Next Article in Journal
Comparative Evaluation of Pyrus Species to Identify Possible Resources of Interest in Pear Breeding
Next Article in Special Issue
Plot Layout Method of Field Experiment for Wheat with Border Irrigation Based on Soil Water Content Heterogeneity
Previous Article in Journal
Effect of Applying an Organic Amendment on the Persistence of Tebuconazole and Fluopyram in Vineyard Soils
Previous Article in Special Issue
Multi-Objective Optimization Water–Nitrogen Coupling Zones of Maize under Mulched Drip Irrigation: A Case Study of West Liaohe Plain, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Accurate and Rapid Measurement of Soil Dry Depth Using Ultrasonic Reflection Waves

1
Guangdong Engineering Research Centre for Highly Efficient Utility of Water/Fertilizers and Solar-Energy Intelligent Irrigation, Guangzhou University, Guangzhou 510006, China
2
School of Mechanical and Electrical Engineering, Guangzhou University, Guangzhou 510006, China
3
China-Ukraine Institute of Welding, Guangdong Academy of Sciences, Guangzhou 510650, China
4
School of Physics and Materials Science, Guangzhou University, Guangzhou 510006, China
*
Authors to whom correspondence should be addressed.
Agronomy 2023, 13(5), 1276; https://doi.org/10.3390/agronomy13051276
Submission received: 6 April 2023 / Revised: 27 April 2023 / Accepted: 27 April 2023 / Published: 28 April 2023
(This article belongs to the Special Issue Water-Saving in Agriculture: From Soil to Plant)

Abstract

:
Soil dry depth is a key parameter that determines soil fertility and nutrient availability, ultimately affecting crop yield and quality. However, accurately measuring the dry depth of soil has been a challenge. In this work, we propose using reflective ultrasonic waves to measure dry depth in soil. Four soil types, including clay, sandy loam, silty loam, and sandy were prepared and the feasibility of the method was demonstrated through theoretical analysis. An experimental measurement system was established to verify the consistency between ultrasonic measurements and manually measured values. Two statistics were used in Ordinary Least Squares (OLS) regression to evaluate the model fit: R-square (R2) and Root mean square error (RMSE). The results indicate that the proposed method provides a higher accuracy in estimating the dry depth of sandy loam and silty loam (R2 values of 0.9899 and 0.992 for sandy loam, RMSE values of 1.57% and 1.5% for silty loam) than those of the clay and sandy samples (R2 values of 0.9896 and 0.9874 for clay, RMSE values of 1.66% and 1.77% for sandy). The maximum measurement errors for all the soil type predictions are below 6%; the overall accuracy was acceptable. Our findings suggest that ultrasonic measurement is an efficient and cost-effective approach for measuring soil dry depth, which could enable the precise control of irrigation water usage and the conservation of valuable water resources.

1. Introduction

Water is a crucial and indispensable element for the growth of plants. Appropriate water irrigation can promote the healthy growth of crops and enhance yield and quality. However, the conventional method of flood irrigation used in field production often leads to excessive irrigation and low water use efficiency [1]. Excessive irrigation can result in various problems, such as root hypoxia and root rot [2,3], which ultimately reduce crop yields. Therefore, monitoring soil moisture is essential as it is one of the most important indicators for determining soil fertility and nutrition.
The classic method of measuring soil moisture is the drying method [4], which provides a high accuracy and is independent of soil type and salinity. However, this method does not allow a repetitive measurement [5]. In addition, it is time-consuming and laborious, making it unsuitable for continuous monitoring [6,7]. In recent years, several new methods have been proposed, including the tensiometer [8,9], the neutron probe [10,11,12], gamma radiation [13,14,15], infrared remote sensing [16,17], and dielectric methods [18,19,20,21]. For instance, Masoud et al. [22] used the angular distribution of scattered neurons to measure soil moisture with small relative errors ranging from 2 to 10%. Zhao et al. [23] used four algorithms with L-band radiometry at different angles and found that the accuracy of all the algorithms achieved their best performances at intermediate incidence angles of 40° to 45°. Calamita et al. [24] estimated soil moisture in eight different sites using Resistivimeter Syscal Junior and a portable Time domain Reflectometer (TDR), and found that the resistivity measurements were superior in inferring soil moisture spatial and temporal variability with an average RMSE of 4.4%. In general, tensiometers are cost effective and non-destructive, and they can provide continuous measurement without distressing the soil. However, tensiometers require frequent maintenance to supply water to the detector [25]. Although neutron probes and gamma radiation can provide fast and accurate soil moisture measurements at a fixed point, the equipment is expensive and the radiation may bring health risks [26,27]. Infrared remote sensing has a low penetration depth, and the response is significantly affected by environment and climate conditions [28]. Moreover, the dielectric moisture sensors (i.e., TDR, FDR, SWR sensors) are also expensive and their applicability in highly saline soils is limited.
Despite many efforts to measure soil moisture, the dry depth of soil is a more critical parameter in determining how much water and nutrients a crop can absorb. This parameter refers to the depth from the soil surface to the interface of the dry and wet soil, where dry soil has moisture content below 5% and wet soil has a moisture over 5%. The accuracy of the tools and techniques in measuring soil dry depth depends on many factors, such as soil texture, temperature, wind, and measurement methodology. Currently, the measurement of soil dry depth is poorly investigated, and limited efforts have been made to measure frozen soil depth using techniques such as ground penetrating radar [29], electrical resistivity tomography [30], or borehole thermometry [31]. Nevertheless, these methods are expensive and are not suitable for measuring soil dry depth. Developing a low-cost method for measuring dry depth would significantly improve crop yield and quality while conserving water resources. Recent evidence suggests that ultrasonic waves can be a technique to solve this issue. For instance, Chen et al. [32] revealed the correlation mechanism between the law of ultrasonic propagation in coal samples and the migration of water. Tanaka et al. [33] improved the measurement accuracy of soil moisture using an ultrasonic waveguide to predict rainfall-induced slop failure. These studies verified that ultrasonic waves can transmit valuable signals when propagating in substances containing a certain amount of humidity.
Inspired by the aforementioned studies, we have proposed a new method to measure the dry depth of soil using reflective ultrasonic waves. The suitable frequency and propagation time were initially determined through simulation, and the relationship between the reflection time and dry depth was established. Subsequently, an experimental measurement system was implemented to validate the feasibility of the proposed technique. Our findings demonstrate excellent performance, with R2 values exceeding 0.98 and RSEM values below 1.8% for all soil types, surpassing those reported in previous research. This study represents a significant advancement in the field of soil dry depth measurement.

2. Materials and Methods

2.1. Theoretical Analysis

Soil is a complex and heterogeneous mixture of solid, liquid, and gas components, with pores of various sizes within a solid skeleton. Soil moisture is mainly found in small and medium pores, while large pores are filled with air. Dry soils are defined as those with a moisture content below 5%, as depicted in Figure 1. The depth of dry soil on the surface can be used as a critical point for evaluating drought stress on crops. This measurement takes into account factors such as surface soil water loss and crop absorption, and is expressed in millimeters of depth rather than a water content percentage. It provides a more accurate characterization of soil moisture than traditional water content indices, as it has a clear concept and spatial scale dimension.
By measuring the depth of the dry soil layer, the critical point between wet and dry soil can be determined, which can help meet crop water needs under specific conditions. The use of soil dry depth in intelligent irrigation systems can improve water management by accurately determining drought stress duration and regulating deficit irrigation parameters. Moreover, the physical quantity of drying depth is practical and relevant to current engineering practices, making it an attractive option for the intelligent manufacturing of agricultural machinery and equipment.
In the process of ultrasonic pulse waves, the distance covered by the wave in one unit of time is referred to as the wave velocity, which can be expressed by:
c = λ T = λ f
where λ is the wavelength of the ultrasonic pulse wave, T is the ultrasonic pulse wave period, and f is the ultrasonic pulse frequency.
During the propagation of ultrasonic waves in a medium, the energy is attenuated, leading to the formation of an ultrasonic field. The ultrasonic field is an area of a certain size and shape in which the energy of the ultrasonic pulse wave is distributed. The instantaneous pressure difference at a certain point in the ultrasonic field, with or without ultrasonic energy, is called the sound intensity, which is represented by:
P = ρ c A ω sin t x c = P m sin t x c
P m = ρ c A ω = 2 π ρ c A f
where Pm is the amplitude of sound pressure, ρ is the density of the medium, c is the velocity of ultrasonic wave, and A is the amplitude of the ultrasonic wave. ω is the angular frequency, t is the time it takes for the ultrasonic wave to be detected, and x is the current displacement.
The velocities of longitudinal and transverse waves are:
c L = λ + 2 μ ρ 1 2 = 1 v ( 1 + v ) ( 1 v ) × E P 1 2
c T = μ ρ 1 2 = E 2 ρ ( 1 + v ) 1 2
where µ is the Lamé constant related to the elastic properties of the medium, E is acoustic energy, cL and cT is wave velocity of longitudinal and transverse waves, and v is the velocity of wave.
It is evident that the wave velocity in a solid medium is solely dependent on its physical properties. Specifically, an increase in the elastic modulus and Poisson’s ratio of the medium leads to a decrease in its density, resulting in a reduction of the wave velocity of an ultrasonic pulse. Furthermore, it is important to note that the longitudinal wave velocity is greater than the transverse wave velocity. Acoustic impedance is the ratio of sound pressure and vibration of a particle in an ultrasonic field, which is represented by:
Z = P v = ρ c A ω v = ρ c v v = ρ c
where P is acoustic strength, ω is angular frequency. t serves as a measure of a medium’s resistance to the propagation of sound waves during the ultrasonic pulse process. The value of acoustic impedance is equal to the product of the medium’s density and velocity.
Sound intensity is the sum of energy vertically passed by ultrasonic wave in unit time and unit area, and is a physical dimension used to express the intensity of sound wave energy, which is expressed by:
I = 1 2 ρ c A 2 ω 2 = 1 2 Z v 2 = P 2 2 Z
The acoustic pressure and acoustic intensity of the incident wave are P0 and I0, respectively. The sound pressure and intensity of the reflected wave are Pr and Ir, respectively. The sound pressure and sound intensity of the projected wave are Pt and It, respectively. The ratio of Pr to the P0 on the contact surface between the first medium and the second medium is called the sound pressure reflectivity of the contact surface, which is represented by:
r = p r p 0
The ratio of Pt to the P0 is called the sound pressure transmissivity of the contact surface, and is denoted by:
t = p t p 0
It is known that the sound pressure reflectance r and the sound pressure transmittance t are:
r = P r P 0 = Z 2 Z 1 Z 2 + Z 1
t = P t P 0 = 2 Z 2 Z 2 + Z 1
where Z1, Z2 are the acoustic impedances of the first and second media, respectively.
The ratio of Ir to I0 is called the acoustic intensity reflectivity of the contact surface, and is denoted by:
R = I r I 0 = P r 2 2 Z 1 P 0 2 2 Z 2 = P r 2 P 0 2 = r 2 = Z 2 Z 1 Z 2 + Z 1 2
The ratio of It to I0 is referred to as the sound intensity reflectivity of the contact surface, and is denoted by:
T = I t I 0 = P t 2 2 Z 1 P 0 2 2 Z 2 = P t 2 P 0 2 = r 2 = 4 Z 2 Z 1 Z 2 + Z 1 2
It can be seen from the above formula that when Z1 > Z2, the acoustic pressure reflectivity r < 0, and the transmittance of the second medium is smaller than that of the first medium, part of the incident ultrasonic wave will be reflected back, and part of the incident ultrasonic wave will pass through the contact surface of the two media.
According to the propagation characteristics of ultrasonic waves, when an ultrasonic wave travels from dry soil to wet soil, it encounters a boundary or interface between the two media. The propagation speed of ultrasonic waves is different in dry and wet soil, and the density of the two media is also different. As a result, the ultrasonic wave has different acoustic impedance in the two media. At the interface between the two media, a reflected wave is generated in the opposite direction of the incident wave, and a transmitted wave travels in the same direction as the incident wave, as shows in Figure 2.
As a mechanical elastic wave, ultrasonic energy will be attenuated in the process of propagation along soil medium. This attenuation occurs due to the friction caused by particle vibration and thermal sensing within the medium. In addition to absorption attenuation, the ultrasonic beam will also diffuse in the medium as it propagates, leading to a reduction in ultrasonic intensity within the soil medium.
Figure 3 illustrates the method used for ultrasonic sensing, which involves measuring the ultrasonic reflection wave at the interface between dry and wet soil. This method allows for direct measurement of the depth of dry soil without the need for artificial destruction, as the transmitter and receiver are placed at the same level. This approach is convenient and efficient, as it eliminates errors caused by human disturbance of the soil layer during sensing. Furthermore, the accuracy of the sensing results is significantly improved through this non-invasive approach.

2.2. Simulation Description

To investigate the relationship between the dry depth of soil and ultrasonic reflection time, the propagation of ultrasonic waves in soil has been simulated in Comsol Multiphysics software via a finite element method (FEM). A 2D soil model with a dimension of 180 mm × 180 mm was created and divided into two layers, as depicted in Figure 4. The left and right sides corresponded to dry soil and wet soil, respectively, and the bisector of the model is set as the dry–wet separation interface. The four sides of the square were set to a low reflection boundary condition by applying an absorbing layer to reduce the non-physical reflections in the physical domain. The dry and wet soil are set with a moisture content of 5% and 25%, respectively, and the corresponding soil density and sound velocity are listed in Table 1.

2.3. Sample Preparation

Soil sample preparation was performed at the campus of Guangzhou university (23.030 1.23″ N, 113.240 3.92″ E), and the location is shown in Figure 5. Four different types of soil were prepared, including sandy, sandy loam, clay, and silty loam (See Figure 6). These soils were collected by a standard shovel and then placed in a seal pocket with a label. Forty samples were collected for each soil type and a total of 160 soil samples were prepared. The collected samples were transported to the Solar Energy Intelligent Irrigation Equipment Technology Innovation Center Laboratory of Guangzhou University for processing. All the samples were initially air-dried for 72 h with a light intensity of 1.5 × 104 lux and a wind speed of 15 m/s, and then passed through a 2 mm soil sieve to remove gravel and plant roots. The filtered soil samples were placed in rectangular glass tanks, with soil depth set at 60~300 mm. A Dn20 drip irrigation belt with a flow rate of 0.2 L/s was deployed to create different dry depth conditions. After irrigation, the soil was stood for 2.5 h to allow the surface soil moisture to volatilize and sink. A soil-dry-depth measuring system was subsequently established, as depicted in Figure 7. Ultrasonic sensors were placed on the soil surface and an ultrasonic controller was used to collect the data. The wire communication module realized the transmission of the information, and Raspberry Pi was used to for reading and computing sensing data. The practical dry depth of the soil was measured manually using soil moisture sensor (LD-SW, Lynd Intelligent Technology Co., Ltd., Weifang, China).

3. Results

Figure 8 reports the propagation velocity of the ultrasonic waves of different frequencies in soil. It was observed that the propagation speed is relatively low at 40 Hz, and the ultrasonic wave cannot reach the interface within 20 µs. As the frequency increases from 40 Hz to 80 Hz, the propagation speed gradually increases. This phenomenon can be explained by the fact that the propagation speed of an ultrasonic wave in the same medium is proportional to the wavelength and frequency. Typically, the propagation direction of the ultrasonic wave is along the x-axis. However, when an ultrasonic wave encounters an interface between two soil mediums, a portion of it is reflected. This reflection is mainly due to the difference in acoustic impedance between the two mediums.
Figure 9 reports the pressure distribution of the ultrasonic waves propagating in soil at different frequencies. It was found that the pressure was increased as the propagation time increased from 20 µs to 40 µs. This can be attributed to the fact that the ultrasonic wave passes through two media with different acoustic impedance, leading to the formation of a superimposed wave of direct waves and reflected wave and resulting in an increase in local pressure. A slightly defected wave was observed during the propagation of a second medium, which also can be attributed to the difference in acoustic impedance. As time increased to 60 μs, the ultrasonic wave had almost completely passed through the interface, and the pressure in the second medium dropped significantly. When time reached 80 μs, the pressure in most regions became zero, except for the sound source area of the ultrasonic waves on the left. This is because if the boundary is set as air medium, the ultrasonic wave will slowly decay in the finite space until it disappears, and there will be no second reflection.
In general, although a higher frequency of ultrasonic wave results in a high propagation speed, it also induces a high pressure. To ensure the accuracy of the measurement, the propagation time of the reflected ultrasonic wave in the soil medium is preferably more than 20µs, and a frequency of 60 Hz was selected for these experiments.
Figure 10 illustrates the relationship between reflection time and dry depth for different soil samples, accompanied by a linear fit. It can be seen that the reflection time is almost proportional to the dry depth, especially for sandy loam and silty loam samples, which exhibited high determination coefficients of 98.24% and 97.4%, respectively. Notably, for a fixed dry depth, the reflection time of clay is comparatively lower than that of sandy loam, silty loam, and sand samples. This can be attributed to its fine particles, numerous pores, and weak water permeability, which contribute to an accelerated propagation speed of ultrasonic waves.
Figure 11 shows the error analysis between the ultrasonic and manual measurement of dry depth for different soil samples. It was found that the ultrasonically and manually measured values reach a great agreement for the sandy loam and silty loam samples with only a minor discrepancy observed in the regression model. The corresponding R2 values are 0.9899 and 0.0990, and the RMSE values are 0.01572 and 0.01599, respectively. Although a high accuracy was also obtained for sandy loam (R2 values of 0.09896, RMSE value of 0.0166) and silty loam (R2 values of 0.09874, RMSE value of 0.0177) samples, significant differences were observed for clay and sand samples in certain measurements. Specifically, errors were prominent in the fifth, eleventh, and fourteenth measurements for clay samples and the thirty-fifth and thirty-sixth measurements for sandy samples. The effectiveness of the proposed method can vary across different soil types, which is attributed to variation in the physical properties of soil. The speed of an ultrasonic wave through soil depends on its density and elasticity, which are impacted by texture and compaction. These differences can cause the wave to scatter or to be absorbed differently, resulting in variations in the quality of the reflected signal. In clay and sandy soils, ultrasonic waves are subject to more scattering and attenuation when propagating inside the soil due to the tightly packed arrangement of particles and relatively low porosity. As a result, the reflected signals are weak. In contrast, sandy loam and silt loam soils have larger soil particles, higher porosity, and a lower soil density, which allows ultrasonic waves to propagate in a more direct and clear path within the soil. This leads to relatively stronger reflected signals. In addition, since sandy and clay soil contain a high content of sand, the water in these soils evaporates easily, resulting in higher ultrasonic measurement values compared with manual measurements.
The maximum measurement errors for each soil sample is listed in Table 2. The maximum measurement error in sandy loam and sandy is larger than in the silt and clay samples with values of 5.97% and 5.98, respectively, while the maximum measurement error of clay and silty loam is 5.67% and 5.89, respectively. Despite these measurement errors, the overall accuracy is acceptable. These findings demonstrated the feasibility of using ultrasonic waves to measure the dry depth of the soil.

4. Discussion

Table 3 presents a comparison of the measurement accuracy of the proposed method with other techniques. TDR and capacitance sensor show a good performance with an R2 value > 0.87, but the maximum RMSE values exceed 2%. The neutron probe method exhibits some uncertainties, with RMSE values ranging from 2%t to 6.7%. The electrical resistivity method shows relatively low R2 values (i.e., 0.21~0.68), although a lower RMSE value can be realized (i.e., 1.15%). In contrast, the proposed method in this work achieved R2 values of over 0.98 and RMSE values less than 1.8%. This outstanding performance further demonstrates the potential of the proposed method for water management, soil remediation, and precision farming. Moreover, the non-invasive and low-cost implementation make it an attractive option for field applications.
However, there are limitations to the proposed method. First, the measurement accuracy could be impacted by soil moisture of dry oil. In extremely dry soil with less than 1% moisture, the lack of moisture can create a relatively uniform density throughout the soil, making it difficult for the ultrasonic waves to reflect back to the surface. This can lead to inaccurate readings or an inability to measure the dry depth of the soil at all. Second, soil type can affect the accuracy of ultrasonic reflective wave measurements, with high levels of clay or organic matter scattering ultrasonic waves, and making it difficult to obtain accurate readings.
It is worth noting that this study only considered four soil types and carried out measurements at room temperature, while many factors, such as wind, temperature, and sunlight were not taken into account. Future research should aim to improve the applicability of this method by considering all soil types and accounting for the effect of environmental factors.

5. Conclusions

In this study, a method based on a reflective ultrasonic wave was proposed to measure the dry depth of soil. A theoretical investigation was initially conducted to demonstrate the feasibility of such a method. A frequency of 60 Hz and a suitable propagation time of greater than 20 µs was selected. Then, an experimental measurement system was established to validate the consistency between ultrasonic measurements and manually measured values. The results show that the ultrasonic measurements show a great agreement with the manual measurements. The resulted R2 values and RSME values are over 0.98 and below 0.3%, respectively, which is better than some of the methods previously reported. Our findings provide evidence that the proposed ultrasonic measurement could be an efficient and low-cost method for measuring soil dry depth, which helps to improve the usage of water resources.

Author Contributions

Conceptualization and methodology, J.X.; software, validation, and formal analysis, S.L.; resources, data curation, and visualization, C.Z.; investigation and writing—original draft preparation, Z.L.; writing—review and editing, Y.Z.; supervision, project administration, and funding acquisition, Z.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (51975136, 52075109), the Science and Technology Innovative Research Team Program in Higher Educational Universities of Guangdong Province (2017KCXTD025), the Industry-University-Research Collaborative Innovation Base of Ministry of Education (220903950010408), Special Research Projects in the Key Fields of Guangdong Higher Educational Universities (2019KZDZX1009), Natural Science Foundation of Guangdong Province (2023A1515011723), the Tertiary Education Scientific research project of Guangzhou Municipal Education Bureau (202235139), and Guangzhou University Research Project (YJ2021002).

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

We thank the editors for their hard work and the referees for their comments and valuable suggestions that helped to improve this paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jaiswal, S. Fuzzy inference based irrigation controller for agricultural demand side management. Comput. Electron. Agric. 2020, 175, 105537. [Google Scholar] [CrossRef]
  2. Bhattarai, S.P.; Su, N.; Midmore, D.J. Oxygation unlocks yield potentials of crops in oxygen-limited soil environments. Adv. Agron. 2005, 88, 313–377. [Google Scholar]
  3. Bhattarai, S.P.; Midmore, D.J.; Pendergast, L. Yield, water-use efficiencies and root distribution of soybean, chickpea and pumpkin under different subsurface drip irrigation depths and oxygation treatments in vertisols. Irrig. Sci. 2008, 26, 439–450. [Google Scholar] [CrossRef]
  4. Schmugge, T.J.; Jackson, T.J.; McKim, H.L. Survey of methods for soil moisture determination. Water Resour. Res. 1980, 16, 961–979. [Google Scholar] [CrossRef]
  5. Zazueta, F.S.; Xin, J.N. Soil moisture sensors. Florida cooperative extension service. Bulletin 1994, 292, 1–11. [Google Scholar]
  6. Patil, V.S.; Shinde, S.A.; Dhawale, N.M. A review on determination of soil organic matter and soil moisture content using conventional methods and image processing techniques. In Proceedings of the 2021 IEEE Pune Section International Conference (PuneCon), Pune, India, 16–19 December 2021; pp. 1–6. [Google Scholar]
  7. Fu, Y.; Taneja, P.; Lin, S.; Ji, W.; Adamchuk, V.; Daggupati, P.; Biswas, A. Predicting soil organic matter from cellular phone images under varying soil moisture. Geoderma 2020, 361, 114020. [Google Scholar] [CrossRef]
  8. Singh, D.N.; Kuriyan, S.J. Estimation of hydraulic conductivity of unsaturated soils using a geotechnical centrifuge. Can. Geotech. J. 2002, 39, 684–694. [Google Scholar] [CrossRef]
  9. Vanapalli, S.K.; Fredlund, D.G.; Pufahl, D.E. Influence of soil structure and stress history on the soil-water characteristics of a compacted till. Geotechnique 2001, 51, 573–576. [Google Scholar] [CrossRef]
  10. Jayawardane, N.S.; Meyer, W.S.; Barrs, H.D. Moisture measurement in a swelling clay soil using neutron moisture meters. Soil Res. 1984, 22, 109–117. [Google Scholar] [CrossRef]
  11. Li, J.; Smith, D.W.; Fityus, S.G. The effect of a gap between the access tube and the soil during neutron probe measurements. Soil Res. 2003, 41, 151–164. [Google Scholar] [CrossRef]
  12. Evett, S.R. Some aspects of time domain reflectometry, neutron scattering, and capacitance methods for soil water content measurement. In Proceedings of the Joint FAO/IAEA Consultants Meeting on Comparison of Soil Water Measurement Using the Neutron Scattering, Time Domain Reflectometry and Capacitance Methods, Vienna, Austria, 23–25 November 1998. [Google Scholar]
  13. Jones, W.K.; Carroll, T.R. Error analysis of airborne gamma radiation soil moisture measurements. Agric. Meteorol. 1983, 28, 19–30. [Google Scholar] [CrossRef]
  14. Cho, E.; Jacobs, J.M.; Schroeder, R.; Tuttle, S.E.; Olheiser, C. Improvement of operational airborne gamma radiation snow water equivalent estimates using smap soil moisture. Remote Sens. Environ. 2020, 240, 111668. [Google Scholar] [CrossRef]
  15. Bogena, H.R.; Huisman, J.A.; Güntner, A.; Hübner, C.; Kusche, J.; Jonard, F.; Vey, S.; Vereecken, H. Emerging methods for noninvasive sensing of soil moisture dynamics from field to catchment scale: A review. Wiley Interdiscip. Rev. Water 2015, 2, 635–647. [Google Scholar] [CrossRef]
  16. Robinson, D.A.; Campbell, C.S.; Hopmans, J.W.; Hornbuckle, B.K.; Jones, S.B.; Knight, R.; Ogden, F.; Selker, J.; Wendroth, O. Soil moisture measurement for ecological and hydrological watershed-scale observatories: A review. Vadose Zone J. 2008, 7, 358–389. [Google Scholar] [CrossRef]
  17. Terhoeven-Urselmans, T.; Schmidt, H.; Joergensen, R.G.; Ludwig, B. Usefulness of near-infrared spectroscopy to determine biological and chemical soil properties: Importance of sample pre-treatment. Soil Biol. Biochem. 2008, 40, 1178–1188. [Google Scholar] [CrossRef]
  18. Stacheder, M.; Koeniger, F.; Schuhmann, R. New dielectric sensors and sensing techniques for soil and snow moisture measurements. Sensors 2009, 9, 2951–2967. [Google Scholar] [CrossRef]
  19. Gaskin, G.J.; Miller, J.D. Measurement of soil water content using a simplified impedance measuring technique. J. Agric. Eng. Res. 1996, 63, 153–159. [Google Scholar] [CrossRef]
  20. Kelleners, T.J.; Soppe, R.W.O.; Robinson, D.A.; Schaap, M.G.; Ayars, J.E.; Skaggs, T.H. Calibration of capacitance probe sensors using electric circuit theory. Soil Sci. Soc. Am. J. 2004, 68, 430–439. [Google Scholar] [CrossRef]
  21. Fares, A.; Polyakov, V. Advances in crop water management using capacitive water sensors. Adv. Agron. 2006, 90, 43–47. [Google Scholar]
  22. Ghaemifard, M.; Ghal-Eh, N.; Najafabadi, R.I.; Vega-Carrillo, H.R. Angular distribution of scattered neutrons as a tool for soil moisture measurement: A feasibility study. Appl. Radiat. Isot. 2020, 160, 109131. [Google Scholar] [CrossRef]
  23. Zhao, T.; Hu, L.; Shi, J.; Lü, H.; Li, S.; Fan, D.; Wang, P.; Geng, D.; Kang, C.S.; Zhang, Z. Soil moisture retrievals using l-band radiometry from variable angular ground-based and airborne observations. Remote Sens. Environ. 2020, 248, 111958. [Google Scholar] [CrossRef]
  24. Calamita, G.; Brocca, L.; Perrone, A.; Piscitelli, S.; Lapenna, V.; Melone, F.; Moramarco, T. Electrical resistivity and tdr methods for soil moisture estimation in central italy test-sites. J. Hydrol. 2012, 454, 101–112. [Google Scholar] [CrossRef]
  25. Chow, L.; Xing, Z.; Rees, H.W.; Meng, F.; Monteith, J.; Stevens, L. Field performance of nine soil water content sensors on a sandy loam soil in new brunswick, maritime region, canada. Sensors 2009, 9, 9398–9413. [Google Scholar] [CrossRef] [PubMed]
  26. Van Bavel, C.H.M.; Underwood, N.; Swanson, R.W. Soil moisture measurement by neutron moderation. Soil Sci. 1956, 82, 29–42. [Google Scholar] [CrossRef]
  27. Greacen, E.L. Soil Water Assessment by the Neutron Method; CSIRO: Melbourne, Australia, 1981. [Google Scholar]
  28. Yin, Z.; Lei, T.; Yan, Q.; Chen, Z.; Dong, Y. A near-infrared reflectance sensor for soil surface moisture measurement. Comput. Electron. Agric. 2013, 99, 0168–1699. [Google Scholar] [CrossRef]
  29. Hinkel, K.M.; Doolittle, J.A.; Bockheim, J.G.; Nelson, F.E.; Paetzold, R.; Kimble, J.M.; Travis, R. Detection of subsurface permafrost features with ground-penetrating radar, barrow, alaska. Permafr. Periglac. Process. 2001, 12, 179–190. [Google Scholar] [CrossRef]
  30. Dafflon, B.; Hubbard, S.; Ulrich, C.; Peterson, J.; Wu, Y.; Wainwright, H.; Kneafsey, T.J. Geophysical estimation of shallow permafrost distribution and properties in an ice-wedge polygon-dominated arctic tundra regionimaging permafrost soil properties. Geophysics 2016, 81, WA247–WA263. [Google Scholar] [CrossRef]
  31. Lachenbruch, A.H.; Marshall, B.V. Changing climate: Geothermal evidence from permafrost in the alaskan arctic. Science 1986, 234, 689–696. [Google Scholar] [CrossRef]
  32. Chen, J.; Cheng, W.; Wang, G.; Li, H. Correlation mechanism between the law of ultrasonic propagation in coal samples and the migration of water. Fuel 2022, 310, 122264. [Google Scholar] [CrossRef]
  33. Tanaka, K.; Hiraoka, N.; Nakano, S.; Kameda, T.; Fujimoto, M.; Fukagawa, R. Improved measurement of soil moisture using an ultrasonic waveguide to predict rainfall-induced slope failure. Jpn. J. Appl. Phys. 2015, 54, 10ND04. [Google Scholar] [CrossRef]
  34. Huisman, J.A.; Hubbard, S.S.; Redman, J.D.; Annan, A.P. Measuring soil water content with ground penetrating radar: A review. Vadose Zone J. 2003, 2, 476–491. [Google Scholar] [CrossRef]
  35. Robinson, D.A.; Jones, S.B.; Wraith, J.M.; Or, D.; Friedman, S.P. A review of advances in dielectric and electrical conductivity measurement in soils using time domain reflectometry. Vadose Zone J. 2003, 2, 444–475. [Google Scholar] [CrossRef]
  36. Ochsner, T.E.; Cosh, M.H.; Cuenca, R.H.; Dorigo, W.A.; Draper, C.S.; Hagimoto, Y.; Kerr, Y.H.; Larson, K.M.; Njoku, E.G.; Small, E.E. State of the art in large-scale soil moisture monitoring. Soil Sci. Soc. Am. J. 2013, 77, 1888–1919. [Google Scholar] [CrossRef]
  37. Cao, D.; Shi, B.; Zhu, H.; Wei, G.; Chen, S.-E.; Yan, J. A distributed measurement method for in-situ soil moisture content by using carbon-fiber heated cable. J. Rock Mech. Geotech. Eng. 2015, 7, 700–707. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of soil dry depth.
Figure 1. Schematic diagram of soil dry depth.
Agronomy 13 01276 g001
Figure 2. Ultrasonic propagation properties in irrigated soil.
Figure 2. Ultrasonic propagation properties in irrigated soil.
Agronomy 13 01276 g002
Figure 3. The principle of ultrasonic sensor to test dry depth.
Figure 3. The principle of ultrasonic sensor to test dry depth.
Agronomy 13 01276 g003
Figure 4. The boundary condition imposed for modeling ultrasonic waves propagating in soil.
Figure 4. The boundary condition imposed for modeling ultrasonic waves propagating in soil.
Agronomy 13 01276 g004
Figure 5. The location of Guangzhou university for collecting soil samples.
Figure 5. The location of Guangzhou university for collecting soil samples.
Agronomy 13 01276 g005
Figure 6. Soil samples prepared for experiments: (a) clay, (b) sandy, (c) silty loam, (d) sandy loam.
Figure 6. Soil samples prepared for experiments: (a) clay, (b) sandy, (c) silty loam, (d) sandy loam.
Agronomy 13 01276 g006
Figure 7. The established soil dry depth measuring system.
Figure 7. The established soil dry depth measuring system.
Agronomy 13 01276 g007
Figure 8. The propagation velocity distribution of ultrasonic waves of different frequencies in soil: (ad) are the velocity distribution of ultrasonic wave of 40 Hz in the soils with the propagation time of 20 µs, 40 µs, 60 µs, and 80 µs, respectively. (eh) are the velocity distribution of ultrasonic wave of 60 Hz in the soils with the propagation time of 20 µs, 40 µs, 60 µs, and 80 µs, respectively. (il) are the velocity distribution of ultrasonic wave of 80 Hz in the soils with the propagation time of 20 µs, 40 µs, 60 µs, and 80 µs, respectively.
Figure 8. The propagation velocity distribution of ultrasonic waves of different frequencies in soil: (ad) are the velocity distribution of ultrasonic wave of 40 Hz in the soils with the propagation time of 20 µs, 40 µs, 60 µs, and 80 µs, respectively. (eh) are the velocity distribution of ultrasonic wave of 60 Hz in the soils with the propagation time of 20 µs, 40 µs, 60 µs, and 80 µs, respectively. (il) are the velocity distribution of ultrasonic wave of 80 Hz in the soils with the propagation time of 20 µs, 40 µs, 60 µs, and 80 µs, respectively.
Agronomy 13 01276 g008
Figure 9. The propagation pressure distribution of ultrasonic waves of different frequencies in soil: (ad) are the distribution of ultrasonic wave of 40 Hz in the soils with the propagation time of 20 µs, 40 µs, 60 µs, and 80 µs, respectively. (eh) are the distribution of ultrasonic wave of 60 Hz in the soils with the propagation time of 20 µs, 40 µs, 60 µs, and 80 µs, respectively. (il) are the distribution of ultrasonic wave of 80 Hz in the soils with the propagation time of 20 µs, 40 µs, 60 µs, and 80 µs, respectively.
Figure 9. The propagation pressure distribution of ultrasonic waves of different frequencies in soil: (ad) are the distribution of ultrasonic wave of 40 Hz in the soils with the propagation time of 20 µs, 40 µs, 60 µs, and 80 µs, respectively. (eh) are the distribution of ultrasonic wave of 60 Hz in the soils with the propagation time of 20 µs, 40 µs, 60 µs, and 80 µs, respectively. (il) are the distribution of ultrasonic wave of 80 Hz in the soils with the propagation time of 20 µs, 40 µs, 60 µs, and 80 µs, respectively.
Agronomy 13 01276 g009
Figure 10. The relationship between reflection time (μs) and dry depth (mm) for (a) clay, (b) sandy loam, (c) silty loam, and (d) sand, respectively.
Figure 10. The relationship between reflection time (μs) and dry depth (mm) for (a) clay, (b) sandy loam, (c) silty loam, and (d) sand, respectively.
Agronomy 13 01276 g010
Figure 11. The error analysis between ultrasonically and manually measured dry depth for (a) clay, (b) sandy loam, (c) silty loam and (d) sandy, respectively.
Figure 11. The error analysis between ultrasonically and manually measured dry depth for (a) clay, (b) sandy loam, (c) silty loam and (d) sandy, respectively.
Agronomy 13 01276 g011
Table 1. Material characteristic parameters of each surface layer of soil.
Table 1. Material characteristic parameters of each surface layer of soil.
Soil Moisture ContentSoil Density (kg/m3)Sound Velocity (m/s)
5%1.49520
25%1.67200
Table 2. The maximum measurement error for different soil samples.
Table 2. The maximum measurement error for different soil samples.
Soil TypesManually Measured Ultrasonic MeasuredThe Maximum Error
Clay96.23 mm90.78 mm5.67%
Sandy loam197.85 mm186.05 mm5.97%
Silty loam120.19 mm113.51 mm5.89%
Sandy62.54 mm58.8 mm5.98%
Table 3. The accuracy of the proposed method compared with the techniques previously reported.
Table 3. The accuracy of the proposed method compared with the techniques previously reported.
TechniquesR2RMSE
This workReflected ultrasonic wave0.98~0.991.5%~1.7%
Huisman et al. [34]TDR0.87~0.991.4%~4.4%
Robinson et al. [35]Capacitance sensors0.88~0.941.4%~2.4%
Ochsner et al. [36]Neutron probe0.74~0.992%~6.7%
Cao et al. [37]Carbon-fiber heated cable0.722.85%
Calamita et al. [24]Electrical resistivity0.21~0.681.15%~6.03%
Zhao et al. [23] L-band Radiometry Not mentioned4.1%
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liang, Z.; Zhao, C.; Zhang, Y.; Long, S.; Xiao, J.; Zhao, Z. Accurate and Rapid Measurement of Soil Dry Depth Using Ultrasonic Reflection Waves. Agronomy 2023, 13, 1276. https://doi.org/10.3390/agronomy13051276

AMA Style

Liang Z, Zhao C, Zhang Y, Long S, Xiao J, Zhao Z. Accurate and Rapid Measurement of Soil Dry Depth Using Ultrasonic Reflection Waves. Agronomy. 2023; 13(5):1276. https://doi.org/10.3390/agronomy13051276

Chicago/Turabian Style

Liang, Zhongwei, Chunhui Zhao, Yupeng Zhang, Sheng Long, Jinrui Xiao, and Zhuan Zhao. 2023. "Accurate and Rapid Measurement of Soil Dry Depth Using Ultrasonic Reflection Waves" Agronomy 13, no. 5: 1276. https://doi.org/10.3390/agronomy13051276

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop