In so-called advanced nuclear fuels, it has been shown that by adding small amounts of Cr-dopant to the UO2 matrix the grain size is increased, resulting in a longer pathway for fission gasses to migrate and escape the fuel matrix, thus reducing pellet swelling during reactor operations. Other favourable properties are promoted upon Cr doping, including improved plasticity and reduced pellet-cladding interactions, all of which reduce the risk of fuel failure1,2,3,4,5. The introduction of Mn as a dopant in UO2 fuels has been hypothesised to offer an improved alternative to commercially available Cr-doped UO2 fuel6, warranting further study of the crystal chemistry of Mn-doped UO2.

Atomistic modelling of the Mn-doped UO2 crystal lattice and defect formation during heat treatment suggested that grain growth is enhanced by the formation of U vacancy defects (Uv) that aid diffusion during the sintering process7. These defects are predicted to increase in concentration with the presence of dopants when compared to undoped UO2. Through atomistic modelling, the Uv concentration was shown to be up to five times greater in Mn-doped UO2 than Cr-doped UO2, strongly suggesting that Mn-doped UO2 should exhibit larger grains and, therefore, enhanced in-reactor properties with regards to fission gas migration. Experimental investigation of the synthesis of Mn-doped UO2 via a sol–gel method found grain size enhancement of up to 80 μm in samples when compared to UO2, with a measured Mn content of 490 ppm. In comparison, grains of ~ 50 μm were achieved in 350 ppm Cr-doped UO2 when both were sintered at 1700 °C for 6 h under a reducing Ar-4%H2 atmosphere8.

Grain size growth has also been postulated to be the result of the formation of a liquid phase in Mn-doped UO2, where MnO powder was dry mixed with an additional dopant, Al2O3, to form a 1000 ppm Mn/Al-doped UO29. When sintered at 1860 °C, which is close to the melting temperature of MnO (1945 °C), a maximum grain size of ~ 50 μm was achieved, attributed to a MnO-Al2O3 binary liquid phase. The mechanism of grain growth in Mn-doped UO2 is, therefore, unclear and further investigation into dopant concentration and solubility of Mn in the UO2 lattice is warranted to optimise fuel fabrication routes.

Additionally, post-operation of a nuclear reactor, the safe storage and disposal of spent fuel is dependent on the long-term durability of the UO2 matrix, most prominently when exposed to geological disposal conditions10,11. Matrix oxidation converts the relatively insoluble U(4+)O2 to soluble U(6+)O22+ and, as such, under geological disposal conditions, may result in radionuclide release into groundwater. Therefore, any alteration to UO2 that may result in oxidation will undoubtedly influence dissolution behaviour and so understanding how dopants affect these properties is crucial to ensuring safe disposal of radioactive waste. For example, it was recently found that the addition of Cr to UO2, reduced the dissolution rate of U, through a galvanic coupling effect between Cr2+ and U6+ 12. A thorough understanding of the crystal chemistry of Mn-doped UO2 in comparison to undoped UO2, including an investigation of defect formation, is thus required to fully underpin predictions of spent fuel dissolution behaviour over the geological disposal period (100,000 + years).

While dry-synthesis and sol–gel methods have been demonstrated for the synthesis of Mn-doped UO2, this work improves upon these processes using an alternative wet chemical co-precipitation synthesis route. To aid a detailed study of the local structure of Mn in UO2, the samples were sintered in a consistent reducing environment. Analysis of Mn-doped UO2, prior to, and after sintering, was performed by X-ray absorption spectroscopy (XAS), X-ray diffraction (XRD), and Raman spectroscopy. The effects of sintering in a reducing environment on the final microstructure and crystal chemistry is also discussed, completing an in-depth experimental study of Mn-doped UO2 crystal chemistry and providing a key foundation for future investigations of the use of Mn-doped UO2 as an advanced nuclear fuel.

Results & discussion

Development of Mn-doped UO2 synthesis and fabrication method

A nitrate co-precipitation method was developed to prepare Mn-doped UO2 with an initial target concentration of 1200 ppm Mn, in which nitrates of Mn (Mn(NO3)3·4H2O) and U (UO2(NO3)2·6H2O) were used. Concentrated NH4OH was added to this mixture in 0.1 mL instalments until a pH of 7, 9, 10 or 11 was reached and precipitation of a mixed compound was achieved. Analysis of the supernatants via Inductively Coupled Plasma-Optical Emission Spectroscopy indicated the success of co-precipitation of Mn for each pH, where ~ 99.00 ± 1.00% U was precipitated for all pH values. The Mn incorporation in the resulting solid upon precipitation varied, with 84.98 ± 0.01%, 99.96 ± 0.03%, 99.80 ± 0.03% and 99.9 ± 0.6% Mn incorporated for pH 7, 9, 10 and 11, respectively (Fig. 1a). This indicates that a pH of ≥ 9 is required to fully incorporate Mn within co-precipitated product. Consideration of both the supernatant analysis and a complete nitric acid digest of the solids confirmed that the optimal synthesis was achieved at pH 10 and, as such, the remainder of the samples discussed herein were fabricated using material precipitated at pH 10.

Figure 1
figure 1

Initial investigation of an optimal synthesis and fabrication route for Mn-doped UO2. (a) Complete digest of target 1200 ppm Mn-doped UO2 synthesised at pH 7, 9, 10 and 11; (b) Thermogravimetric analysis of precursor material synthesised at pH 10 and; (c) X-ray diffraction patterns of precursor and calcined material (pH 10).

An optimal calcination temperature was determined by Thermogravimetric analysis of the precipitated precursor material (Fig. 1b), where mass losses due to the removal of water (100–190 °C)13, U/Mn and ammonia nitrate from synthesis (200–300 °C)14,15, and the decomposition of nitrate oxides formed during heating (300–525 °C)15, occurred up to a temperature of ~ 600 °C. The DTA curve shows endothermic peaks, A and C, relating to water loss and decomposition of nitrate oxides, respectively, and an exothermic peak, B, attributed to the formation nitrate oxides during heating. A calcination temperature of 750 °C was selected and conversion from precursor material (identified as 3UO3·2NH3·4H2O) to UO2 after heat treatment for 4 h under a reducing atmosphere (95% N2/5% H2) was confirmed via XRD of the calcined powder (Fig. 1c).

Crystal chemistry of Mn-doped UO2

Mn K-edge X-ray absorption near edge spectroscopy (XANES) analysis of Mn-doped UO2 showed a unique coordination environment when compared to standards of known speciation (Fig. 2a). Of the different Mn standards measured, the E0 position (white line, taken from the peak of the first derivative) of Mn-doped UO2 (6551 ± 0.3 eV) was closest to that of the Mn2+ standards (6550–6552 ± 0.6 eV) (Table 1); however, spectrally, each Mn2+ standard was clearly different to the samples, suggesting that Mn in UO2 does not possess a local environment similar to those of the Mn standards.

Figure 2
figure 2

Mn K-edge XANES spectra of Mn doped UO2 compared to known standards. (a) complete spectra and; (b) pre-edge region. Standards are depicted in black, with calcined samples in blue and sintered samples in orange.

Table 1 XANES analysis for Mn-doped UO2 samples and Mn standards.

Analysis of the pre-edge feature in the Mn K-edge XANES (Fig. 2b and Supplementary Material Fig. 1) can be used to probe the local symmetry environment16,17. This pre-edge feature arises from the forbidden 1s to 3d electronic transition and becomes more intense when there is stronger mixing of the d and p orbitals resulting from a lack of local symmetry in the Mn environment18. Here, the centroid position and integrated area of the pre-edge feature were similar for all Mn-doped UO2 samples (6540.42–6540.64 eV and 0.0490–0.0994, respectively) (Fig. 2b and Table 1). These values are in accordance with those determined for standards that contain Mn2+ in an octahedral local symmetry environment (6540.36–6540.83 eV and 0.055–0.285 for the centroid position and integrated area, respectively). The values for Mn-doped UO2 were also in agreement with literature values for (Mn2+,Fe)3Al2Si3O12 (6540.66 eV, 0.0745)17, which exhibits a cubic Mn local environment; however, the same study reports similar values for octahedrally-coordinated Mn2+O (6540.62 eV, 0.0745). This highlights that differentiation in the local symmetry of Mn2+ by pre-edge analysis is not simple and that additional Extended X-ray Absorption Fine Structure (EXAFS) analysis is required to confirm the coordination environment.

Since there were no changes in the XANES spectral features with increasing Mn content, nor between calcined oxide powder and sintered samples (Fig. 2), the 1200 ppm and 2400 ppm Mn-doped UO2 calcined powder samples were selected as representative of the local structure of Mn in UO2 for EXAFS fitting of the Mn K-edge. In agreement with the models of Mn-doping in UO2 developed by Cooper et al.7, the fitted data for both concentrations (Fig. 3, Table 2; corresponding data for the sintered material is given in Supplementary Material Fig. 2), indicated that the first O nearest neighbour environment of Mn is split between 6 O atoms at 2.27–2.29 ± 0.01 Å and 2 O atoms at 2.55–2.57 ± 0.03 Å in an eightfold coordination environment, consistent with direct substitution of Mn onto the U4+ site, or onto an interstitial site7. Slight contraction in the first Mn–O distance, when compared to 2.36 ± 0.02 Å U–O distance, is to be expected if Mn is substituted onto the U4+ site, due to the smaller cation size of Mn2+ (0.96 Å) compared to U4+ (1.00 Å). This contraction also creates a small amount of distortion in the local Mn environment, which is likely the reason for the first coordination shell splitting into two different O distances.

Figure 3
figure 3

Mn K-edge spectra (black lines) and EXAFS model fits (red lines) for nominally doped 1200 ppm and 2400 ppm Mn calcined powder samples. (a) k3-weighted EXAFS; and (b) Fourier transform of the k3-weighted EXAFS, using a Hanning window function.

Table 2 Fitting parameters and results for 1200 ppm and 2400 ppm Mn-doped UO2 oxide powders.

Substitution of Mn2+ on the U4+ lattice site is further supported by a decrease in the lattice parameter with increasing Mn content, consistent with Vegard’s Law (Fig. 4). Both calcined powder and sintered samples maintained the fluorite crystal structure of UO2 (Supplementary Material Fig. 3). Beyond the first coordination shell, a further 5 U atoms were fit at distance of 3.78 ± 0.01 Å (Table 2). The decreased coordination from the expected 12 U atoms at 3.85 ± 0.06 Å in crystalline UO2 may be attributed to a combination of lattice distortion from Mn-doping and of the limited data range available (2–10.5 Å−1). The combination of these two factors makes delineation of the distorted U-U backscatterers challenging; extension of the data range, not practically possible in this experiment due to the low Mn concentrations within a highly absorbing UO2 matrix, could resolve this.

Figure 4
figure 4

Lattice parameter values for calcined powder and sintered material, as determined by XRD. Errors are the standard deviation of triplicate analysis.

The bond valence sum (BVS) was calculated and used to indicate the oxidation state. Values of 1.80 and 1.88 were calculated for the 1200 ppm and 2400 ppm Mn-doped UO2 samples, respectively, supporting the presence of Mn2+ (Table 2). This is in agreement with recently published studies of the coordination of Cr doped in the UO2 structure, which was found as Cr2+ in calcined material19.

A charge balance mechanism is required in the substitution of Mn2+ on the U4+ site, provided by the formation of positive defects, U5+ and/or oxygen vacancies (Ov). Experimental evidence for U5+ formation can be determined using the High Energy Resolution Fluorescence Detected (HERFD) XANES method, with data acquired at the U M4-edge; however, given that the doping level of Mn was low (300–2400 ppm), the amount of corresponding U5+ (up to 0.24%) is at, or below, the limits of detection of the technique. Nevertheless, principal component analysis of the data series revealed that, for the calcined powder, two components, attributed to U5+ (closely matching a standard of CrUO4) and U4+ (UO2) oxidation states, were required to accurately reconstruct each of the samples in the series (Fig. 5a and Supplementary Material Fig. 4).

Figure 5
figure 5

U M4-edge HERFD XANES of Mn-doped UO2 materials. (a) calcined material; (b) quantification of the U4+/5+ ratio in calcined powder determined from fitting of U M4-edge HERFD XANES data in (a); and (c) sintered material.

Iterative target transformation factor analysis was used to indicate the approximate fraction of U4+ and U5+ in each of the oxide powders (Fig. 5b), with results indicating ~ 21–26 ± 2% U5+ present in Mn-doped UO2 (and, therefore, ~ 74–79 ± 1% U4+). Given the arguments above, the presence of U5+ seems unlikely to arise from a charge compensation mechanism, but from oxidation of the samples. Despite being exposed to air post-synthesis for the same period of time as the calcined samples, the sintered materials did not show any strong evidence for the presence of U5+ (Fig. 5c).

The formation of Ov defects was observed through Raman spectroscopy (Fig. 6a). The main T2g peak (~ 445 cm−1) of UO2 and the defect band (~ 500–700 cm−1), indicative of lattice distortion due to defect formation, was present for all samples. The relative intensity of individual peak contributions of the defect bands, where U1 is attributed to Ov, U2 to the LO phonon mode and U3 to Oi, was compared to the relative intensity of the T2g peak20 (Fig. 6b). There were minor fluctuations in the relative contributions as the concentration of Mn was increased; however, Ov were always found to be present (Fig. 6c), confirming their formation as a charge balance mechanism for Mn doping in UO2.

Figure 6
figure 6

Raman analysis of Mn-doped UO2 calcined powders. (a) Raman spectra; (b) deconvolution of the 2400 ppm Mn-doped UO2 spectra into U1, U2 and U3 peaks; and (c) defect content realised by the area ratio of the defect peak to the T2g peak.

As noted previously, there were no significant differences in the crystal structure of sintered Mn-UO2 when compared to the calcined oxide powders, indicated by consistency in spectral features observed in the Mn K-edge XANES (Fig. 2) and the Fourier Transform of the EXAFS region (Fig. 3 and Supplementary Material Fig. 2). The reduction in UO2 lattice parameter, due to Mn2+ substitution on the U4+ site, was maintained; although, the absolute values of lattice parameter were 0.002–0.004 Å greater in the sintered samples than the calcined samples (Fig. 4). This could be the result of volatilisation of Mn during sintering (see below) and/or reduction of U5+ (0.84 Å) to U4+ (1.00 Å) imposed by the reducing high temperature environment, as observed by HERFD-XANES (Fig. 5c), in which no U5+ was present.

Influence of reducing sintering conditions on the microstructure of Mn-doped UO2.

When comparing the Mn concentrations of calcined oxide powders with those of materials sintered in a reducing (95% N2(g) / 5% H2(g)) environment, it is evident that significant volatilisation of Mn occurred (Fig. 7a)8. For UO2 doped with 300 ppm Mn, the loss was ~ 30%, while for the remaining concentrations (600, 1200 and 2400 ppm), it was ~ 80%.

Figure 7
figure 7

Microstructural evolution of undoped and Mn-doped UO2 sintered in a reducing environment. (a) the measured Mn content in Mn-doped UO2 calcined powder and sintered pellets; and SEM images and average grain size of ~ 500 grains measured across the sample surface in (b) undoped UO2 and (cf) Mn-doped UO2 of increasing Mn content.

No significant influence on the grain size was observed when the average measurements of ~ 500 grains were considered in comparison to undoped UO2 (Fig. 7b–e); however, there was a marginal increase with increasing dopant concentration. An average grain size of 13.9–14.8 ± 0.5 μm was recorded, in comparison with undoped UO2, sintered under the same conditions, which had an average grain size of 14.0 ± 0.2 μm. This is in contrast to previous studies8 and modelling predictions6, most likely due to the nature of the sintering atmosphere, which contained no O2.

Studies of commercially available Cr-doped UO2 have shown that the sintering atmosphere is the most important factor contributing to grain growth, with the addition of small quantities of oxygen thought to promote the formation of a ‘CrO’ eutectic phase. At sintering temperatures (> 1675 °C), this phase is liquid and aids grain growth, evidenced as precipitates of Cr2O3 formed upon cooling21,22,23,24. The effects of a eutectic phase have been reported for Mn/Al doped UO2 sintered in a H2 atmosphere at 1860 °C, where a MnO/Al2O3 phase was observed by EDX mapping10. However, contrary to Cr-doped UO2, the addition of oxygen to the sintering atmosphere in Mn/Al doped UO2 resulted in a smaller grainsize, attributed to increased solubility of MnO in the UO2 matrix10.

In the present study, under fully reducing conditions, much of the Mn2+ incorporated into the UO2 was volatilised, and no precipitates were observed, which indicates that the solubility limit of Mn in UO2 was not reached and that an equivalent MnO liquid phase was not likely formed. Moreover, no significant increase in the grain growth was observed with increasing Mn concentration (Fig. 7). One reason for this is that the temperature of sintering (1700 °C) was below the melting point of MnO2 (1945 °C), therefore, the formation of any possible eutectic phase was restricted. It has been previously shown, in Cr-doped UO2, that, if the sintering atmosphere required for the eutectic phase formation is not met, the presence of precipitates at grain boundaries will have the opposite effect and grain growth will be hindered by grain boundary “pinning”21; however, this reasoning is excluded here due to the absence of precipitates in the Mn-doped UO2 samples.

Despite the observation that the average grain size did not significantly increase upon increased levels of Mn doping, when compared with the undoped UO2, individual grain sizes of up to 50 μm were observed for all Mn-doped UO2 samples studied (Fig. 7b–e). This ‘localised’ grain growth has previously been reported for Mn-doped UO2 prepared via sol–gel synthesis and sintered under reducing conditions (95% Ar(g) / 5% H2(g), 6h, 1700 °C)8. Such behaviour may be due to the increased diffusivity of U atoms predicted by atomistic models of the defect concentrations7. Further work to underpin the solubility limit of Mn in UO2, and to develop an understanding of the influence of sintering conditions on Mn–UO2, is required to fully elucidate the mechanism of grain size growth.

These results have given insight into the crystallographic basis for interpretation of grain growth as well as oxidation and dissolution behaviours in Mn-doped UO2 and support development for future use as accident tolerant fuel.

Methods

Sample preparation

A wet synthesis method was developed using uranium nitrate hexahydrate (UO2(NO3)2·6H2O, (The British Drug House (BDH). B.D.H Laboratory Chemicals Division, > 98%, 0.3 mol L−1). This was mixed with manganese(II) nitrate tetrahydrate (Mn(NO3)3·4H2O, Sigma Aldrich, 99.99%, 1.6 mol L−1) in the proportion required to give the desired concentration of dopant. An initial synthesis study was performed in which a number of samples were assessed to determine the optimal pH and calcination temperature for fabrication of Mn-doped UO2. The final fabrication route is shown in Fig. 8 and is detailed as follows. Concentrated ammonium hydroxide, NH4OH (5 mol L−1), was added until a pH of 10 was reached. The resultant precipitate was vacuum filtered, washed in ultra-high quality (18 MΩ cm) water, and dried for 24 h at 90 °C. Thermogravimetric analysis was carried out on precursor material using a Netzsch TG 449 F3 Jupiter instrument coupled with a 64-channel QMS 403 D Aeolos mass spectrometer. Samples were heated to 1000 °C at a rate of 10 °C min−1 under a constant Ar(g) flow. Analysis of the mass loss over time determined that a calcination temperature of 750 °C for 4 h under a reducing (95% N2(g) / 5% H2(g)) atmosphere was sufficient to convert the precursor material to oxide. Successful co-precipitation of Mn with U oxide was measured by inductively coupled plasma optical emission spectroscopy (ICP-OES) analysis of the supernatant, where > 99.9% precipitation was observed for both U and Mn for samples precipitated at pH 10.

Figure 8
figure 8

Optimised synthesis and fabrication route for Mn- doped UO2.

Homogenised oxide powders, prepared via milling at 35 Hz for 15 min, were uniaxially pressed at 2.5 tonnes into 6 mm green pellets and sintered at 1700 °C for 8 h in a reducing atmosphere (95% N2(g) / 5% H2(g)). Complete digest of precursor and calcined material, as well as sintered pellets (crushed to powder using a mortar and pestle), was carried out to assess the Mn content. In this method, 20 mg of sample was completely dissolved in 2M nitric acid (ultrapure HNO3) at 90 °C, with the aid of a magnetic stirrer, and an aliquot was removed and analysed by ICP-OES (Thermofisher Scientific iCAPDuo6300) for Mn and U concentration. Samples were dissolved in triplicate and an average measurement of Mn, with errors of 1 standard deviation, for each sample reported.

Characterisation

Secondary electron images of thermally etched pellet sample surfaces were taken using a Hitachi TM3030 scanning electron microscope at an accelerating voltage of 15 kV. Samples were thermally etched at 80–90% sintering temperature under a reducing atmosphere (95% N2(g) / 5% H2(g)) and the revealed grains imaged at 500 × magnification. The Image J software was used to obtain the average grain size of ~ 500 grains in images taken from 5 different regions across the pellet surface.

A PANalytical Xpert3 diffractometer was used in reflection mode with a 45 keV/40 mA generator to characterise powder samples by XRD (XRD). Data were collected between 5° and 100° 2θ with a step size of 0.013° and a step time of 40 s, a fixed slit size of 0.5 was used. LaB6 (20–30 wt%) was used as an internal standard for data alignment, corrected using the WinXPow software. Accurate determination of lattice parameters was carried out in the Topas software, using Le Bail refinements. To avoid oxidation, calcined powders were measured immediately post heat treatment (i.e. within 10 min of removal from the furnace) and sintered samples were crushed upon removal from the furnace, within a controlled inert atmosphere (N2(g)), and immediately measured to avoid oxidation.

Mn K-edge (6.539 keV) and U M4-edge (3.728 keV) X-ray Absorption Spectroscopy measurements were performed, at room temperature, on both calcined powder and sintered Mn-doped UO2. The I20-scanning beamline at the Diamond Light Source (DLS), UK, was used in fluorescence mode to measure the Mn K-edge using a 64-element Ge detector with Xspress4 signal processing and a beam size of 400 × 300 µm (FWHM). A high flux and excellent energy resolution is given by the wiggler-sourced I20-scanning beamline due to the use of a Si(111) four-bounce crystal monochromator, allowing the detection of low concentrations of Mn dopant in the heavily absorbing UO2 matrix.

Multiple scans were taken across a number of energy ranges to improve data quality. Energy steps of 5 eV were taken from 6439 to 6535 eV, 0.2 eV from 6536 to 6560 eV with a time step of 1 s step−1 in the XANES region for all samples. In the EXAFS region for calcined Mn-doped UO2 samples, an energy step of 0.04 Å−1 was taken from 6560 – 6980 eV with a time step of 1–6 s step−1. The spectral features in the sintered samples were identical to those of the calcined powers. A number of standards of known Mn valence state and coordination environment were measured in transmission mode, including a range of Mn2+ standards: Mn2+O; (Mn2+, Fe)(Ta, Nb)2O6; Mn2+SiO6; and Mn2+2SiO4.

The E0 (white line, taken from the peak of the first derivative) positions of the raw data were aligned to the E0 position of a standardised reference Mn foil that was measured simultaneously with Mn-doped UO2 samples. A cubic spline background subtraction and normalisation procedure25 was then carried out and multiple scans merged using the Athena software26. To isolate the pre-edge region for analysis, an arctangent function was applied across the 6535.0–6548.0 eV energy range to remove the photoelectric background absorption27 and three Gaussian peak functions used to model the pre-edge features (Supplementary Material Fig. 1). The energy position and normalised height were refined using the non-linear least-square fitting SOLVER function in Excel and the pre-edge centroid position assigned by taking the intensity weight average of each Gaussian fit.

To create models of the local environment realised in the EXAFS region, absorption data (eV) were converted to wavenumber, k (Å−1), and a Fourier Transform of the resulting k space was used for fitting. The possible scattering paths of the Mn central absorber atom to the surrounding atoms in the Mn-doped UO2 local environment were generated using a FEFF6 algorithm28 in Artemis26 where a UO2 CIF file was modified to include Mn as the central absorbing atom at the first U lattice position. The range of k (Å−1), r (Å), and the amplitude reduction factor (S02) were optimised for the data and the following parameters allowed to refine: degeneracy (N); fitted bond length (R) (Å); shift from Mn K-edge position (6.539 keV) (∆E0) and; the Debye–Waller factor (σ2). The F-test was applied to each fitted path, the result (α) indicated the confidence of adding the path to improve the fit (> 67% gives a confidence of 1σ, > 95% gives a confidence of 2σ)29. The bond valence sum (BVS) was calculated for the Mn–O coordination environment, which is an evaluation of bond distances in a coordination shell that are equal to the formal oxidation state of the cation absorber30.

HERFD XANES was carried out at the U M4-edge using the BM20 beamline at the European Synchrotron (ESRF), France31, with the use of an X-ray emission spectrometer32. Calcined powders were mixed with poly-ethylene glycol (PEG) and pressed into 6 mm pellets for analysis, while the as-sintered surfaces of pellets were measured. Raw U M4-edge data was pre-processed in Athena, as discussed above, and a principal component analysis was used to determine the possible number of spectral components, followed by Iterative Target Transformation Factor Analysis (ITTFA)33 to indicate the relative concentration of each component using U(4+)O2 and CrU(5+)O4 standards measured alongside Mn-doped UO2 at the BM20 beamline.

Raman measurements were performed using a Renishaw inVia Reflex confocal spectrometer equipped with a Leica DM2500 microscope. Calcined Mn-doped UO2 powders were pressed into pellets before analysis using a 514.5 nm green argon laser with 1800 lines mm−1 grating to acquire a spectral acquisition between 200 and 800 cm−1. An average of 10 points across the compacted green pellet surface of each pellet was taken, to confirm their reproducibility and homogeneity of the composition across the sample. Lorentz function fitting was used to obtain information of peaks attributed to defects (U1, U2 and U3) in the UO2 structure, relative to the main T2g band of UO2 at 445 cm−1. In addition to the U2 LO phonon band at ~ 574 cm−1, the U1 peak at ~ 527 cm−1 attributed to Ov defects, and U3 peak at ~ 634 cm−1 attributed to Oi defects, were identified19.