can reach considerably greater depths of down to 150 cm. Micro-meteorological measurements were made again at the end of the dry season (4-14 October 1999) above a cassava field invaded by weeds near the standard climatological
100

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA station (a few hundred metres from the experimental site) to recalibrate the instruments. Calculations along the lines explained in Section 6.2.3 suggested Et during these ten days of 16.7 mm (36% of E0), while the model simulated a very similar Et of 17.4 mm. These findings provide additional support for the models Et estimates. 6.4.4. Total water use and evapotranspiration components Crop coefficients (kc sensu Doorenbos and Pruitt, 1977) calculated as the ratio of daily Et over Penman open water evaporation E0 suggested kc values to decrease from ca. 0.95 for an almost bare soil to around 0.77 during maize senescence, and remaining rather constant for the mulched cassava crop (Fig. 6.6). For the entire measuring period (27 November 1998-24 April 1999; 10-158 DAP) an average value of 0.84 was found, or 1.10 when interception losses were added. By contrast, FAO guidelines suggest an increase from 0.3-0.5 to 1.05-1.2 during canopy development for maize, decreasing again to 0.55-0.6 during senescence (Doorenbos and Pruitt, 1977). This difference arguably reflects drier topsoil conditions prevailing in the FAO crop irrigation requirement studies. Moreover, estimates of intercepted rainfall (or sprinkler irrigation water) are not taken into account explicitly in the FAO approach. As such, the usefulness of FAO crop coefficients is questionable under humid tropical conditions where the soil is wetted frequently by rain and interception makes up an important part of total ET. The present ET estimates correspond well with results from an earlier lysimeter study in the same catchment during the 1994/95 wet season (Van Dijk, 1996). Measurements of rainfall and drainage on three 1.3 m2 percolation tanks planted with inter-cropped maize, cassava and upland rice suggested an average ET of 4.14 mm d-1 during 152 days (20-171 DAP) which compares favourably to the 4.11 mm d-1 for the same cropping phase in the current study. However, simultaneous measurements of throughfall and stemflow on and near the percolation tanks suggested interception to make up 46% of total ET during the 1994/95 measuring period, compared with only 24% for the same phase in the current study. Although the interception losses estimated in the lysimeter study were suspiciously high (representing ca. 13% of gross rainfall), these could not be attributed to measurement errors (see Chapter 5 for discussion). Instead, it was suggested that the exposed position of the 1994/95 experiment (near the top of a ridge) may have enhanced local advection of energy, although the current study does not provide support for this hypothesis. The presently modelled annual ET of 1228 mm is comparable to the 1264 mm yr-1 derived by Sommer (2000) for cassava in Eastern Amazonia using similar measuring and modelling approaches. Blackie (1979) reported an average annual water use of 1030 mm for mixed subsistence crops at a montane site in Kenya. Lower values are normally found for short vegetation types in the equatorial tropics. For example, Williams et al. (1997) estimated an annual ET of 667 mm for a catchment in Central Amazonia of which 80% had been converted to crops and bush fallow. Interestingly, Sommer (2000) derived ET values for similar (3-4 year old) fallow vegetation in Amazonia that were more than twice as high (1411-1458 mm yr-1), again using a methodology similar to the present one. It cannot be excluded that the catchment-based estimate of Williams and Melack (1997) represents an under-estimate: earlier work by Lesack (1993) on the same catchment but in a forested state also yielded a lower ET figure than would be expected for forest, at 1120 mm versus 1310 mm (Shuttleworth, 1988). The presently derived annual total water use of mixed crops also exceeds the water use of tropical grassland, be it well-watered

101

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA short grass in the Congo (1070 mm yr-1; Riou, 1984) or tall natural grassland in the seasonal outer tropics of Fiji (750-900 mm yr-1 ; Waterloo et al., 2002). Soil evaporation appeared to be an important component of overall ET: at the beginning of the cropping cycle (November 1998) the almost bare soil evaporated water at rates as high as 3.5 mm d-1 or 92% of open water evaporation (Fig. 6.6). A similar value of 3.0 mm was measured using small percolation tanks filled with bare soil in Bogor, West Java by Gonggrijp (1941). By contrast, Klinge et al. (2001) estimated annual ET from a cleared and burned patch of rain forest in Eastern Amazonia to be a mere 150-350 mm yr-1 (0.4-1.0 mm d-1), depending on the degree of regrowth. However, soil water availability at the latter site was much lower, whereas the reflection coefficient of the sandy soil was presumably much higher (Klinge, 1998). For the whole year, soil evaporation was estimated to contribute 368-440 mm (30-36%) to total ET. The application of surface mulch after maize harvesting appeared to reduce soil evaporation and total ET by decreasing net radiation (by as much as 14%), but possibly also by increasing the resistance to transfer of vapour between the soil and atmosphere. The effect of mulch application on soil physical conditions such as temperature and water content are well documented (e.g. Lal, 1978; Bussire and Cellier, 1994). The potential importance of soil mulch cover on radiation and energy budgets is self-evident, but little researched in the humid tropical context. This stresses the point made by Giambelluca (1996) that with the exception of pasture (Gash et al., 1996) post-forest land cover parameterisation for GCM modelling is often based on limited data. To assess the effect of mulching on annual water use, Penman-Monteith Et calculations were made accounting for changes in the radiation balance for situations with year-round bare soil and mulch cover, respectively, using corresponding albedo values (cf. Section 6.2.3). This suggested an annual ET for a situation without any mulching that was 57-78 mm (56%) higher than presently estimated, whereas it was 39 mm (3%) lower for year-round mulch cover. These numbers provide a rough indication of the likely effect of mulching on total water use. Annual crops like maize and cassava often replace rain forests (Jepma, 1995), whose annual ET totals are generally in the order of 1350-1450 mm yr-1 (Bruijnzeel, 1990). A value of 1480 mm yr-1 was inferred for mature secondary rain forest elsewhere in West Java but at a lower altitude (120 m a.s.l.) and with a slightly wetter rainfall regime (2850 mm; Calder et al., 1986). No more than 60% of this amount (885 mm) was transpired by the forest, the rest (595 mm) being lost as interception. A similar figure was reported for lowland forest in Peninsular Malaysia (ET=1555 mm yr-1; Et=915 mm; Ei=640 mm; Abdul Rahim et al., 1995). When compared with the present results these findings suggest that, under the maritime tropical conditions prevailing over much of Southeast Asia, the conversion of rain forest to rain-fed annual cropping will result in a decrease in annual ET of at least 150-300 mm (10-20%). Moreover, a major shift in the importance of the evaporation components of interception, vegetation transpiration and soil evaporation can be expected. Interestingly, the presently computed values for annual crop water use exceed ET totals for mature coniferous upland plantation forests elsewhere in Java using the catchment water budget technique (900-1070 mm yr-1; Bruijnzeel, 1988). The contrast is maintained after correcting for differences in elevation: at 0.79-0.84, ET/E0 ratios for these forests were below the presently derived value (0.88; Table 6.5). Nevertheless, it is to be expected that soil water reserves during the dry season will be lowered after reforestation of formerly cropped uplands because the deeper-rooted trees can sustain higher transpiration rates (Bruijnzeel, 1989; Calder, 1998). Simulated rates of soil water
102

CHAPTER 6 - WATER BUDGET OF INTER-CROPPED MAIZE AND CASSAVA percolation to the groundwater reservoir during the dry season were already very limited or even indicated capillary rise under cassava (cf. Fig. 6.10, Table 6.5). The amounts of rainfall needed to replenish the soil at the onset of the rainy season would therefore have to be correspondingly larger after reforestation, also because of the associated higher rainfall interception (Bruijnzeel, 1989). As a result, the time lag between the return of the rains and the subsequent rise in stream baseflow levels would be expected to be longer. Having said so, infiltration opportunities under annual cropping are generally less than under established plantation forest. During the transition months of October and November 14-29 mm (2-4% of 673-734 mm gross rainfall in 1998 and 1999) were observed to run off superficially from the agricultural bench-terraced hillsides in the study area (Chapter 14) and therefore lost to recharge, whereas this was less than 1 mm (0.1% of 697 mm in 2000) below a fast-growing Paraserianthes falcataria plantation nearby. Ultimately, the effect of land cover change on soil water reserves and baseflow levels will depend on the balance between changes in vegetation water use on the one hand, and in infiltration opportunities on the other (Bruijnzeel, 1989, 2002). In the present case, the former seems to outweigh the latter, but local measurements of (plantation) forest water use are needed to quantify the overall effect.

103

Anda mungkin juga menyukai