Next Article in Journal
Changes in the Occurrence of Late Spring Frost in Poland
Previous Article in Journal
Leaf Response to Different Light Spectrum Compositions during Micropropagation of Gerbera Axillary Shoots
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Distribution of TiO2 Nanoparticles in Acidic and Alkaline Soil and Their Accumulation by Aspergillus niger

1
Institute of Laboratory Research on Geomaterials, Faculty of Natural Sciences, Comenius University in Bratislava, Ilkovičova 6, 842 15 Bratislava, Slovakia
2
Department of Soil Science and Geology, Faculty of Agrobiology and Food Resources, Slovak University of Agriculture in Nitra, Trieda A. Hlinku 2, 949 76 Nitra, Slovakia
3
Nanotechnology Centre, VŠB Technical University of Ostrava, 17. listopadu 15/2172, 708 33 Ostrava, Czech Republic
*
Author to whom correspondence should be addressed.
Agronomy 2020, 10(11), 1833; https://doi.org/10.3390/agronomy10111833
Submission received: 29 September 2020 / Revised: 15 November 2020 / Accepted: 19 November 2020 / Published: 22 November 2020
(This article belongs to the Section Soil and Plant Nutrition)

Abstract

:
The nanoparticles of TiO2 (TiO2 NP) have been used as a plant-growth stimulant or catalyst in pesticide formulas. However, due to high resistance of TiO2 NP to abiotic weathering, dissolved Ti is unlikely to act as an active compound in these preparations. Even if soil is acidic, TiO2 NP do not dissolve easily and preferably remain as undissolved particles. The low dissolution rates of inorganic nanoparticles in the soil environment make Ti in TiO2 NP largely unavailable for plants and soil microorganisms. To characterize the behavior of TiO2 NP in soil under different pH conditions, we analyzed TiO2 NP-size distribution in two soil materials, an alkaline and acidic one. We also cultivated Aspergillus niger, a fungus ubiquitously found in soils, in the growth medium spiked with TiO2 NP to assess accumulation of the nanoparticles in fungus. In soil suspensions, the dissolved Ti was present in low concentrations (up to 0.010 mg L−1). Most of the TiO2 NP remained in particulate form or appeared as aggregates sized 100–450 nm. In experiment on Ti accumulation by A. niger, TiO2 NP either settled down to the bottom of the flask with growth medium or were actually accumulated by the fungus; about 7.5% of TiO2 NP were accumulated in fungal mycelia. Most of the TiO2 NP remain in particulate form in soil solutions, regardless of soil pH. Filamentous fungus A. niger has the ability to accumulate bioavailable TiO2 NP, which hints at the possibility that some soil fungi can affect spatial distribution of this type of nanoparticles in soils.

1. Introduction

Titanium (Ti) is one of the most abundant elements in soil, sediments and weathered rocks. In natural waters, however, the levels of soluble Ti range from 4 pM at the surface of ocean waters [1] to 100 µM in hot spring waters [2]. This is due to a very low dissolution rate and high resistance of Ti minerals to chemical weathering [3]. Its essentiality is also disputed [4,5]. It has just recently been found that Ti has a role in the growth and development of plants [4] and it acts as a plant-growth stimulant when in nanoparticulate form (TiO2 NP) [6,7,8,9,10,11,12]. The TiO2 NP are used as additives in pesticides and other complex formulations as either an active ingredient or a carrier. They increase adhesion or accelerate the degradation of the organic pesticides [13,14,15,16,17]. Yet, they can negatively affect roots and symbiotic mycorrhizae of plants and fungi [18]. The recommended dosage varies with application and plant species. Generally, a single application of TiO2 NP that does not increase nanoparticle concentration above the value of approx. 60 mg kg−1 is considered to be safe for use [19,20].
The stability of nanoparticles in soils is strongly affected by pH. Their dissolution rate mostly increases with increasing soil acidity [21,22]. During the process of breathing, plants release CO2 into the soil where it dissolves and forms carbonic acid. That acid reduces the pH of soil in the root area [23]. Plants also release organic acids into the environment to better mineralize and solubilize poorly available elements [24]. Soil microorganisms also partially contribute to this process [25] via complexolysis by excreting extracellular biogenic chelating metabolites. Moreover, a larger area of nanoparticles’ surfaces is exposed to chemical weathering, which increases the dissolution rate. Therefore, non-aggregated smaller particles are affected by soil pH and microorganisms instantly and are more prone to dissolution [21,26,27]. Thus, TiO2 NP may be among those Ti forms which dissolve first after being exposed to the weathering effects of acidic soil conditions, plant and microbial activity. During weathering, a significant mobilization of Ti in the soil due to acidification by inorganic and organic acids may occur [5,28].
Soil properties affect nanoparticle solubility, dispersion and aggregation, with soil pH being the dominant factor. Therefore we have turned our attention towards particle size distribution and solubility of TiO2 NP in acidic and alkaline soil. The selected particle size ranges were 1–100, 100–450 and 450–1000 nm. The concentration of dissolved Ti was measured by centrifugal ultrafiltration. The distribution of TiO2 NP in soils is not only affected by nanoparticle dissolution and aggregation but also by their accumulation in the tissues of soil organisms. The filamentous fungus Aspergillus niger has been found to be capable of producing a high amount of organic compounds that chelate metals and solubilize mineral phases. Furthermore, it can reduce pH levels in the surrounding soil environment [29,30,31] and thus has a greater potential to accumulate TiO2 NP and Ti released from the nanoparticles. Moreover, A. niger has been observed to transform and mineralize biogenic TiO2 NP [32,33,34]. For all these reasons, this particular species appeared to be a suitable one on which to experiment for the purpose of evaluation of fungal ability to accumulate TiO2 NP from a liquid growth medium.

2. Materials and Methods

2.1. Soils Used in the Experiment

Samples of two soils differing in their characteristics were collected from 5 to 15 cm depth at two sites in the western part of Slovakia. The soils were classified according to the Morphogenetic Soil Classification System of Slovakia [35] and the World Reference Base for Soil Resources (WRB) [36] as Cernozem kultizemna karbonatova/Calcic Chernozem (CH-cc) and Kambizem luvizemna pseudoglejova/Dystric Stagnic Cambisol (CM-st.dy). CH-cc is an agricultural soil and was selected as a representative of soils with neutral to alkaline pH, while CM-st.dy represented natural forest soils with acidic pH. Digestion of 1 g of soil by the acid mixture of HF + HNO3 + HClO4 + H2O2 (4:3:4:1, 60 mL) in an open system at 200 °C was used to determine total concentrations of Al, Fe, Mn, Ti and Zn. Oxalate-extractable fractions of Al, Fe and Mn were determined using extraction with 0.2 mol L−1 ammonium oxalate stabilized at pH 3 on samples ground to a particle size below 0.5 mm [37]. The method by Walkley and Black [38] was used to determine total organic carbon (TOC). Soil pH was determined in 1 mol L−1 KCl (pHKCl) and deionized water (pHH₂O) at a 1:2.5 soil to solution ratio. The concentration of CaCO3 was determined by gas volumetric measurement of CO2 evolved upon 10% HCl treatment using calcimeter apparatus [39]. The size distribution of soil particles was determined by a pipette method [39]. The USDA-FAO soil-texture triangle was used to classify soil samples into textural classes [40]. Concentrations of humic substances, humic acids and fulvic acids in soil samples were determined by the method developed by Kononova and Belchikova [41]. The soil cation exchange capacity (CEC) was measured by the Kappen method [42]. The characteristics of the two soils are shown in Table S1 (Supplementary Material).

2.2. Batch Experiment with Soil

TiO2 NP (nanopowder, 21 nm primary particle size (TEM), ≥99.5% trace metals basis, Prod. No 718467, Sigma-Aldrich, Saint-Louis, MO, USA) used in the experiment were characterized with X-ray diffraction (XRD) (Bruker, MA, USA) and scanning electron microscopy (SEM) (JEOL JSM-7610F Plus, JEOL, Tokyo, Japan) and their detailed characteristics were reported in the previous work [11]. Results from SEM showed that TiO2 NP were of nearly spherical shape with some nanocrystals being prismatic with bi-pyramidal termination. The XRD results revealed two distinct mineral phases, anatase and rutile, with their mean sizes calculated by XRD TOPAS 3.0 software (Burker, MA, USA) being 30.0 ± 2.0 and 19.6 ± 0.2 nm, respectively.
Suspension with the TiO2 NP concentration of 100 mg L−1 was prepared by adding 0.50 mL of artificial rainwater concentrate (See Supplementary Material Section 2) and 50 mg of TiO2 NP to a 500 mL volumetric flask that was then filled up to the mark with distilled water. The TiO2 NP suspension was prepared shortly before the experiment and was placed in an ultrasonic bath for 15 min. This concentration was selected based on previous reports using a similar concentration in other experiments related to the transport of nanoparticles in soils [43].
Batch sorption experiments were conducted to explore the distribution of TiO2 NP in liquid- and solid-phase and to determine the amount of dissolved Ti in soil solutions. The 15 mL centrifuge tubes were filled with 0.790 mg of either CH-cc or CM-st.dy. Then, 15.8 mL of 100 mg L−1 TiO2 NP suspension was added to the tubes. In a control sample, 15.8 mL of artificial rainwater was added. Each experimental set-up was done in triplicate. The sealed centrifugal tubes were mounted in an overhead shaker and shaken at 5 rpm for 24 h. After 24-h batch sorption, the centrifuge tubes were sequentially centrifuged at 1000× g for 1 min, 2300× g for 2 min and 3000× g for 30 min to separate particles or aggregates smaller than 1000 nm (C1000), 450 nm (C450) and 100 nm (C100), respectively. Selected centrifugal forces and times were calculated according to the Stokes’ law for particles made of TiO2 [44,45,46]. After each step of centrifugation, 2 mL of supernatant was removed for the analyses of Ti content. After the last centrifugation step, 6 mL of supernatant was removed and added to ultrafiltration centrifugation units (Sartorius Vivaspin® 6 mL, 3 kDa), and the units were centrifuged for 20 min at 3500× g. Then, 1.5 mL of filtrate was removed to analyze the sample for Ti content (C1). From the acquired Ti concentrations, the concentrations of Ti present in different particle size fractions were calculated: 450–1000 nm = C1000–C450; 100–450 nm = C450–C100; 1–100 nm = C100–C1; and dissolved Ti (C1). The concentrations of Ti added in soil in the form of TiO2 NP were calculated as a difference between Ti concentrations determined in the experiments with TiO2 NP and concentrations of Ti naturally present on a site (control samples).

2.3. Cultivation of Fungus Treated with TiO2 Nanoparticles

For the experiment on a microscopic filamentous fungus Aspergillus niger (Tiegh.) supplemented with different concentrations of TiO2 NP, the fungus was grown in Sabouraud growth medium (HiMedia, Mumbai, India) in Erlenmeyer flasks for 7 days at 25 °C in the dark (static cultivation). Aspergillus niger (Tiegh.) strain CBS 140,837 used in the experiments was originally isolated from the mercury-contaminated soil [30]. To each of the Erlenmeyer flasks, 50 mL of Sabouraud growth medium was added. Different amounts of TiO2 NP, 10, 20, 40, 80, 160, 320, 640 and 1280 mg were added to the growth media to obtain growth media with concentrations of 200, 400, 800, 1600, 3200, 6400, 12,800 and 25,600 mg L−1 TiO2 NP, respectively. Every concentration and a control sample with no added TiO2 NP were tested in triplicate. Before inoculation, all of the Erlenmeyer flasks were placed in an ultrasonic bath for 15 min. Inoculation was accomplished by adding 50 μL of A. niger spore suspension to the growth media. After a 7-day cultivation period, the pH of the growth media was measured, and fungal mycelia were removed, dried out at 60 °C and weighed.
For an experiment in which the rates of Aspergillus-mediated dissolution of TiO2 NP were measured, a concentration of 400 mg L−1 TiO2 NP in growth media was used. It was the highest concentration that did not reduce the dry weight of fungal mycelia significantly. Erlenmeyer flasks were filled with 50 mL of Sabouraud growth medium and 20 mg of TiO2 NP. The flasks were then placed in an ultrasonic bath for 15 min prior to inoculation. A. niger was inoculated and grown in the same manner as in the experiment with different concentrations of TiO2 NP. The control sample was not spiked with TiO2 NP. The experiment was conducted in six replicates for both TiO2 NP amended growth media and the control. After the 7-day cultivation period, the pH of the growth media was measured, and fungal mycelia were carefully removed by tweezers, dried out at 60 °C and weighed. The dry biomass was then analyzed for Ti content. Growth media were removed and centrifuged at 1000× g for one minute to separate TiO2 NP particles and aggregates smaller than 1000 nm. After the centrifugation, 2 mL of supernatant was pipetted off and sent for the analysis of Ti content. The dissolved Ti was obtained by transferring 6 mL of supernatant that was added to ultrafiltration centrifugation units (Sartorius Vivaspin®, Göttingen, Germany, 6 mL, 3 kDa) and the units were then centrifuged for 20 min at 3500× g. Afterwards, 1.5 mL of filtrate was removed to analyze the sample for Ti content.

2.4. Determination of Ti Concentration by ICP-MS

The collected supernatants, filtrates and the dry biomass were digested in the mixture of HNO3 and HF acids by microwave digestion at 1200 W for 30 min (Anton Paar Multiwave 3000, Anton Paar, Graz, Austria). The digested samples were analyzed for Ti content by ICP-MS (SCIEX ELAN 6000, Perkin-Elmer SCIEX, Quebec, Canada). The isotope of 50Ti was used to collect the data and 103Rh was used as an internal standard.

2.5. Statistical Analysis

Concentrations of Ti in dry biomass and growth media, the dry weight of mycelia and pH were compared for all treatment conditions via two-tail t-test at a significance level α = 0.05. Prior to t-test, data were analyzed for differences in variances by F-test and then either t-test for equal or t-test for unequal variances was used. The statistical evaluation was completed using Analysis ToolPak add-in for Microsoft Excel (Redmond, WA, USA).

3. Results

3.1. Batch Experiment with Soil

After being treated with TiO2 NP, two soils of different pH values, alkaline agricultural Calcic Chernozem (CH-cc) and strongly acidic forest Dystric Stagnic Cambisol (CM-st.dy) (Table S1) were studied for the distribution of mobile Ti in aggregates in different particle size ranges. A substantial difference in soil pH was selected deliberately to investigate the extent to which a pH affects the TiO2 NP dissolution.
Distribution of colloidal and ionic forms in mobile Ti fraction was measured in aqueous suspensions of untreated (TiO2 NP-free soil) and TiO2 NP-treated soils (contaminated soil). These experimental results can be seen in Figure 1, which shows the concentrations of Ti in different particle size fractions and the retention of TiO2 NP in CM-st.dy and CH-cc soils.
The concentration of dissolved or ionic Ti in untreated soils was below the 0.005 mg L−1 limit of quantification (Figure 1a) and only up to 0.12 mg L−1 of Ti was associated with particles in 1–100 nm size range (Figure 1b). Most of the Ti was found in colloidal fraction with the highest volume of Ti bound to particles or aggregates within a 450 to 1000 nm size range (Figure 1d,e). The fraction of particles sized 100–450 nm contained the second-largest amount of Ti (Figure 1c,e).
In the TiO2 NP-treated soils, a concentration of an ionic Ti reached 0.010 and 0.005 mg L−1 in CH-cc and CM-st.dy suspensions, respectively (Figure 1a). A large portion of Ti was bound to particles and aggregates sized between 1 to 100 nm (Figure 1b,e,f) and to those in a 100–450 nm size range (Figure 1c,e,f). Lesser amounts of TiO2 NP were accumulated in 450–1000 nm size fraction of soil suspension (Figure 1d,e,f). Less than 30% of added TiO2 NP was retained in both soils (Figure 1g).

3.2. Mutual Interactions of Fungus and TiO2 NP

Various doses of TiO2 NP were supplemented into the growth media to evaluate the fungal growth response to nanoparticles (Figure 2). The addition of TiO2 NP, up to a concentration of 400 mg L−1, caused no statistically significant weight changes in fungal biomass. The pH of the nanoparticle-amended growth medium was similar to that of TiO2 NP-free medium. However, concentrations over 800 mg L−1 were found to have statistically significant negative effects on the development of A. niger and reduced the dry weight of the mycelia. Hence, the 400 mg L−1 TiO2 NP concentration was selected for further experiment.
At an initial TiO2 NP concentration of 400 mg L−1, fungus A. niger acidified its environment (i.e. growth medium) to a value as low as 2.5. The fungus accumulated 3.6 mg g−1 of Ti in the dry mycelia. The remaining Ti was mainly associated with aggregated sediments. Approximately 90% of the TiO2 NP settled during cultivation (Figure 3), while 7.5% were bioaccumulated or incorporated into the growing biomass mechanically. Less than 2.5% remained in the form of particles and aggregates with a size up to 1000 nm and only 0.01% (0.030 mg L−1) was dissolved in the growth media during cultivation.

4. Discussion

4.1. Batch Experiment with Soils

Our results suggest that the properties of alkaline agricultural soil and properties of an acidic forest soil affect the distribution of TiO2 NP in soil suspension in a similar way (Figure 1f). If pH was the only factor affecting the TiO2 NP distribution, this finding would be quite unexpected (lower TiO2 NP retention would be expected for CH-cc compared to CM-st.dy) [47]. Since both soils differ significantly in clay and organic carbon content, we hypothesize that two concurrent factors affected the observed behaviour: (1) Humic acids stabilize nanoparticles and their aggregates by adsorbing onto their surfaces. This process is accompanied by the development of a negative surface charge which leads to the stabilization of the nanoparticles in a wide soil pH range [48,49,50]. (2) High content of clay and silt fraction in CH-cc provides a larger surface area on which TiO2 NP can be easily attached to or TiO2 NP can make aggregates with this soil fraction and, thus, remain suspended in soil solutions [51,52]. Therefore, even though the pH values differ significantly, the combined effect of the two factors mentioned above results in a similar size fractionation of Ti in both soil solutions.
Some level of homoaggregation of TiO2 NP may have also occurred [53] and it may have resulted in a higher content of Ti detected in 1–100 nm and 100–450 nm size fractions in TiO2 NP-treated soils (Figure 1e,f). Only a negligible amount of Ti from the TiO2 NP was released into solution in both soil types studied. This is in line with observations where even in very acidic soils, the Ti dissolution is minimal, although it does occur naturally [3].
Compared to real soil conditions, soil batch experiments cannot simulate some factors that influence the transport of nanoparticles in soil. Under natural conditions, soils have a complex structure with micro- and macropores. Moreover, soil organisms and plants help to form additional pores in soil. These factors heavily affect the porosity and hydraulic conductivity, and, thus, the distribution of nanoparticles in the soil environment. Failing to account for these factors can lead to both underestimation and overestimation of the nanoparticle mobility in soils [21,27,54].

4.2. Fungal Accumulation of TiO2 Nanoparticles

Microscopic filamentous fungi mechanically and chemically disintegrate soil particles and fasten the weathering process [55]. However, Ti minerals are one of the most stable solid soil phases [4]. The accumulation of Ti in fungi or plants should thus be minimal. Even with the biologically enhanced mobilization induced by acidolysis and complexolysis, only a negligible amount of the dissolved Ti is released to the soil solutions [4,5,28,56].
A. niger acidified its environment to pH 2.5. It has been reported that under such conditions, the fungus also releases large amounts of organic acids and other biomolecules that enable the chelation and complexolysis [29,31,57]. However, in our experiment, only 0.01% of the initial content of Ti was detected in growth media in a solubilized form (Figure 3). Thus, we hypothesize that A. niger strain’s mycelia accumulated most of Ti in the form of TiO2 NP aggregates. However, the design of the experiment has not allowed for an accurate estimation of the volume of TiO2 NP that has been accumulated on the mycelial surface and the amount of TiO2 NP that have been absorbed. Likewise, we do not know whether Ti present in the fungus was first dissolved and then absorbed and bound to organic molecules in fungal cells or if A. niger reprecipitated the dissolved portion of Ti as biogenic TiO2 NP inside its cells. Such bioaccumulation of TiO2 NP may affect the spatial distribution of Ti in soils, and soil fungal biomass can act as a source of Ti for other organisms. The concentration of Ti accumulated in the mycelia (3.6 g kg−1 dry weight) was higher than Ti concentration in roots of barley grown in the hydroponic system (0.26 to 1.49 g kg−1 dry weight) at TiO2 NP content of 100 to 1000 mg L−1 [12].
The aggregation of TiO2 NP was also accompanied by the settling of agglomerated particles to the bottom of the Erlenmeyer flask; the aggregation was likely enhanced by secretion of extracellular polymeric substances by A. niger. These exudates attach to nanoparticle surfaces and promote aggregation via alteration of surface properties [58,59,60,61].

5. Conclusions

Our results indicate that Aspergillus niger strain and other similar soil fungi are organisms capable of redistributing Ti in soil horizons. They can bind relatively high amounts of TiO2 NP into their mycelia. Soil fungi and other soil microorganisms able to produce organic and inorganic acids should be considered when evaluating distribution and environmental risks of Ti and other elements conventionally referred to as immobile since they contribute to the accumulation and redistribution of these elements in soils. Future research should refine the estimate of the amounts of Ti attached to mycelial surfaces and Ti accumulated in the mycelia using more precise methods. Moreover, since A. niger has a potential to dissolve and remineralize TiO2 NP, laboratory experiments able to adequately capture this process would provide unique insights into the behaviour of fungus during bioaccumulation.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4395/10/11/1833/s1, Table S1: Characteristics of Calcic Chernozem (CH-cc) and Dystric Stagnic Cambisol (CM-st.dy) soil samples.

Author Contributions

Conceptualization, L.N., M.Š. and M.U.; data curation, M.Š.; formal analysis, M.Š.; funding acquisition, M.U. and P.M.; investigation, M.Š.; methodology, M.Š., M.U. and P.M.; project administration, P.M.; resources, M.K., M.B. and I.H.; supervision, M.U., M.K. and P.M.; validation, M.B.; visualization, M.Š.; writing—original draft, M.Š.; writing—review and editing, L.N. and M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Scientific Grant Agency of the Ministry of Education, Science, Research and Sport of the Slovak Republic and the Slovak Academy of Sciences (Vedecká grantová agentúra MŠVVaŠ SR a SAV) under the contracts Nos. VEGA 1/0153/17, VEGA 1/0164/17, and VEGA 1/0146/18. And by Cultural and Educational Agency of the Ministry of Education, Science, Research and Sport of the Slovak Republic (Kultúrna a edukačná grantová agentúra MŠVVaŠ SR) under the contract KEGA 013SPU-4/2019.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Orians, K.J.; Boyle, E.A.; Bruland, K.W. Dissolved titanium in the open ocean. Nature 1990, 348, 322–325. [Google Scholar] [CrossRef]
  2. Van Baalen, M.R. Titanium mobility in metamorphic systems: A review. Chem. Geol. 1993, 110, 233–249. [Google Scholar] [CrossRef]
  3. Cornu, S.; Lucas, Y.; Lebon, E.; Ambrosi, J.P.; Luizão, F.; Rouiller, J.; Bonnay, M.; Neal, C. Evidence of titanium mobility in soil profiles, Manaus, central Amazonia. Geoderma 1999, 91, 281–295. [Google Scholar] [CrossRef]
  4. Zierden, M.R.; Valentine, A.M. Contemplating a role for titanium in organisms. Metallomics 2016, 8, 9–16. [Google Scholar] [CrossRef]
  5. Buettner, K.M.; Valentine, A.M. Bioinorganic Chemistry of Titanium. Chem. Rev. 2012, 112, 1863–1881. [Google Scholar] [CrossRef]
  6. Tan, W.; Peralta-Videa, J.R.; Gardea-Torresdey, J.L. Interaction of titanium dioxide nanoparticles with soil components and plants: Current knowledge and future research needs—A critical review. Environ. Sci. Nano 2018, 5, 257–278. [Google Scholar] [CrossRef]
  7. Hrubý, M.; Cígler, P.; Kužel, S. Contribution to understanding the mechanism of titanium action in plants. J. Plant Nutr. 2002, 25, 577–598. [Google Scholar] [CrossRef]
  8. Kužel, S.; Hrubý, M.; Cígler, P.; Tlustoš, P.; Van Nguyen, P. Mechanism of physiological effects of titanium leaf sprays on plants grown on soil. Biol. Trace Elem. Res. 2003, 91, 179–189. [Google Scholar] [CrossRef]
  9. Kolenčík, M.; Nemček, L.; Šebesta, M.; Urík, M.; Ernst, D.; Kratošová, G.; Konvičková, Z. Effect of TiO2 as Plant-Growth Stimulating Nanomaterial on Crop Production. In Plant Responses to Nanomaterials; Singh, V.P., Singh, S., Prasad, S.M., Chauhan, D.K., Tripathi, D.K., Eds.; Springer Nature: Berlin/Heidelberg, Germany, 2020; In Press. [Google Scholar]
  10. Šebesta, M.; Illa, R.; Zvěřina, O.; Šeda, M.; Diviš, P.; Kolenčík, M. Positive Effects of TiO2 Nanomaterials on Crop Growth. In Nanotechnology for Plant Diseases; Ingle, A.P., Ed.; Wiley-VCH: Weinheim, Germany, 2020; In Press. [Google Scholar]
  11. Kolenčík, M.; Ernst, D.; Urík, M.; Ďurišová, Ľ.; Bujdoš, M.; Šebesta, M.; Dobročka, E.; Kšiňan, S.; Illa, R.; Qian, Y.; et al. Foliar Application of Low Concentrations of Titanium Dioxide and Zinc Oxide Nanoparticles to the Common Sunflower under Field Conditions. Nanomaterials 2020, 10, 1619. [Google Scholar] [CrossRef]
  12. Kořenková, L.; Šebesta, M.; Urík, M.; Kolenčík, M.; Kratošová, G.; Bujdoš, M.; Vávra, I.; Dobročka, E. Physiological response of culture media-grown barley (Hordeum vulgare L.) to titanium oxide nanoparticles. Acta Agric. Scand. Sect. B—Soil Plant Sci. 2017, 67, 285–291. [Google Scholar] [CrossRef]
  13. Rajwade, J.M.; Chikte, R.G.; Paknikar, K.M. Nanomaterials: New weapons in a crusade against phytopathogens. Appl. Microbiol. Biotechnol. 2020, 104, 1437–1461. [Google Scholar] [CrossRef]
  14. Yan, J.; Huang, K.; Wang, Y.; Liu, S. Study on anti-pollution nano-preparation of dimethomorph and its performance. Chin. Sci. Bull. 2005, 50, 108–112. [Google Scholar] [CrossRef]
  15. Guan, H.; Chi, D.; Yu, J.; Li, H. Dynamics of residues from a novel nano-imidacloprid formulation in soyabean fields. Crop Prot. 2010, 29, 942–946. [Google Scholar] [CrossRef]
  16. Mohamed, M.M.; Khairou, K.S. Preparation and characterization of nano-silver/mesoporous titania photocatalysts for herbicide degradation. Microporous Mesoporous Mater. 2011, 142, 130–138. [Google Scholar] [CrossRef]
  17. Cartwright, A.; Jackson, K.; Morgan, C.; Anderson, A.; Britt, D.W. A Review of Metal and Metal-Oxide Nanoparticle Coating Technologies to Inhibit Agglomeration and Increase Bioactivity for Agricultural Applications. Agronomy 2020, 10, 1018. [Google Scholar] [CrossRef]
  18. Priyanka, K.P.; Harikumar, V.S.; Balakrishna, K.M.; Varghese, T. Inhibitory effect of TiO2 nanoparticles on symbiotic arbuscular mycorrhizal fungi in plant roots. IET Nanobiotechnology 2016, 11, 66–70. [Google Scholar] [CrossRef]
  19. Rafique, R.; Zahra, Z.; Virk, N.; Shahid, M.; Pinelli, E.; Park, T.J.; Kallerhoff, J.; Arshad, M. Dose-dependent physiological responses of Triticum aestivum L. to soil applied TiO2 nanoparticles: Alterations in chlorophyll content, H2O2 production, and genotoxicity. Agric. Ecosyst. Environ. 2018, 255, 95–101. [Google Scholar] [CrossRef]
  20. Mattiello, A.; Lizzi, D.; Marchiol, L. Influence of Titanium Dioxide Nanoparticles (nTiO2) on Crop Plants: A Systematic Overview. Nanomater. Plants Algae Microorg. 2018, 277–296. [Google Scholar] [CrossRef]
  21. Šebesta, M.; Kolenčík, M.; Matúš, P.; Kořenková, L. Transport and distribution of engineered nanoparticles in soils and sediments. Chem. List. 2017, 111, 322–328. [Google Scholar]
  22. Šebesta, M.; Matúš, P. Separation, Determination, and Characterization of Inorganic Engineered Nanoparticles in Complex Environmental Samples. Chem. List. 2018, 112, 583–589. [Google Scholar]
  23. Ben-Noah, I.; Friedman, S.P. Review and Evaluation of Root Respiration and of Natural and Agricultural Processes of Soil Aeration. Vadose Zone J. 2018, 17, 170119. [Google Scholar] [CrossRef] [Green Version]
  24. Adeleke, R.; Nwangburuka, C.; Oboirien, B. Origins, roles and fate of organic acids in soils: A review. South Afr. J. Bot. 2017, 108, 393–406. [Google Scholar] [CrossRef]
  25. Burford, E.P.; Fomina, M.; Gadd, G.M. Fungal involvement in bioweathering and biotransformation of rocks and minerals. Mineral. Mag. 2003, 67, 1127–1155. [Google Scholar] [CrossRef]
  26. Klaine, S.J.; Alvarez, P.J.J.; Batley, G.E.; Fernandes, T.F.; Handy, R.D.; Lyon, D.Y.; Mahendra, S.; McLaughlin, M.J.; Lead, J.R. Nanomaterials in the environment: Behavior, fate, bioavailability, and effects. Environ. Toxicol. Chem. 2008, 27, 1825–1851. [Google Scholar] [CrossRef]
  27. Peijnenburg, W.; Praetorius, A.; Scott-Fordsmand, J.; Cornelis, G. Fate assessment of engineered nanoparticles in solids dominated media—Current insights and the way forward. Environ. Pollut. 2016, 218, 1365–1369. [Google Scholar] [CrossRef]
  28. Welch, S.A.; Christy, A.G.; Kirste, D.; Beavis, S.G.; Beavis, F. Jarosite dissolution I—Trace cation flux in acid sulfate soils. Chem. Geol. 2007, 245, 183–197. [Google Scholar] [CrossRef]
  29. Polák, F.; Urík, M.; Matúš, P. Low molecular weight organic acids in soil environment. Chem. List. 2019, 113, 307–314. [Google Scholar]
  30. Urík, M.; Hlodák, M.; Mikušová, P.; Matúš, P. Potential of Microscopic Fungi Isolated from Mercury Contaminated Soils to Accumulate and Volatilize Mercury(II). Water Air Soil Pollut. 2014, 225, 2219. [Google Scholar] [CrossRef]
  31. Boriová, K.; Urík, M.; Bujdoš, M.; Pifková, I.; Matúš, P. Chemical mimicking of bio-assisted aluminium extraction by Aspergillus niger’s exometabolites. Environ. Pollut. 2016, 218, 281–288. [Google Scholar] [CrossRef]
  32. Qin, W.; Wang, C.; Ma, Y.; Shen, M.; Li, J.; Jiao, K.; Tay, F.R.; Niu, L. Microbe-Mediated Extracellular and Intracellular Mineralization: Environmental, Industrial, and Biotechnological Applications. Adv. Mater. 2020, 32, 1907833. [Google Scholar] [CrossRef]
  33. Raliya, R.; Biswas, P.; Tarafdar, J.C. TiO2 nanoparticle biosynthesis and its physiological effect on mung bean (Vigna radiata L.). Biotechnol. Rep. 2015, 5, 22–26. [Google Scholar] [CrossRef] [Green Version]
  34. Hulkoti, N.I.; Taranath, T.C. Biosynthesis of nanoparticles using microbes—A review. Colloids Surf. B Biointerfaces 2014, 121, 474–483. [Google Scholar] [CrossRef]
  35. Societas Pedologica Slovaca. Morphogenetic Soil Classification System of Slovakia, Basal Reference Taxonomy, 2nd ed.; Soil Science and Conservation Research Institute: Bratislava, Slovakia, 2014; ISBN 978-80-8163-005-7. [Google Scholar]
  36. IUSS Working Group. WRB World Reference Base for Soil Resources 2014, Update 2015 International Soil Classification System for Naming Soils and Creating Legends for Soil Maps; World Soil Resources Reports No. 106; FAO: Rome, Italy, 2015. [Google Scholar]
  37. McKeague, J.A.; Day, J. Dithionite-and oxalate-extractable Fe and Al as aids in differentiating various classes of soils. Can. J. Soil Sci. 1966, 46, 13–22. [Google Scholar] [CrossRef]
  38. Walkley, A.; Black, I.A. An examination of the Degtjareff method for determining soil organic matter, and a proposed modification of the chromic acid titration method. Soil Sci. 1934, 37, 29–38. [Google Scholar] [CrossRef]
  39. Fiala, K.; Kobza, J.; Matúšková, Ľ.; Brečková, V.; Makovníková, J.; Barančíková, G.; Búrik, V.; Litavec, T.; Houšková, B.; Chromaničová, A. Záväzné Metódy Rozborov Pôd; VÚPOP: Bratislava, Slovakia, 1999. [Google Scholar]
  40. FAO. Guidelines for Soil Description, 4th ed.; FAO: Rome, Italy, 2006; ISBN 92-5-105521-1. [Google Scholar]
  41. Kononova, M.M.; Belčikova, N.P. Uskorennyje metody opredelenija sostava gumusa minerálnych počv. Počvovedenije 1962, 10, 75–87. [Google Scholar]
  42. Kappen, H. Die Bodenazidität; Springer: Berlin/Heidelberg, Germany, 1929; ISBN 978-3-642-89928-7. [Google Scholar]
  43. Chowdhury, I.; Hong, Y.; Honda, R.J.; Walker, S.L. Mechanisms of TiO2 nanoparticle transport in porous media: Role of solution chemistry, nanoparticle concentration, and flowrate. J. Colloid Interface Sci. 2011, 360, 548–555. [Google Scholar] [CrossRef]
  44. Gibbs, R.J.; Matthews, M.D.; Link, D.A. The relationship between sphere size and settling velocity. J. Sediment. Res. 1971, 41, 7–18. [Google Scholar]
  45. Schwertfeger, D.M.; Velicogna, J.R.; Jesmer, A.H.; Saatcioglu, S.; McShane, H.; Scroggins, R.P.; Princz, J.I. Extracting Metallic Nanoparticles from Soils for Quantitative Analysis: Method Development Using Engineered Silver Nanoparticles and SP-ICP-MS. Anal. Chem. 2017, 89, 2505–2513. [Google Scholar] [CrossRef]
  46. Šebesta, M.; Nemček, L.; Urík, M.; Kolenčík, M.; Bujdoš, M.; Vávra, I.; Dobročka, E.; Matúš, P. Partitioning and stability of ionic, nano- and microsized zinc in natural soil suspensions. Sci. Total Environ. 2020, 700, 134445. [Google Scholar] [CrossRef]
  47. Fang, J.; Xu, M.; Wang, D.; Wen, B.; Han, J. Modeling the transport of TiO2 nanoparticle aggregates in saturated and unsaturated granular media: Effects of ionic strength and pH. Water Res. 2013, 47, 1399–1408. [Google Scholar] [CrossRef]
  48. Loosli, F.; Le Coustumer, P.; Stoll, S. TiO2 nanoparticles aggregation and disaggregation in presence of alginate and Suwannee River humic acids. pH and concentration effects on nanoparticle stability. Water Res. 2013, 47, 6052–6063. [Google Scholar] [CrossRef]
  49. Chen, G.X.; Liu, X.Y.; Su, C.M. Distinct Effects of Humic Acid on Transport and Retention of TiO2 Rutile Nanoparticles in Saturated Sand Columns. Environ. Sci. Technol. 2012, 46, 7142–7150. [Google Scholar] [CrossRef]
  50. Zhang, R.; Tu, C.; Zhang, H.; Luo, Y. Stability and transport of titanium dioxide nanoparticles in three variable-charge soils. J. Soils Sediments 2020, 20, 1395–1403. [Google Scholar] [CrossRef]
  51. Buffle, J.; Wilkinson, K.J.; Stoll, S.; Filella, M.; Zhang, J. A Generalized Description of Aquatic Colloidal Interactions: The Three-colloidal Component Approach. Environ. Sci. Technol. 1998, 32, 2887–2899. [Google Scholar] [CrossRef]
  52. Fang, J.; Shan, X.; Wen, B.; Lin, J.; Owens, G. Stability of titania nanoparticles in soil suspensions and transport in saturated homogeneous soil columns. Environ. Pollut. 2009, 157, 1101–1109. [Google Scholar] [CrossRef]
  53. Zehlike, L.; Peters, A.; Ellerbrock, R.H.; Degenkolb, L.; Klitzke, S. Aggregation of TiO2 and Ag nanoparticles in soil solution—Effects of primary nanoparticle size and dissolved organic matter characteristics. Sci. Total Environ. 2019, 688, 288–298. [Google Scholar] [CrossRef]
  54. Cornelis, G. Fate descriptors for engineered nanoparticles: The good, the bad, and the ugly. Environ. Sci. Nano 2015, 2, 19–26. [Google Scholar] [CrossRef]
  55. Gadd, G.M. (Ed.) Fungi in Biogeochemical Cycles; Cambridge University Press: Cambridge, UK, 2006; ISBN 9780511550522. [Google Scholar]
  56. Jonglertjunya, W.; Rubcumintara, T. Titanium and iron dissolutions from ilmenite by acid leaching and microbiological oxidation techniques. Asia-Pac. J. Chem. Eng. 2013, 8, 323–330. [Google Scholar] [CrossRef]
  57. Polák, F.; Urík, M.; Bujdoš, M.; Matúš, P. Aspergillus niger enhances oxalate production as a response to phosphate deficiency induced by aluminium(III). J. Inorg. Biochem. 2020, 204, 110961. [Google Scholar] [CrossRef]
  58. Crampon, M.; Hellal, J.; Mouvet, C.; Wille, G.; Michel, C.; Wiener, A.; Braun, J.; Ollivier, P. Do natural biofilm impact nZVI mobility and interactions with porous media? A column study. Sci. Total Environ. 2018, 610, 709–719. [Google Scholar] [CrossRef]
  59. Jiang, X.; Wang, X.; Tong, M.; Kim, H. Initial transport and retention behaviors of ZnO nanoparticles in quartz sand porous media coated with Escherichia coli biofilm. Environ. Pollut. 2013, 174, 38–49. [Google Scholar] [CrossRef]
  60. He, J.-Z.; Li, C.-C.; Wang, D.-J.; Zhou, D.-M. Biofilms and extracellular polymeric substances mediate the transport of graphene oxide nanoparticles in saturated porous media. J. Hazard. Mater. 2015, 300, 467–474. [Google Scholar] [CrossRef]
  61. Tarafdar, J.C.; Agrawal, A.; Raliya, R.; Kumar, P.; Burman, U.; Kaul, R.K. ZnO Nanoparticles Induced Synthesis of Polysaccharides and Phosphatases by Aspergillus Fungi. Adv. Sci. Eng. Med. 2012, 4, 324–328. [Google Scholar] [CrossRef]
Figure 1. Concentrations (ad) and distribution (e,f) of Ti in various particle size fractions in untreated and treated soil suspensions and the retention of Ti in contaminated soils (g). Lower-case letters indicate no significant differences between the bars that represents soils and their different treatments (p < 0.05).
Figure 1. Concentrations (ad) and distribution (e,f) of Ti in various particle size fractions in untreated and treated soil suspensions and the retention of Ti in contaminated soils (g). Lower-case letters indicate no significant differences between the bars that represents soils and their different treatments (p < 0.05).
Agronomy 10 01833 g001
Figure 2. Growth of Aspergillus niger at different TiO2 NP concentrations. * indicates significant differences between treatments and control (p < 0.05).
Figure 2. Growth of Aspergillus niger at different TiO2 NP concentrations. * indicates significant differences between treatments and control (p < 0.05).
Agronomy 10 01833 g002
Figure 3. Distribution of Ti in the cultivation system after incubation of A. niger in the presence of 400 mg L−1 TiO2 NP.
Figure 3. Distribution of Ti in the cultivation system after incubation of A. niger in the presence of 400 mg L−1 TiO2 NP.
Agronomy 10 01833 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Šebesta, M.; Nemček, L.; Urík, M.; Kolenčík, M.; Bujdoš, M.; Hagarová, I.; Matúš, P. Distribution of TiO2 Nanoparticles in Acidic and Alkaline Soil and Their Accumulation by Aspergillus niger. Agronomy 2020, 10, 1833. https://doi.org/10.3390/agronomy10111833

AMA Style

Šebesta M, Nemček L, Urík M, Kolenčík M, Bujdoš M, Hagarová I, Matúš P. Distribution of TiO2 Nanoparticles in Acidic and Alkaline Soil and Their Accumulation by Aspergillus niger. Agronomy. 2020; 10(11):1833. https://doi.org/10.3390/agronomy10111833

Chicago/Turabian Style

Šebesta, Martin, Lucia Nemček, Martin Urík, Marek Kolenčík, Marek Bujdoš, Ingrid Hagarová, and Peter Matúš. 2020. "Distribution of TiO2 Nanoparticles in Acidic and Alkaline Soil and Their Accumulation by Aspergillus niger" Agronomy 10, no. 11: 1833. https://doi.org/10.3390/agronomy10111833

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