Next Article in Journal
Polish Adaptation of the Yale Physical Activity Survey: Measurement Properties
Next Article in Special Issue
Seasonal Release Potential of Sediments in Reservoirs and its Impact on Water Quality Assessment
Previous Article in Journal
A Mosquito Workshop and Community Intervention: A Pilot Education Campaign to Identify Risk Factors Associated with Container Mosquitoes in San Pedro Sula, Honduras
Previous Article in Special Issue
Impacts of Artificial Underground Reservoir on Groundwater Environment in the Reservoir and Downstream Area
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Application of Dynamically Constrained Interpolation Methodology to the Surface Nitrogen Concentration in the Bohai Sea

1
Key Laboratory of Physical Oceanography, Qingdao Collaborative Innovation Center of Marine Science and Technology (CIMST), Ocean University of China, Qingdao 266100, China
2
Qingdao National Laboratory for Marine Science and Technology, Qingdao 266100, China
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2019, 16(13), 2400; https://doi.org/10.3390/ijerph16132400
Submission received: 30 April 2019 / Revised: 30 June 2019 / Accepted: 4 July 2019 / Published: 6 July 2019

Abstract

:
Observations of ocean pollutants are usually spatiotemporally dispersive, while it is of great importance to obtain continuous distribution of ocean pollutants in a certain area. In this paper, a dynamically constrained interpolated methodology (DCIM) is proposed to interpolate surface nitrogen concentration (SNC) in the Bohai Sea. The DCIM takes the pollutant transport advection diffusion equation as a dynamic constraint to interpolate SNCs and optimizes the interpolation results with adjoint method. Feasibility and validity of the DCIM are testified by ideal twin experiments. In ideal experiments, mean absolute gross errors between interpolated observations and final interpolated SNCs are all no more than 0.03 mg/L, demonstrating that the DCIM can provide convincing results. In practical experiment, SNCs are interpolated and the final interpolated surface nitrogen distribution is acquired. Correlation coefficient between interpolated and observed SNCs is 0.77. In addition, distribution of the final interpolated SNCs shows a good agreement with the observed ones.

1. Introduction

The Bohai Sea is the largest and the only semiclosed inland sea in China, surrounded by land on its three sides. It connects to the north of the Yellow Sea through the Bohai Strait. The weak water exchange causes a poor self-purification ability of the Bohai Sea, making it difficult to be restored in a short time if the marine ecosystem is severely damaged. According to statistics, the sewage water of over 40 rivers flows into the Bohai Sea and the volume is nearly 890 × 108 m3 [1], which inevitably aggravates the environmental problems, such as ocean eutrophication (the accumulation of nutrients: nitrogen, phosphorus, etc.). Moreover, the severe deterioration in marine environment has badly affected the development of fishery and the Bohai Sea is gradually losing its function as a fishing ground [2]. To maintain sustainable development, relevant researches about marine pollutants have been conducted. The mathematical models are considered as the most direct and effective way for quantification [3], and with help of a mathematical model knowing more about the temporal and spatial distributions of pollutants in the Bohai Sea plays an important role in environment restoration.
Activities in coastal oceans can help to speed up economic construction, but meantime it will cause serious pollution to marine ecosystems. A number of numerical studies have been carried out to simulate pollutant dispersion [4,5,6,7]. A two-dimensional water quality model was developed and applied to analyze and optimize the ecological programs, and it can simulate key model variables (NH4+-N, PO43−-P, chemical oxygen demand, and water level) [8]. Lee et al. [9] established an advection-dispersion model for pollutant transport simulation to analyze the influence of tidal currents on the concentration distribution. Gupta et al. [10] utilized numerical modeling to determine the sewage assimilative ability and found that the water quality was badly deteriorated due to the multiple sewage discharge. Periáñez [11] developed a particle-tracking model constituted by an off-line running hydrodynamic module to simulate the dispersion of pollutants. A three-dimensional numerical model of gravity flows was introduced in the research of Huang et al. [12], which was used to investigate the distribution features of various pollutants discharged at different positions in a wide river. Li et al. [13] simulated the temporal and spatial distribution of pollutants of the Bohai Sea in twin experiments with the adjoint assimilation method.
Interpolation methods, such as the Kriging, Cressman, spline, and polynomial interpolations, are widely used in the numerical model to obtain an integrated field based on sparse observations. Bargaoui and Chebbi [14] applied Kriging methods to evaluate the spatial and temporal variability of rainfall. Jeffrey et al. [15] adopted a thin plate smoothing spline and the ordinary Kriging to get daily climate variables and rainfall, respectively. A modified Cressman method was proposed in the study of Liu et al. [16], where the influence radius was modified to produce relatively accurate distributions. Wang et al. [17] applied the Cressman interpolation method to calculate the monthly mean distribution of total nitrogen to study the initial filed of pollution in the Bohai Sea. Guo et al. [18] introduced the surface spline interpolation to a two-dimensional tidal model and illustrated the feasibility and practicability of the method.
The adjoint assimilation method has been applied in oceanography for decades [19,20,21,22]. Zhang et al. [23] used the adjoint method in a two-dimensional tidal model to study the characteristics of bottom friction parameterizations. A three-dimensional cohesive sediment transport model with the adjoint assimilation method was established in Wang et al. [24] to get better simulation results of parameters. Furthermore, Mao et al. [25] developed the dynamically constrained interpolation methodology (DCIM), where the dynamic constraints were combined with the statistical information of observations to interpolate the suspended sediment concentrations. In this paper, DCIM will be applied to interpolate the surface nitrogen concentration in the Bohai Sea. The rest of the paper is organized as follows. Section 2 describes the dynamic constraint model, observation information, and details of the DCIM. Section 3 gives the results of numerical experiments. Section 4 concludes the whole work.

2. Materials and Methods

2.1. The Dynamical Model

Considering the convection and diffusion processes, the governing equation of marine pollutant transport model is presented as follows [17,26,27]
C t + u C x + v C y + w C z = x ( A H C x ) + y ( A H C y ) + z ( K H C z ) r C
where C denotes the concentration of pollutants; t and x, y, z are the symbols of time and space, respectively; u and v represent the horizontal velocities (in x and y directions, respectively) and w represents the vertical velocity (in z direction); AH and KH denote the horizontal and vertical diffusion coefficients (AH = 100 m2/s, KH = 0.00001 m2/s), respectively; r is the pollutant attenuation coefficient, and r = 0, which means that the pollutant is treated as conservative substance [17]. For the finite difference scheme readers can be referred to the Appendix A.
The open boundary of the model is set at 122.5° E, where a no-gradient condition and constant condition are used at the outflow boundary and the inflow boundary, respectively.

2.2. Observations and Model Setting

The marine environmental monitoring data in the Bohai Sea and the north Yellow Sea, provided by the North China Sea Environmental Monitoring Center, State Oceanic Administration, includes the data of February, May, August and October of each year. Nitrate, phosphate, pH, etc. are monitored in order to investigate the spatiotemporal distribution of different pollutant elements and then diagnose marine pollution matter [16]. The distribution of observation points is shown in Figure 1 and the date of observations is given in Table 1.
The monitoring data in 2009 are analyzed in practical experiments in this paper. The computational domain is the Bohai Sea (37° N–41° N, 117.5° E–122.5° E) with a 4′ × 4′ grid resolution. The computing time is 30 days and the time step is set to be 6 h. The three-dimensional Regional Ocean Model System (ROMS) provides the hydrodynamic flow field used in numerical experiments of present study [27].

2.3. Dynamically Constrained Interpolation Methodology

According to the study of Yaremchuk and Sentchev [28], the dynamically unconstrained interpolation method and the dynamically constrained interpolation method (DCIM) are two parts of interpolation methods, and the DCIM is used in the model to obtain the interpolation of observations. In addition, the adjoint method is used to optimize the interpolation results.

2.3.1. The Adjoint Methods

To optimize the interpolation results, the misfit between interpolation results and observations should be gradually reduced, which is described by the cost function and defined as [17]
J = 1 2 K C ( C i , j , k C ¯ i , j , k ) 2
where Ci,j,k and C ¯ i , j , k denote the interpolation results and the observation data at the point (i,j,k), respectively; KC represents the weighting matrix whose element equals to 1 when the observations are available; otherwise, KC = 0.
The governing equation of marine pollutant transport model (1) can be written as
F = C t + u C x + v C y + w C z x ( A H C x ) y ( A H C y ) z ( K H C z ) + r C
Based on the Lagrange multiple method, the Lagrange function can be written as
L = J + Ω ( C * F ) d Ω
where C* represents the adjoint variable of C; Ω denotes the computational domain.
The adjoint model of the pollution transport model is calculated from Equation (5). The gradients of the cost function with respect to model parameters can be calculated by Equation (6):
L C = 0
L p = 0
where p stands for the model parameters.
In this paper, the adjoint equation and the gradient can be written as Equations (7) and (8), respectively.
C * t z ( K H C * z ) = u C * x + v C * y + w C * z + x ( A H C * x ) + y ( A H C * y ) - K C ( C C ¯ )
J C 1 = ( C * t ) 1 + ( u C * x ) 1 + ( v C * y ) 1 + ( w C * z ) 1 + x ( A H C * x ) 1 + y ( A H C * y ) 1 + z ( K H C * z ) 1
where the superscript 1 denotes the SNC at the first iteration step.

2.3.2. The Process of DCIM

The DCIM contains the following steps, described as [25] follows.
Step 1. Propose a guess value of the parameters in the model.
Step 2. Acquire the interpolation of observations through forward model.
Step 3. Calculate the cost function and obtain the Lagrange multiple through adjoint model.
Step 4. Based on Equation (6), acquire the gradients of the cost function with respect to the parameters of the model and adjust the parameters along the opposite direction of the gradient.
Step 5. Stop calculating when the preset ending condition is satisfied; otherwise, go to step 2 and continue iterating.

3. Numerical Results

3.1. Verification of the DCIM

In this part, we testified the feasibility and validity of the DCIM by ideal experiments. The observations used in ideal experiments were generated by integrating the given distribution of nitrogen over time. In order to maintain the universality, the initial guess values were set to be half of the max value of nitrogen concentration. Similar conclusions were drawn when other initial guess values were taken. Statistic results of other initial guess values are given in Appendix B.

3.1.1. Application of the DCIM in Ideal Twin Experiments

As mentioned by Elbern et al. [29], the validity of the assimilated or interpolated results can only be testified by the observations that were not assimilated or interpolated. Therefore, one-fifth of the total observations were randomly selected as observations that were not interpolated but only used for verification, and these observations were named as checking observations. The other observations were named as interpolated observations, which were to be interpolated with the DCIM. By this cross-validation, it can be distinguished that whether the interpolated observations were overfitted or not. If the interpolated observations were overfitted, there would be large misfit between simulation results and checking observations [30].
To eliminate the contingency induced by selection of checking observations, all idealized observations were randomly divided into five subsets and every subset was taken as the checking observations by turns. Therefore, there were five twin experiments, which were named as IE_11–IE_15, respectively. The statistics of these twin experiments are listed in Table 2. The Cressman interpolation method [31] was introduced to the ideal twin experiments IE_11-IE_15 so that the quality of results can be assessed. The comparison is presented in Appendix C. In order to quantify the difference between interpolated SNCs and observations, mean absolute gross error (MAGE) and mean normalized gross error (MNGE) were calculated as follows
MAGE = 1 N i = 1 N | I i O i |
MNGE = 1 N i = 1 N [ ( | I i O i | ) / O i ]
where N is the number of observations and I and O are the interpolated SNCs and observations, respectively.
In the five twin experiments, the rate of decline in MAGEs between checking observations and corresponding interpolated SNCs (K3) were 66.7%, 70.8%, 63.6%, 57.1%, and 71.4%, respectively; and the errors were no more than 0.12 mg/L. What is more, the MNGEs between checking observations and interpolated SNCs (K4) were all reduced by at least 50%. Besides, at the first iteration step, the MNGEs between interpolated observations and corresponding interpolated SNCs (K2) were all larger than 120%, while after applying the DCIM, K2 were all less than 6%. Thus, it can be demonstrated that the interpolated observations were not overfitted. Figure 2 shows that most dots were near the 1:1 line, no matter whether the dot stands for interpolated observations or checking observations, which indicates that the DCIM was an effective tool to interpolate observations.
To make the fullest use of all observations, another twin experiment IE_21 was conducted. In IE_21 all observations were used to interpolate and the final MAGE was 0.02 mg/L, which was reduced by 92.9%, while the final MNGE was 6.06% (see Table 2). Comparison between interpolated SNCs and prescribed observations is shown in Figure 2f. The correlation coefficient was near 1.00 on the whole, meaning that the final interpolated results were almost equal to the artificial observations. Thus, we can say that the DCIM was a feasible and effective method to interpolate the SNCs.

3.1.2. Sensitivity to Observational Errors

In the real ocean environments, the observations can be contaminated by noises. Therefore, another three twin experiments, named by IE_31, IE_32, and IE_33, respectively were conducted, in which random perturbations were added to the prescribed observations. The maximum percentages of observation errors were 10%, 20%, and 30% in three experiments, respectively. The comparison between interpolated SNCs and observations were shown in Figure 2g–i, and the results indicated that the final interpolated SNCs were close to the observations in all three twin experiments. Moreover, the statistics of MAGEs and MNGEs shown in Table 2 also demonstrated that although the observations contained noises, the DCIM can still perform well when used to interpolate the SNCs. This means that the interpolation results may still be convincing when the DCIM was adopted in the practical situation.

3.2. Practical Applications

In this section, the observed data of SNCs were used to carry out practical experiments. The final MAGE and MNGE were 0.21 mg/L and 47.9%, respectively (Table 3), which were both reduced by more than 55%. The mean value and the standard deviation of the observed SNCs were 0.69 and 0.55 mg/L, respectively, while those of the interpolated SNCs were 0.69 and 0.46 mg/L, respectively. The results indicated that the interpolated SNCs were almost equal to the observed SNCs. Figure 3 showed the scatterplot to compare interpolated SNCs and observed SNCs visually. The 2:1, 1.25:1, 1:1, 0.85:1, and 1:2 lines were shown for reference. For 84.3% of the observations, the ratio of interpolated SNCs to the observed was between 0.5 and 2; for 11.6%, the ratio was over 2 and for 4.1%, the ratio was below 0.5. It was obvious that the closer the ratio was to 1, the close the interpolated SNCs were to the observed SNCs. For 53.7% of the observations, the ratio was between 0.85 and 1.25. What is more, the correlation coefficient between the interpolated SNCs and the observed SNCs was 0.77.
The statistical results mentioned above indicated that the interpolated SNCs with DCIM were coherent with the observed SNCs. The final distribution of the interpolated surface nitrogen concentration was given in Figure 4. The MAGE between each interpolated observation and interpolated SNC was shown in Figure 5. Statistics of MAGEs was shown in Figure 6. By statistics, we can know that 46.3% (56/121) of the MAGEs were no more than 0.1 mg/L and only 20.7% (25/121) of the MAGEs were over 0.3 mg/L. Figure 4 showed that high concentration appears in the three bays, while in the central Bohai Sea the concentration was low, and comparing with Figure 1 it showed a good agreement with the observed nitrogen concentration distribution.

4. Conclusions

In this paper, we interpolated the surface nitrogen concentration with the dynamically constrained interpolation methodology (DCIM). The pollutant transport model was taken as dynamic constraint and the interpolated results were optimized iteratively with the adjoint method.
The feasibility and validity of DCIM were testified with prescribed observations in ideal twin experiments. The statistics and the scatterplot of twin experiments illustrated that the interpolated SNCs with DCIM were close to the prescribed observations and that the interpolated results were still convincing when noises were added to the prescribed observations. In practical experiment, the observed data were used to interpolate the surface nitrogen concentration with DCIM. The correlation coefficient between interpolated SNCs and observed SNCs was 0.77. The distribution of final interpolated surface nitrogen concentration shows a good agreement with the observations. The interpolated results in ideal experiment and in practical experiment demonstrated that the DCIM can be an effective method to interpolate the spatial and temporal distributing observations.

Author Contributions

Conceptualization, X.L. (Xianqing Lv); Formal analysis, X.L. (Xiaona Li); Funding acquisition, X.L. (Xianqing Lv); Investigation, Q.Z.; Methodology, Q.Z.; Software, Q.Z. and X.L. (Xiaona Li); Supervision, X.L. (Xianqing Lv); Writing—original draft, Q.Z.; Writing—review & editing, X.L. (Xiaona Li).

Funding

This work is supported by the National Natural Science Foundation of China (Grant No. U1806214) and the National Key Research and Development Plan of China (Grant No. 2016YFC1402705 and 2016YFC1402304).

Acknowledgments

We are grateful to the reviewers and editors whose comments helped to improve this paper. We would like to thank Haibo Chen, Daosheng Wang, Haidong Pan, and Yafei Nie for their valuable discussions and suggestions.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Difference Form of Equation (1) and Equation (5)

For Equation (1), the difference form was as follows
C i , j , k l + 1 C i , j , k l Δ t [ K H ( C i , j , k + 1 l + 1 C i , j , k l + 1 ) Δ z k + 1 / 2 Δ z k + 1 K H ( C i , j , k l + 1 C i , j , k 1 l + 1 ) Δ z k + 1 / 2 Δ z k ] = u i , j , k l ( C i + 1 , j , k l C i 1 , j , k l ) 2 Δ x j v i , j , k l ( C i , j + 1 , k l C i , j 1 , k l ) 2 Δ y w i , j , k l ( C i , j , k + 1 l C i , j , k 1 l ) 2 Δ z k + [ A H ( C i + 1 , j , k l C i , j , k l ) Δ x j + 1 / 2 Δ x j + 1 A H ( C i , j , k l C i 1 , j , k l ) Δ x j + 1 / 2 Δ x j ] + [ A H ( C i , j + 1 , k l C i , j , k l ) ( Δ y ) 2 A H ( C i , j , k l C i , j 1 , k l ) ( Δ y ) 2 ]
For Equation (5), the difference form can be described as
C i , j , k * l 1 C i , j , k * l Δ t [ K H ( C i , j , k + 1 * l 1 C i , j , k * l 1 ) Δ z k + 1 / 2 Δ z k + 1 K H ( C i , j , k * l 1 C i , j , k 1 * l 1 ) Δ z k + 1 / 2 Δ z k ] = ( u i + 1 , j , k l C i + 1 , j , k * l u i 1 , j , k l C i 1 , j , k } * l ) 2 Δ x j + ( v i , j + 1 , k l C i , j + 1 , k * l v i , j 1 , k l C i , j 1 , k * l ) 2 Δ y + ( w i , j , k + 1 l C i , j , k + 1 * l w i , j , k 1 l C i , j , k 1 * l ) 2 Δ z k + [ A H ( C i + 1 , j , k * l C i , j , k * l ) Δ x j + 1 / 2 Δ x j + 1 A H ( C i , j , k * l C i 1 , j , k * l ) Δ x j + 1 / 2 Δ x j ] + [ A H ( C i , j + 1 , k * l C i , j , k * l ) ( Δ y ) 2 A H ( C i , j , k * l C i , j 1 , k * l ) ( Δ y ) 2 ] K C ( C i , j , k l C i , j , k l ¯ )

Appendix B. Statistic Results of Other Initial Guess Values

During calculation, there are three kinds of initial guess values, including the minimum, half and maximum of the nitrogen concentration. The simulation results were given in Table A1.
Table A1. Simulation results of different initial guess values.
Table A1. Simulation results of different initial guess values.
Initial GuessK1K2
InitialFinalInitial (%)Final (%)
Minimum0.370.0285.395.48
Half0.280.02132.476.06
Maximum0.900.02344.038.58
K1 is MAGEs between the interpolated observations and the interpolated SNCs (mg/L); K2 is MNGEs between the interpolated observations and the interpolated SNCs (mg/L).
From Table A1 we know that after iterative optimization the final MAGE of the three initial guess values were all 0.02 mg/L. It meant that no matter what the initial guess value was, the final simulation results were almost the same. The reason why the final percentages of K2 were different was that due to the different initial guess values, the initial errors were different. So in the paper, half of the max value of nitrogen concentration was taken as the initial guess value.

Appendix C. Comparison between the DCIM and Cressman Interpolation Method

In order to access the quality of results, the Cressman interpolation (CI) method was introduced to the ideal twin experiments IE_11-IE_15. Comparison of MAGEs between CI and DCIM was shown in Table A2.
Table A2. Comparison of MAGEs between the two methods (unit: mg/L).
Table A2. Comparison of MAGEs between the two methods (unit: mg/L).
IE_11IE_12IE_13IE_14IE_15
DCIM0.110.070.120.090.08
CI0.230.220.290.280.35
In experiments, from IE_11 to IE_15, the MAGEs were reduced by 52.2%, 68.1%, 58.6%, 67.9%, and 77.1%, respectively, and errors were reduced by almost an order of magnitude. Through comparison, we can draw the conclusion that the DCIM was a much more effective interpolation method than the CI.

References

  1. Sündermann, J.; Feng, S. Analysis and modelling of the Bohai sea ecosystem—A joint German-Chinese study. J. Mar. Syst. 2004, 44, 127–140. [Google Scholar] [CrossRef]
  2. Gao, X.; Zhou, F.; Chen, C.T.A. Pollution status of the Bohai Sea: An overview of the environmental quality assessment related trace metals. Environ. Int. 2014, 62, 12–30. [Google Scholar] [CrossRef] [PubMed]
  3. Fanghua, H.; Shengtian, Y.; Hongguang, C.; Qingsong, B.; Lingfang, Z. A method for estimation of non-point source pollution load in the large-scale basins of China. Acta Sci. Circumst. 2006, 26, 375–383. [Google Scholar]
  4. Duan, J.G.; Nanda, S.K. Two-dimensional depth-averaged model simulation of suspended sediment concentration distribution in a groyne field. J. Hydrol. 2006, 327, 426–437. [Google Scholar] [CrossRef]
  5. Sámano, M.L.; Pérez, M.L.; Claramunt, I.; García, A. Assessment of the zinc diffusion rate in estuarine zones. Mar. Pollut. Bull. 2016, 104, 121–128. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, J.; Kitazawa, D. Numerical analysis of particulate organic waste diffusion in an aquaculture area of Gokasho Bay, Japan. Mar. Pollut. Bull. 2015, 93, 130–143. [Google Scholar] [CrossRef] [PubMed]
  7. Xu, M.; Chua, V.P. A numerical study on land-based pollutant transport in Singapore coastal waters with a coupled hydrologic-hydrodynamic model. J. Hydro. Environ. Res. 2017, 14, 119–142. [Google Scholar] [CrossRef]
  8. Chen, Q.; Tan, K.; Zhu, C.; Li, R. Development and application of a two-dimensional water quality model for the Daqinghe River Mouth of the Dianchi Lake. J. Environ. Sci. 2009, 21, 313–318. [Google Scholar] [CrossRef]
  9. Lee, M.E.; Seo, I.W. Analysis of pollutant transport in the Han River with tidal current using a 2D finite element model. J. Hydro. Environ. Res. 2007, 1, 30–42. [Google Scholar] [CrossRef]
  10. Gupta, I.; Dhage, S.; Chandorkar, A.A.; Srivastav, A. Numerical modeling for Thane creek. Environ. Model. Softw. 2004, 19, 571–579. [Google Scholar] [CrossRef]
  11. Periáñez, R. GISPART: A numerical model to simulate the dispersion of contaminants in the Strait of Gibraltar. Environ. Model. Softw. 2005, 20, 797–802. [Google Scholar] [CrossRef]
  12. Huang, H.; Chen, G.; Zhang, Q.F. The distribution characteristics of pollutants released at different cross-sectional positions of a river. Environ. Pollut. 2010, 158, 1327–1333. [Google Scholar] [CrossRef] [PubMed]
  13. Li, X.; Xu, M.; Lv, X.; Fu, K. A Study of the Transport of Marine Pollutants Using Adjoint Method of Data Assimilation with Method of Characteristics. Adv. Math. Phys. 2018, 2018, 14. [Google Scholar] [CrossRef]
  14. Kebaili Bargaoui, Z.; Chebbi, A. Comparison of two kriging interpolation methods applied to spatiotemporal rainfall. J. Hydrol. 2009, 365, 56–73. [Google Scholar] [CrossRef]
  15. Jeffrey, S.J.; Carter, J.O.; Moodie, K.B.; Beswick, A.R. Using spatial interpolation to construct a comprehensive archive of Australian climate data. Environ. Model. Softw. 2001, 16, 309–330. [Google Scholar] [CrossRef]
  16. Liu, Y.; Yu, J.; Shen, Y.; Lv, X. A modified interpolation method for surface total nitrogen in the Bohai Sea. J. Atmos. Ocean. Technol. 2016, 33, 1509–1517. [Google Scholar] [CrossRef]
  17. Wang, C.; Li, X.; Lv, X. Numerical study on initial field of pollution in the Bohai Sea with an adjoint method. Math. Probl. Eng. 2013, 2013, 1–10. [Google Scholar] [CrossRef]
  18. Guo, Z.; Pan, H.; Fan, W.; Lv, X. Application of surface spline interpolation in inversion of bottom friction coefficients. J. Atmos. Ocean. Technol. 2017, 34, 2021–2028. [Google Scholar] [CrossRef]
  19. Peng, S.Q.; Xie, L. Effect of determining initial conditions by four-dimensional variational data assimilation on storm surge forecasting. Ocean Model. 2006, 14, 1–18. [Google Scholar] [CrossRef] [Green Version]
  20. Li, X.; Wang, C.; Fan, W.; Lv, X. Optimization of the spatiotemporal parameters in a dynamical marine ecosystem model based on the adjoint assimilation. Math. Probl. Eng. 2013, 2013, 1–12. [Google Scholar] [CrossRef]
  21. Chen, H.; Cao, A.; Zhang, J.; Miao, C.; Lv, X. Estimation of spatially varying open boundary conditions for a numerical internal tidal model with adjoint method. Math. Comput. Simul. 2014, 97, 14–38. [Google Scholar] [CrossRef]
  22. Pan, H.; Guo, Z.; Lv, X. Inversion of tidal open boundary conditions of the M2constituent in the bohai and yellow seas. J. Atmos. Ocean. Technol. 2017, 34, 1661–1672. [Google Scholar] [CrossRef]
  23. Zhang, J.; Lu, X.; Wang, P.; Wang, Y.P. Study on linear and nonlinear bottom friction parameterizations for regional tidal models using data assimilation. Cont. Shelf Res. 2011, 31, 555–573. [Google Scholar] [CrossRef]
  24. Wang, D.; Cao, A.; Zhang, J.; Fan, D.; Liu, Y.; Zhang, Y. A three-dimensional cohesive sediment transport model with data assimilation: Model development, sensitivity analysis and parameter estimation. Estuar. Coast. Shelf Sci. 2018, 206, 87–100. [Google Scholar] [CrossRef]
  25. Mao, X.; Wang, D.; Zhang, J.; Bian, C.; Lv, X. Dynamically constrained interpolation of the sparsely observed suspended sediment concentrations in both space and time: A case study in the Bohai Sea. J. Atmos. Ocean. Technol. 2018, 35, 1151–1167. [Google Scholar] [CrossRef]
  26. Zong, X.; Pan, H.; Liu, Y.; Lv, X. Improved estimation of pollutant emission rate in an ocean pollutant diffusion model by the application of spline interpolation with the adjoint method. J. Atmos. Ocean. Technol. 2018, 35, 1961–1975. [Google Scholar] [CrossRef]
  27. Zong, X.; Xu, M.; Xu, J.; Lv, X. Improvement of the ocean pollutant transport model by using the surface spline interpolation. Tellus A Dyn. Meteorol. Oceanogr. 2018, 70, 1–13. [Google Scholar] [CrossRef] [Green Version]
  28. Yaremchuk, M.; Sentchev, A. Interpolation of the Radial Velocity Data From Coastal Hf Radars. Radar Syst. 2005, 19, 298–301. [Google Scholar]
  29. Elbern, H.; Strunk, A.; Schmidt, H.; Talagrand, O. Emission rate and chemical state estimation by 4-dimensional variational inversion. Atmos. Chem. Phys. 2007, 7, 3749–3769. [Google Scholar] [CrossRef] [Green Version]
  30. Mattern, J.P.; Fennel, K.; Dowd, M. Estimating time-dependent parameters for a biological ocean model using an emulator approach. J. Mar. Syst. 2012, 96, 32–47. [Google Scholar] [CrossRef]
  31. Cressman, G.P. An Operational Objective Analysis System. Mon. Weather Rev. 1959, 87, 367–374. [Google Scholar] [CrossRef]
Figure 1. Topography of the Bohai Sea (depth in meters) and distribution of observation points in May 2009. The size of each point indicates surface nitrogen concentration.
Figure 1. Topography of the Bohai Sea (depth in meters) and distribution of observation points in May 2009. The size of each point indicates surface nitrogen concentration.
Ijerph 16 02400 g001
Figure 2. Comparison of simulated and observed surface nitrogen concentrations (SNCs), including interpolated observations (red dots) and checked observations (blue dots), for (a) IE_11, (b) IE_12, (c) IE_13, (d) IE_14, (e) IE_15, (f) IE_21, (g) IE_31, (h) IE_32, and (i) IE_33. The 1:1 lines are shown for reference in all the panels.
Figure 2. Comparison of simulated and observed surface nitrogen concentrations (SNCs), including interpolated observations (red dots) and checked observations (blue dots), for (a) IE_11, (b) IE_12, (c) IE_13, (d) IE_14, (e) IE_15, (f) IE_21, (g) IE_31, (h) IE_32, and (i) IE_33. The 1:1 lines are shown for reference in all the panels.
Ijerph 16 02400 g002
Figure 3. Comparison of interpolated and observed SNCs. The 2:1, 1.85:1, 1:1, 0.85:1, and 1:2 lines are shown for reference.
Figure 3. Comparison of interpolated and observed SNCs. The 2:1, 1.85:1, 1:1, 0.85:1, and 1:2 lines are shown for reference.
Ijerph 16 02400 g003
Figure 4. The final distribution of the interpolated surface nitrogen concentration (unit: mg/L).
Figure 4. The final distribution of the interpolated surface nitrogen concentration (unit: mg/L).
Ijerph 16 02400 g004
Figure 5. Mean absolute gross error (MAGE) between each interpolated observation and interpolated SNC (unit: mg/L).
Figure 5. Mean absolute gross error (MAGE) between each interpolated observation and interpolated SNC (unit: mg/L).
Ijerph 16 02400 g005
Figure 6. Statistics of MAGEs.
Figure 6. Statistics of MAGEs.
Ijerph 16 02400 g006
Table 1. Observation information: location and date.
Table 1. Observation information: location and date.
Longitude (°N)Latitude (°E)DateLongitude (°N)Latitude (°E)Date
119.705139.9282119.788137.786413
119.37539.67785121.594440.776413
120.927837.82926119.533338.22514
120.631937.88196119.161138.186114
119.470839.75427120.341737.7514
120.423640.10837119.548637.566714
120.088940.08337119.42537.770814
118.333338.27120.263937.795814
120.344440.15427120.286137.645814
118.129839.13077121.330640.679214
118.634739.07688121.541740.559714
120.806940.6083811839.133314
121.084740.58198117.838939.111114
118.787438.96529118.959738.311115
120.494440.227810118.945338.621715
120.620840.302810118.548338.348915
122.140340.544410118.5538.562515
122.056940.654210117.808339.029215
120.797240.486111118.006939.011115
119.362539.533311117.804238.951415
118.286138.96411122.081940.241716
118.430338.947211118.538.840316
121.798640.658311118.316738.708316
121.804240.779211117.616738.762516
121.040340.719412121.77540.17516
121.108340.813912118.802838.152818
121.241740.852812118.900838.148318
121.405640.8512122.131940.338919
119.996937.999413121.911140.468119
120.722238.133313122.187540.429219
120.32538.366713118.013938.518120
119.913937.605613119.488938.847221
119.079239.098621121.147238.718124
119.540339.808921121.283339.018124
118.883338.906921121.091738.818124
121.616740.011121119.258337.190324
121.752839.934721121.220840.244525
118.183338.221120.072239.066726
121.906940.094421120.883338.919426
118.219438.388921121.08239.130526
119.495637.153121117.986138.659726
117.683338.501421117.997238.836126
119.397837.411121117.633338.627826
117.822238.344421118.611738.144426
118.013938.270821119.040337.379227
119.604237.280621119.082537.531927
119.326439.663922119.016737.520827
119.466739.31822119.215337.898627
121.245839.629222119.188937.536127
121.431939.820822120.966737.991728
121.258339.483322120.302838.679228
119.172639.312323120.716738.436128
118.980539.158323121.030638.2528
121.255639.311123120.938.616728
121.555639.294423121.097238.508328
121.477839.184723119.011138.022228
121.633339.093123120.122237.477830
120.645839.68224119.861137.248630
120.245839.92524119.848137.386130
119.8539.8524119.766737.179230
119.312539.413924
Table 2. Statistics of the ideal experiments.
Table 2. Statistics of the ideal experiments.
K1K2K3K4
ExptInitialFinalInitial (%)Final (%)InitialFinalInitial (%)Final (%)
IE_110.260.01124.965.290.330.11157.2077.67
IE_120.280.01135.454.000.240.07118.4428.00
IE_130.270.01134.514.910.330.12123.4747.18
IE_140.300.01131.713.740.210.09135.6350.93
IE_150.270.01135.355.130.280.08121.8937.02
IE_210.280.02132.476.06
IE_310.280.02134.468.20
IE_320.280.02134.277.68
IE_330.290.03140.1810.62
K1 is MAGEs between the interpolated observations and the interpolated SNCs (mg/L); K2 is MNGEs between the interpolated observations and the interpolated SNCs; K3 is MAGEs between the checking observations and the interpolated SNCs (mg/L); K4 is MNGEs between the checking observations and the interpolated SNCs.
Table 3. Statistics of practical experiment.
Table 3. Statistics of practical experiment.
K1K2
InitialFinalInitial (%)Final (%)
0.480.21120.7447.90
K1 is MAGEs between the interpolated observations and the interpolated SNCs (mg/L); K2 is MNGEs between the interpolated observations and the interpolated SNCs.

Share and Cite

MDPI and ACS Style

Zheng, Q.; Li, X.; Lv, X. Application of Dynamically Constrained Interpolation Methodology to the Surface Nitrogen Concentration in the Bohai Sea. Int. J. Environ. Res. Public Health 2019, 16, 2400. https://doi.org/10.3390/ijerph16132400

AMA Style

Zheng Q, Li X, Lv X. Application of Dynamically Constrained Interpolation Methodology to the Surface Nitrogen Concentration in the Bohai Sea. International Journal of Environmental Research and Public Health. 2019; 16(13):2400. https://doi.org/10.3390/ijerph16132400

Chicago/Turabian Style

Zheng, Quanxin, Xiaona Li, and Xianqing Lv. 2019. "Application of Dynamically Constrained Interpolation Methodology to the Surface Nitrogen Concentration in the Bohai Sea" International Journal of Environmental Research and Public Health 16, no. 13: 2400. https://doi.org/10.3390/ijerph16132400

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