Next Article in Journal
Extended π-Systems in Diimine Ligands in [Cu(P^P)(N^N)][PF6] Complexes: From 2,2′-Bipyridine to 2-(Pyridin-2-yl)Quinoline
Previous Article in Journal
Effects of Interfacial Reactions on Microstructures and Mechanical Properties of 3003 Al/T2 Cu and 1035 Al/T2 Cu Brazed Joints
 
 
Article
Peer-Review Record

Investigation of the Structural Changes and Catalytic Properties of FeNi Nanostructures as a Result of Exposure to Gamma Radiation

Crystals 2020, 10(4), 254; https://doi.org/10.3390/cryst10040254
by Daryn B. Borgekov 1,2, Maxim V. Zdorovets 1,2,3, Dmitriy I. Shlimas 1,2 and Artem L. Kozlovskiy 2,4,*
Reviewer 1: Anonymous
Reviewer 2: Anonymous
Crystals 2020, 10(4), 254; https://doi.org/10.3390/cryst10040254
Submission received: 10 February 2020 / Revised: 24 March 2020 / Accepted: 26 March 2020 / Published: 27 March 2020
(This article belongs to the Section Crystalline Materials)

Round 1

Reviewer 1 Report

The authors present the effect of gamma irradiation of NiFe nanotubes on the crystal- and micro- structure, as well as the impact on chemical resistance and catalytic activity as a function of gamma irradiation dose. The crystal- and micro- structure is investigated by SEM, EDX, and XRD. The degradation and corrosion of the nano structures in PBS is investigated. The catalytic activity as a function of gamma irradiation dose is investigated for the para-nitroaniline to para-phenylenediamine conversion using UV-VIS absorption spectroscopy. The authors show clear evidence of the positive effect of gamma irradiation on degradation/corrosion and catalytic activity.

I recommend publication after major revisions, as all experimental methods need more detailed description.

-- What type of SEM/EDX is used and what are the imaging conditions (acc. voltage, current, detection mode)?

-- What acceleration voltage is used for EDX? 

-- How is the composition Ni21Fe79 determined? Is it a point measurement or an average of one of the maps? Is it a standard based quantification or reference free method? What is the measurement error?

-- The scale bar is missing for all EDX maps.

-- What is actually displayed in the EDX maps? For a NiFe alloy? I would expect a homogeneous distribution of Fe an Ni across the entire nanotube. The maps look more like Ni tubes with Fe patches?

-- More details about the XRD measurements has to be given. What X-ray source is used? Cu-Kalpha? Is it a Bragg-Brentano geometry? How do the authors extract information like crystal size from the XRD data?

-- I am a bit concerned about the extraction / over interpretation of size, strain, micro strain, and defect density from the XRD data at high dose and therefore high texture. Even with WPPM methods, it is almost impossible to separate size and micro strain based on only one observable diffraction peak.

I do agree though that from a qualitative point of view the interpretation of the XRD data seems to make sense.

-- What are the solid red lines in Fig. 4 and 8, is it a fit with some sort of theory?

-- How is the degradation determined and what is measured experimentally (kinetic curve of what?) to determine the degradation of the nanostructures? More details needs to be given.

-- It would be nice to get more details about the UV-VIS measurements, e.g. what is the starting concentration, is it an in-situ or ex-situ experiment, ... .

Author Response

 

The morphological features, as well as the elemental composition of the synthesized nanostructures, were studied using the scanning electron microscopy method performed using “Hitachi TM3030” scanning electron microscope (Hitachi Ltd., Chiyoda, Tokyo, Japan). Shooting mode - LEI, Current – 20 μA, Accelerate voltage – 2 kV, WD= 8 mm. 

The study of the elemental composition, as well as the mapping of the structures under study in order to determine the equiprobable distribution of elements in the structure, was carried out using the energy dispersive analysis method performed using the EDA Bruker Flash MAN SVE installation (Bruker, Karlsruhe, Germany), при ускоряющем напряжении 15 kV.

The influence of the ratio of the components of the electrolyte solution and the difference in the applied potentials on the elemental composition and isotropy of the distribution of the components in the structure of the nanotubes was determined using the energy dispersive analysis method by taking 10 spectra from different sections of the synthesized nanostructures, as well as mapping the elements along the entire length. After taking 10 spectra from different sections of the nanostructures, we calculated the average values of both components, as well as determine the magnitude of the error, which was no more than 2 at. %.

A scale bar has been added.

The data are presented in a more complete form, which displays the distribution of each element separately.

 

As can be seen from the data presented, the elemental ratio of the components of iron and nickel in the structure is close to permalloy compounds, which are characterized by the ratio of the components Fe:Ni = 20:80. A slight deviation of this ratio for the synthesized nanostructures may be due to nucleation processes that occur during the preparation of nanostructures. According to the mapping data, the distribution of iron in nickel is equally probable over the entire length of the nanostructures, which indicates the isotropy of the growth rate of nanostructures and the same rate of reduction of metal ions from electrolyte solutions during synthesis.

 

The crystallographic characteristics of nanostructures before and after directed modification were studied using a D8 Advance Eco powder X-ray diffractometer (Bruker, Karlsruhe, Germany). Conditions for recording diffractograms: 2θ = 35–80°, step 0.01°, Bragg-Brentano geometry, spectrum acquisition time 3 s, sample table rotation speed 20 rpm, X-ray radiation Cu-Kα, λ = 1.54 Å. For X-ray diffraction, all samples before and after modification were in a polymer matrix, which was mounted on the sample holder perpendicular to the x-ray beam. The rotation of the sample table was carried out to minimize the effect of the difference in textures at different shooting angles. An analysis of the structural characteristics, as well as the phase composition, was carried out using TOPAS v.5.0 software based on the Rietveld method. The size of the crystallites, which was calculated according to the Scherrer equation:

 

(1)

where k = 0.9 is the dimensionless particle shape coefficient (Scherrer constant), λ=1,54Å is the X-ray wavelength, β is the half-width of the reflex at half maximum (FWHM), and θ is the diffraction angle (Bragg angle).

Microstresses were estimated based on an analysis of the displacement of diffraction peaks calculated according to equation (2):

,

(2)

where dpristine and dirr are the interplanar distances before and after irradiation. The lattice stresses and macrostresses in the structure were estimated for all diffraction peaks.

The degree of crystallinity or structural ordering was determined by approximating the diffraction lines by the required number of pseudo-Voigt symmetric functions, and also by measuring the width of the recorded lines at half their height (FWHM). The pseudo-Voigt functions used to approximate the X-ray peak profile in the diffractogram (3) [43,44]:

,

(3)

 

 

 

where x is the variable corresponding to the angle of reflection 2θ; x0 - sets the position of the maximum of the function; η is the specific fraction of the Lorentz function; A is the normalizing factor; bG and bL are the parameters of the Gauss functions G(x, x0, bG) and Lorentz L(x, x0, bL).

 

An analysis of the angular dependence of physical broadening allows one to evaluate the influence of factors such as the size effect associated with crushing or recrystallization of crystallites as a result of external influences, and a distorting factor depending on the degree of deformation of the crystal structure and its change as a result of external influences. To assess the impact, the Williamson-Hall method was used, which is based on relation (4) [45]:

,

,

,

(4)

where β is physical broadening of the diffraction maximum, λ is the x-ray wavelength (1.54 Å), D is the crystallite size, θ is the Bragg diffraction angle, ε is the magnitude of microstresses in the grating.

Microstresses were estimated based on an analysis of the displacement of diffraction peaks calculated according to equation (5):

,

(5)

where dpristine and dirr are the interplanar distances before and after irradiation.

 

In the case of Figure 4, the dots indicate the experimental data, the red lines indicate the approximation lines of the results obtained, which are necessary to determine the rate of degradation reaction.

In the case of Figure 8, the dots indicate the experimental data, the red lines indicate the approximation lines of the results.

Kinetic curves are a collection of data on changes in the elemental composition and mass of samples during oxidation and degradation, obtained using methods of energy dispersive analysis, x-ray phase analysis and by weighing samples before and after corrosion tests.

 

Figure 7 shows the results of changes in mass loss during degradation. According to the data obtained, the change in mass occurs in two stages. The first stage is described by a positive weight gain, which is due to the formation of an oxide film on the surface of the nanotubes. In this case, an increase in the radiation dose and, consequently, a change in the structural characteristics of nanotubes leads to an increase in the number of days at which a positive change in mass is observed, which indicates that the formed oxide film reduces the rate of degradation of nanotubes. The second stage of the change is characteristic of a negative weight gain, which is caused by the processes of degradation and partial destruction of the structure of nanotubes with the dissolution of oxide inclusions and the formation of ulcerative pores. It should be noted that an increase in the test temperature leads to a sharper drop in the mass loss coefficient, which indicates an increase in the corrosion rate of nanotubes.

 

   

a)

b)

 

с)

 

Figure 7. Dynamics of mass loss due to corrosion: a) at a test temperature of 25°C; b) at a test temperature of 36°C; c) at a test temperature of 40°C.

 

According to the data of x-ray phase analysis, the degradation processes are associated with the formation of oxide and hydroxide phases in the structure of nanotubes, the presence of which leads not only to a deterioration in structural characteristics, but also with an increase in the test time to partial amorphization and degradation of nanotubes. Figure 8 presents data on changes in the concentration of impurity oxide inclusions in the structure of nanotubes.

 

   

a)

b)

 

с)

 

Figure 8. Graph of the change in the concentration of oxide phases in the structure of nanotubes: a) at a test temperature of 25°C; b) at a test temperature of 36°C; c) at a test temperature of 40°C

 

As is known, the degradation of iron-containing nanostructures occurs by the following mechanisms:

 

 

 

 

Degradation by nickel oxidation occurs as a result of the loss of electrons in the interaction with the medium and the transition to nickel (II) oxide:

Ni0 – 2e = Ni2+.

The nickel (II) oxide formed refers to bertollides with oxygen stoichiometry. With a large amount of oxygen in the structure, it transforms into nickel (III) oxide Ni2O3·H2O or NiOOH. In this case, nickel ions are completely oxidized:

Ni2+ – 1e = Ni3+.

But since the oxidation state of +3 is not characteristic of nickel, compounds with this valence are unstable; therefore, hydrated forms of nickel (II) oxide fall apart with the elimination of oxygen, which leads to the destruction of crystalline and chemical bonds.

 

The applicability of targeted modification of nanostructures by irradiation with gamma radiation in order to increase the rate of catalytic activity was evaluated using the reaction of the para-nitroaniline (PNA) reducing agent to para-phenyldiamine (PPD). The reaction was carried out by dispersing nanostructures (0.1 mg) in an aqueous solution of a mixture of PNA (10-4 М) and NaBH4 (6х10-2 М). Evaluation of the catalytic reduction ability was carried out on a UV spectrometer at room temperature in the time range from 3 to 30 minutes in increments of 3 minutes. Based on the obtained UV spectra, the reaction rate constant was calculated. Obtaining UV spectra, as well as assessing changes in the catalytic activity of the studied structures, was carried out by the in situ method. Samples in a volume of 10 μl were taken after 3 minutes of measurements and placed in a spectrometer to record UV spectra, based on which the reaction rate constant was calculated. 

Author Response File: Author Response.doc

Reviewer 2 Report

This manuscript uses a handful of assays to quantify effects of 1-5 MeV gamma-ray exposures on polycrystalline FeNi nanotube properties.  The basic finding is that the gamma-ray exposures lead to measurable effects on the polycrystalline structure, which in turn effects degradation rates of the structures and catalytic activity associated with the structures.  These basic findings are believable from the presented material.  However, there are numerous omissions in the manuscript that concern data collection and analysis, and which are important to the quantitative findings in the manuscript.  Without improvements to the manuscript, I cannot be confident in the quantitative results.  Some  of my concerns are described below.

I do not know from the manuscript how the gamma-ray dose was estimated.  I would like to have  sense of how accurate the dose estimation is.

The geometry of the x-ray diffraction measurements is not explained.  I would like to know the direction of the x-ray beam with respect to the axis of the nanotubes.

I would like to understand how texture analysis can be done faithfully from only one sample geometry.  There are many reasons for why selected Bragg peaks could vanish upon structural modification of an anisotropic sample, and the singular (unknown) diffraction geometry presented here does not allow me to conclude which one it is.

Figure 1d is redundant with figure 2.  Also, it looks like all negative values in the intensities have been removed - they should be included in the plots.

There is some discussion surrounding the change in the shape of the diffraction maxima, but there is no analysis of this shown in the figures, table, or described in the text.  It looks like the 111 line narrows with increased dose but it is hard to see a trend from the way the profiles are presented.  I suggest an overlay of the profiles and/or a plot or tabulation of peak widths.

Figure 3 shows a plot of density against dose.  There is no description of how density was estimated from the data.  I would like to know how density is defined.  The plot is labeled “microstrain density” but the text only refers to “density” which adds to my confusion.

I do not know how the “degree of structural ordering” was calculated in table 1.  I would like to see a clear definition of this, and how it was quantified. 

The space group column can be removed from table 1, since it is the same for all rows.  I suggest adding peak widths to this table, and perhaps the ratio of the 111 to the 200 peak.

Near line 200, there is discussion about how FeNi structures are “… the most promising candidates…” for targeted drug delivery.  I looked at reference 51, which is cited in the context of this statement, but I did not find the word “drug” anywhere in the article.  I am not expert in FeNi nanostructures, but it is evident that the literature has been cited appropriately in this manuscript.

Figure 8 shows a plot of reaction rate constants for different gamma-ray doses, but there is no explanation of how the rates were calculated from the spectra in figure 7.  I would like to see at least a brief description of how the rates were calculated (e.g. SVD analysis followed by a fit to 2-state model).  A plot of the populations over time should be shown alongside the spectra.  Also, I think the rate constant should have inverse-time units in figure 8.

The above are some examples of how the authors have not adequately explained important aspects of their measurements and analysis.  I believe the work is meaningful, and that others who work on FeNi nano-structures will be interested, but the present state of the paper only allows me to conclude one thing: gamma rays effect the structure of FeNi nanosctructures.  The quantitative aspects are questionable in my view, and the authors should make significant improvements to strengthen their presentation.

Author Response

 

The radiation dose was controlled by two methods: dose calculation taking into account the structural features of the nanostructures, their geometry and length, and also using film detectors that were placed near the samples and behind the film samples to the nanotubes in order to determine the dose accumulated during the irradiation. Placing film detectors in two places made it possible to compare the accumulated doses of both the nanostructures themselves and the detector without them.

The crystallographic characteristics of nanostructures before and after directed modification were studied using a D8 Advance Eco powder X-ray diffractometer (Bruker, Karlsruhe, Germany). Conditions for recording diffractograms: 2θ = 35–80°, step 0.01°, Bragg-Brentano geometry, spectrum acquisition time 3 s, sample table rotation speed 20 rpm, X-ray radiation Cu-Kα, λ = 1.54 Å. For X-ray diffraction, all samples before and after modification were in a polymer matrix, which was mounted on the sample holder perpendicular to the x-ray beam. The rotation of the sample table was carried out to minimize the effect of the difference in textures at different shooting angles.

 

Figure 2. X-ray diffraction patterns of the studied samples before and after irradiation

 

The degree of crystallinity or structural ordering was determined by approximating the diffraction lines by the required number of pseudo-Voigt symmetric functions, and also by measuring the width of the recorded lines at half their height (FWHM). The pseudo-Voigt functions used to approximate the X-ray peak profile in the diffractogram (3) [43,44]:

 

,

(3)

 

where x is the variable corresponding to the angle of reflection 2θ; x0 - sets the position of the maximum of the function; η is the specific fraction of the Lorentz function; A is the normalizing factor; bG and bL are the parameters of the Gauss functions G(x, x0, bG) and Lorentz L(x, x0, bL).

An analysis of the angular dependence of physical broadening allows one to evaluate the influence of factors such as the size effect associated with crushing or recrystallization of crystallites as a result of external influences, and a distorting factor depending on the degree of deformation of the crystal structure and its change as a result of external influences. To assess the impact, the Williamson-Hall method was used, which is based on relation (4) [45]:

,

,

,

(4)

where β is physical broadening of the diffraction maximum, λ is the x-ray wavelength (1.54 Å), D is the crystallite size, θ is the Bragg diffraction angle, ε is the magnitude of microstresses in the grating.

Microstresses were estimated based on an analysis of the displacement of diffraction peaks calculated according to equation (5):

,

(5)

where dpristine and dirr are the interplanar distances before and after irradiation.

 

This column has been deleted. Diffraction line width data added.

Sample

Lattice parameter, Å

Interplanar distance

Degree of structural ordering

Full width at half maximum of peak (111) (FWHM)

The ratio of the 111 to the 200 peak

(Intensity (111)/Intensity (200))

Initial

3.5492±0.0013

2.04781

86.4

1.494

6.11

100 kGy

3.5415±0.0016

2.04318

87.3

1.246

6.15

200 kGy

3.5381±0.0021

2.04087

91.1

0.917

6.21

300 kGy

3.5367±0.0015

2.04085

93.5

0.820

10.56

400 kGy

3.5353±0.0011

2.04071

94.1

0.812

16.32

500 kGy

3.5221±0.0017

2.03632

94.5

0.745

21.74

The density of the studied nanostructures was calculated using the formula (6):

 

(6)

where V0 is the volume of the crystal cell, Z is the number of atoms in the crystal cell, A is the atomic weight of atoms.

The degree of disorder of the crystal lattice or the so-called integral porosity of the samples under study was found according to formula (7):

 

(7)

where р0 is the density of the reference sample.

 

Corrected, the link is replaced with the corresponding one.

Figures 10a-b show the dynamics of changes in the intensity of spectral maxima characteristic of PNA and PPD compounds. As can be seen from the data presented, complete recovery is observed only for modified samples with doses of 300 and 500 kGy, while for unirradiated samples and irradiated with a dose of 100 kGy, the recovery reaction does not end in the allotted time for testing.

   

a)

b)

   

с)

d)

 

Figure 10. Graphs of the dependence of the changes in CI/C0 (a) and LnCI/C0 (b) reflecting the catalytic activity of the studied nanostructures; c) PNA-PPD reaction scheme; d) The graph of the change in the rate of reaction constant (dots indicate experimental data; red lines indicate approximation of the results)

 

As can be seen from the data presented, for the initial nanostructures, a partial restoration of the PNA–PPD reaction is observed over the entire cycle time (see the diagram in Figure 10c), which indicates a low rate of catalytic activity. In contrast to the initial nanostructures, the modified nanostructures show a complete recovery of the PNA–PPD reaction, as evidenced by an increase in the reduction rate (see the data in Figure 10d). The value of the reaction rate constant was calculated on the basis of data on the changes in the concentration of PNA and PPD compounds, which were estimated from the intensity values of the corresponding peaks in the spectra. As a result of the studies, it was found that an increase in the degree of ordering of the crystal structure, as well as a decrease in defective inclusions, leads to a sharp increase in the productivity of nanostructures for the PNA–PPD reduction reaction, which indicates the promise of using gamma radiation not only to increase corrosion resistance, but also catalytic activity nanostructures.

 

Author Response File: Author Response.doc

Round 2

Reviewer 1 Report

While most of the comments are addressed satisfactorily and the manuscript has significantly improved, the description and interpretation of the XRD experiments has now become even more confusing.

 

-- The authors describe three different methods/formulas to extract crystallite size from XRD patterns (Scherrer-Formula, Williamson-Hall, and Rietveld refinement). Only the last two are able to yield size and micro-strain, so I don't understand what the Scherrer-Formula is used for?

 

-- Regarding Williamson-Hall, and Rietveld refinement, yes they can be used to extract size and micro-strain; however, both methods relay on at least to peaks to be able to separate the two contributions. In the first version there are no (200) or (220) peaks visible for the samples irradiated to more then 300 kGy, but in the revised version they are visible. How is this possible? Why dose the background look different in the two versions?

 

-- How is size and micro-strain determined in Fig. 3, with Williamson-Hall or Rietveld refinement? What are the errors to size and micro-strain (error bars)?

 

-- Equation (4) should read macro-strain not micro-strain.

 

-- check line 108

 

 

Author Response

The authors are grateful for the comments made and will try to clarify the description of the methods.

 

1. The Rietveld method was applied to determine the phase composition of the studied nanostructures, as well as to determine the lattice parameters.

2. The Williamson-Hall method allowed us to estimate the contributions of dimensional factors and distortion of diffraction peaks. In this case, both crystallite size and microstress were calculated using this method.

3. The use of the Scherrer formula was used to estimate crystallite sizes without taking into account distortions of diffraction lines. Moreover, the use of this formula showed that the crystallite size is comparable to the sizes determined by the Williamson-Hall method. Also in the latest edition, the Scherrer equation is removed from the text of the article.

Regarding the representation of diffraction patterns, according to the remark of the previous reviewer, the representation of diffraction patterns was changed to represent them without subtracting the background, as well as smoothing, which led to a more pronounced manifestation of the entire diffraction pattern, as well as the reflection of weak reflections (200) and (220) for irradiated samples with doses above 300 kGy. The latest edition of the article presents diffraction patterns without the use of filters for smoothing and subtracting background lines, which made it possible to display all changes in the diffraction pattern. The authors also presented calculations of texture coefficients of the studied structures before and after irradiation.

 

The refinement of sizes and microstresses was carried out using the Williamson-Hall method.

Macrostrain were estimated based on an analysis of the displacement of diffraction peaks calculated according to equation (3):

,

(3)

Corrected.

Author Response File: Author Response.doc

Reviewer 2 Report

I am glad to see that the authors have included equations that explain how they analyzed the diffraction data. Most of this is fine, but I am still very confused about the diffraction geometry and the "texture" analysis. The authors state that "the rotation of the sample table was carried out to minimize the effect of the difference in textures at different shooting angles." Later in the manuscript, on line 216, the authors state that "A change in the intensities of the diffraction peaks indicates a reorientation of the texture, as well as a change in the orientation of the crystallites." These two statements are contradictory in my view. How did the authors infer orientational information from an isotropic sample?

Regarding dose: The authors wrote that they calculated dose while taking into account the geometry and length of the nanostructures. They gave no details, so I suggest that they either remove this statement or provide sufficient information for the reader to understand what they did. I do not think it is a simple matter to include nanostructures in dose calculations since radiation damage processes are very complicated (a detailed calculation would warrant a separate manuscript - it is certainly not necessary for this manuscript). The authors also stated that they measured the dose by comparing accumulated doses on film detectors with and without the sample in place, which seems fine. It would be helpful to have an estimate of the errors in this measurement.

I think that the statement regarding "texture orientation" needs to be fixed before the manuscript is published. Otherwise, it is probably suitable for publication in this journal.

Author Response

The authors are grateful for the comment provided.

 

Regarding these comments.

 

The change in the orientation of crystallites was evaluated by assessing the change in the intensity of diffraction reflections, and determining the texture coefficients, the values ​​of which show a distinguished direction of orientation. In this case, the use of the rotation of the sample in measuring the diffraction pattern allows us to remove questions about the azimuthal variation of the intensity with respect to the position of the angle φ. In this shooting mode, the intensities of the diffraction peaks are isotropic relative to the position of the angle φ, but their ratio allows us to estimate the contribution to the preferred orientation of the crystallites.

Moreover, for irradiated samples, a decrease in the intensity of (200) and (220) diffraction reflections is observed while maintaining isotropy with respect to φ, which leads to a change in the values of texture coefficients, the values of which are presented in Table 2, reflecting the preferred orientation of crystallites in the structure of nanotubes. Texture coefficients were calculated using the Harris formula (6):

 

,

(6)

 

where I(hkl) – the experimentally obtained relative intensity; I0(hkl) – relative intensity corresponding to a given plane according to the PDF-2 database; n – the number of planes.

 

Table 2. Texture coefficient data

 

TChkl

111

200

220

Initial

1.875*

0.643

0.482

100 kGy

1.942

0.743

0.315

200 kGy

1.964

0.564

0.472

300 kGy

2.053

0.564

0.383

400 kGy

2.153

0.531

0.316

500 kGy

2.534

0.253

0.213

* Coefficients greater than one are shown in italics

 

 

The authors are grateful for the comments made and agree with the reviewer that the radiation dose calculations require separate work. In this regard, the description of dose measurements is corrected and presented in the following form.

 

The radiation dose was controlled by using film detectors that were placed near the samples and behind the film samples to the nanotubes in order to determine the dose accumulated during the irradiation. Placing film detectors in two places made it possible to compare the accumulated doses of both the nanostructures themselves and the detector without them.  The measurement error of the dose was not more than 2-3%.

Author Response File: Author Response.doc

Round 3

Reviewer 1 Report

All comments have been addressed and I support publication.

 

Back to TopTop