Next Article in Journal
Effects of Nano-SiO2 and Nano-CaCO3 on Mechanical Properties and Microstructure of Cement-Based Soil Stabilizer
Previous Article in Journal
Deposition of HfO2 by Remote Plasma ALD for High-Aspect-Ratio Trench Capacitors in DRAM
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Kondo-like Behavior in Lightly Gd-Doped Manganite CaMnO3

1
Institut za fiziku, Bijenička cesta 46, HR-10000 Zagreb, Croatia
2
“Vinča” Institute of Nuclear Sciences, National Institute of the Republic of Serbia, University of Belgrade, Mike Petrovića Alasa 12-14, 11351 Belgrade, Serbia
3
Faculty of Sciences and Mathematics, University of Priština in Kosovska Mitrovica, Lole Ribara 29, 38220 Kosovska Mitrovica, Serbia
4
Department of Bromatology, Faculty of Pharmacy, University of Belgrade, 11000 Belgrade, Serbia
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(11), 784; https://doi.org/10.3390/nano15110784
Submission received: 7 April 2025 / Revised: 7 May 2025 / Accepted: 21 May 2025 / Published: 23 May 2025
(This article belongs to the Topic Magnetic Nanoparticles and Thin Films)

Abstract

:
Manganese oxides (manganites) are among the most studied materials in condensed matter physics due to the famous colossal magnetoresistance and very rich phase diagrams characterized by strong competition between ferromagnetic (FM) metallic and antiferromagnetic (AFM) insulating phases. One of the key questions that remains open even after more than thirty years of intensive research is the exact conductivity mechanism in insulating as well as in metallic phases and its relation to the corresponding magnetic structure. In order to shed more light on this problem, here, we report magnetotransport measurements on sintered nanocrystalline samples of the very poorly explored manganites Ca 1 x Gd x MnO 3 with x = 0.05 and x = 0.10 , in the temperature range 2–300 K, and in magnetic fields up to 16 T. Our results indicate that both compounds at low temperatures exhibit metallic behavior with a peculiar resistivity upturn and a large negative magnetoresistance. We argue that such behavior is consistent with a Kondo-like scattering on Gd impurities coupled with the percolation of FM metallic regions within insulating AFM matrix.

Graphical Abstract

1. Introduction

In contrast to conventional materials, in many advanced materials there is strong coupling between charge, spin, lattice, and orbital degrees of freedom, which highly complicates the understanding of these materials but also leads to many exotic and applicable phenomena. One of the most prominent examples of such material is the family of manganese oxides (manganites) described by a general chemical formula R 1 x A x MnO 3 , where R is a rare-earth element or Bi and A is an alkaline-earth element or Pb. Significant large interest in manganites started in 1994 with the discovery of the famous colossal magnetoresistance (CMR), an enormous decrease in electrical resistivity ( ρ ) in an applied magnetic field (H), which close to the Curie temperature can reach values as high as 127,000 % [1], in contrast to the conventional materials, where the effect of H is usually a few %. Aside from CMR, manganites are also interesting because of their very rich phase diagrams in which the ferromagnetic (FM) phase is usually metallic, while the antiferromagnetic (AFM) phase is insulating [2,3,4,5].
Rich physics of manganites R 1 x A x MnO 3 stems from the mixed valence state Mn 3 + /Mn 4 + caused by the substitution of trivalent R 3 + with a divalent A 2 + . The Mn 3 + /Mn 4 + cations are octahedrally surrounded by O 2 anions within the distorted perovskite crystal structure, which leads to standard splitting of the atomic Mn 3d levels into high-energy e g and low-energy t 2 g orbitals. In the case of Mn 3 + , there is additional splitting caused by Jahn–Teller elongation distortion of the MnO 6 octahedron (Figure 1). Three valence electrons of Mn 4 + occupy the three low-energy t 2 g orbitals, while four valence electrons of Mn 3 + occupy the t 2 g orbitals and the lower of the two e g orbitals. Due to the strong Hund coupling, each orbital is occupied by a single electron, and the spins of all electrons are parallel to each other so that Mn 3 + and Mn 4 + have magnetic moments of 4 μ B and 3 μ B , respectively, where μ B is the Bohr magneton [5].
The strong Hund coupling is the reason why the most relevant hopping process in manganites is the hop of an e g electron from an Mn 3 + to a neighboring Mn 4 + cation. It was shown already in 1951 by Zener [6], and later by Anderson and Hasegawa [7] as well as by Kubo and Ohata [8], that the probability t of such a hop of the conduction e g electron depends on the angle θ between the magnetic moments of the corresponding Mn 3 + and Mn 4 + cations (created by core t 2 g electrons) t cos ( θ / 2 ) . This is the famous double exchange model that is crucial for understanding of the physics of manganites. According to this model, the maximum hopping probability occurs for the parallel arrangement of neighboring Mn magnetic moments ( θ = 0 ), which qualitatively explains why the FM phase in manganites is usually also metallic and why an external magnetic field drastically reduces the resistivity (the CMR effect). Conversely, the minimum (zero) hopping probability occurs for the antiparallel arrangement of neighboring Mn magnetic moments ( θ = π ), which qualitatively explains why the AFM phase in manganites is also insulating.
Aside from this double exchange interaction that favors the FM arrangement of neighboring Mn magnetic moments, in manganites, there is also the more conventional superexchange interaction between t 2 g and e g electrons that favors the AFM arrangement of the neighboring Mn magnetic moments [9,10,11,12]. Additionally, manganites are characterized also by a strong electron–phonon interactions, since each hop of a conduction e g electron is accompanied by a structural change in MnO 6 octahedra induced by the Jahn–Teller effect (see Figure 1), which prefers a polaronic insulating state [13,14,15,16,17,18]. Finally, there are Coulomb interactions among conduction e g electrons, which are believed to be responsible for tendencies towards a charge ordered (CO) insulating state [19,20,21,22,23,24,25].
All these different interactions in manganites are of comparable strength, which is precisely the reason for the aforementioned very rich phase diagrams. Moreover, the competition between different interactions in manganites is so strong that it leads to the so-called nanoscale phase separation, i.e., to the coexistence of nanoclusters of competing phases. The most important is the coexistence of FM and AFM nanoclusters, which is especially pronounced close to the doping x that separates the FM metallic and CO/AFM insulating ground states [2,5,26,27,28]. Such a phase-separated state is extremely sensitive to internal tuning parameters such as doping and R / A disorder, as well as to external tuning parameters such as temperature (T) and magnetic field. In the case of non-single-crystalline samples, the phase-separated state is also sensitive to the sample’s microstructure, i.e., to grain boundaries in polycrystalline samples [29,30,31,32,33,34,35,36] or to the film/substrate interface in thin film samples [37,38,39,40,41,42,43].
The nanoscale phase separation is believed to be key in understanding the physical properties of manganites, most notably the CMR effect and the transition between the FM metallic and CO/AFM insulating phase [2,5,28,43,44]. One of the most important unresolved problems in manganites is concerned with the conduction mechanism in this FM/AFM phase-separated state. The crucial question here is whether this separated state should be regarded as a mixture of metallic and insulating islands within a random resistor network percolation model, or as a magnetic disorder that additionally scatters the conduction e g electrons. Another important question is concerned with the conduction mechanism in high magnetic fields, when all FM clusters are oriented in the direction of the field. A final important issue is the role of other interactions and the structural disorder in the conduction mechanism of manganites.
In order to address these questions, here, we report our magnetotransport study on sintered nanocrystalline samples of very poorly explored manganites Ca 1 x Gd x MnO 3 in high magnetic fields up to 16 T and at low temperatures down to 2 K. We focus on the samples with low Gd doping levels x = 0.05 and x = 0.10 , because for these samples, previous magnetic measurements indicated a significant phase separation into FM regions within the AFM matrix [45]. Additionally, in these samples, there are no confounding effects related to charge ordering (which occurs at higher x values), the AFM phase has the simplest G-type structure, and complications related to the Jahn–Teller distortion of Mn 3 + are highly reduced [46,47], which makes them an excellent playground for the proposed study of the conduction mechanism in a FM/AFM phase-separated state. In contrast to common beliefs, our results imply that both studied samples x = 0.05 and x = 0.10 at low temperatures exhibit metallic rather than insulating behavior. We find that this metallic behavior features a peculiar resistivity upturn at the lowest temperatures that can be best fitted to the Kondo-like behavior. The evolution of the fitting parameters with magnetic field points towards a classic electrical percolation picture where the conducting component refers to the FM metallic regions and the non-conducting component refers to the insulating AFM matrix.

2. Materials and Methods

Nanopowders of Ca 1 x Gd x MnO 3 with x = 0.05 and x = 0.10 were prepared using a modified glycine nitrate procedure described in refs. [48,49]. The as-prepared nanopowders were then uniaxially cold pressed at a pressure of 50 MPa and further isostatically pressed at a pressure of 250 MPa. The final sintered nanocrystalline samples in the form of compacts were obtained via thermal pre-etching at temperatures of 1300 and 1400 for 15 min [45]. Both x = 0.05 and x = 0.10 sintered samples were reported to have a single phase with a perovskite crystal structure (orthorhombic space group P n m a , No. 62) and a chemical composition that is in good agreement with the nominal values. The grain size was found to be a few μ m , and the crystallite size was found to be around 30 nm.
Resistivity measurements were performed using the standard four-contact dc technique within a commercial Quantum Design Physical Property Measurement System (PPMS) in the temperature range of 2–300 K and in magnetic fields of up to 16 T. Typical sample dimensions for transport measurements were 6 × 3 × 1 mm 3 . The annular electric contacts were prepared by applying silver paste directly to the sample surface. The current was applied along the long axis of the sample, and the magnetic field was applied perpendicular to the current.

3. Results

The resistivity as a function of temperature for both Ca 1 x Gd x MnO 3 samples x = 0.05 and x = 0.10 is shown in Figure 2. For comparison, the T-dependence of resistivity for the parent compound CaMnO 3 is also shown [50], which, according to the conventional phase diagram, has an insulating ground state with a single magnetic order, which has the simplest G-type AFM structure [2,3]. As we can see, the Gd doping drastically reduces the resistivity of CaMnO 3 , especially at low T values, and more effectively for x = 0.10 than for x = 0.05 . Note that the resistivity derivative d ρ / d T for both samples remains < 0 , which is usually in the literature referred to as insulator-like behavior. However, as we will argue later, both samples with cooling exhibit an insulator-to-metal (IM) crossover, even at H = 0 , where d ρ / d T < 0 .
The effect of an external magnetic field on the resistivity of the Ca 1 x Gd x MnO 3 samples is shown in Figure 3. As we can see, there is a significant decrease in resistivity in the magnetic field at low T values, which for x = 0.10 in the maximum field of 16 T even leads to real metallic behavior ( d ρ / d T > 0 ) below ≈100 K. In other words, there is an IM crossover at ≈100 K where ρ ( T ) reaches a local maximum (Figure 3a). The IM crossover slightly shifts to lower temperatures with decreasing fields and evolves into a knee-like feature at low fields. A similar knee-like feature, although barely visible, is present also for x = 0.05 at all fields. Later, we will argue that this knee-like feature is also related to the IM crossover. Note that the metallic behavior in high fields for x = 0.10 exhibits a peculiar resistivity upturn at ≈20 K, the temperature at which the ρ ( T ) curve passes through a local minimum.
Let us now focus on the most interesting ρ ( T ) curve in the maximum field of 16 T for the Ca 1 x Gd x MnO 3 sample x = 0.10 . The complex T-dependence of that curve can be much more easily analyzed by taking the first derivative d ρ / d T , which is shown in Figure 4a. As can be seen, there are four characteristic temperatures in d ρ / d T . The sharp minimum at T N 120 K can be readily identified as the Néel temperature, i.e., as the phase transition to the G-type AFM ground state reported in the literature [46,47]. The zero point at T I M 100 K corresponds to the local resistivity maximum in Figure 3a, i.e., to the IM crossover. The wide maximum at T i n f 70 K corresponds to the inflection point of resistivity, and the zero point at T m i n 20 K corresponds to the local resistivity minimum (see Figure 3a), the physical meaning of which will be discussed later.
The evolution of d ρ / d T with magnetic field for the x = 0.10 sample is shown in the inset of Figure 4a. As can be seen, the d ρ / d T curves in the zero field and in the field 1 T are negative across the whole T-range, i.e., in contrast to the rest of the curves, they do not have the two zero points associated with T I M and T m i n . On the one hand, such behavior might indicate that the conduction mechanism at low fields is different than the conduction mechanism at high fields. On the other hand, a smooth field evolution of d ρ / d T , specifically of the wide maximum at T i n f , the shape of which does not change with the field, points towards the same conduction mechanism at all fields, which will be further corroborated later.
The minimum at T N and the maximum at T i n f are present also in the d ρ / d T curve for the Ca 1 x Gd x MnO 3 sample x = 0.05 in 16 T (Figure 4b). While T N is the same as for x = 0.10 , T i n f has a significantly lower value ≈20 K. Nevertheless, the similarity with the maximum in d ρ / d T for x = 0.10 in Figure 4a implies that the conduction mechanism in x = 0.05 is qualitatively the same, at least in 16 T. As shown in inset of Figure 4b, the maximum in d ρ / d T for x = 0.05 becomes less pronounced with decreasing fields, and in the zero field (and 1 T), it evolves into a knee-like feature, which, we will later argue, is related to the IM crossover, i.e., to the same conduction mechanism.
Our detailed analysis shows that the complex behavior of resistivity at low temperatures can be best fitted to the simple Kondo model:
ρ ( T ) = ρ 0 + q T 2 + ρ K ( T = 0 ) T K 2 T 2 + T K 2 s ,
where ρ 0 is residual resistivity, q is the strength of the standard Fermi liquid (FL) term, ρ K ( T = 0 ) is zero-temperature Kondo resistivity, T K = T K / ( 2 1 / s 1 ) 1 / 2 , T K is the Kondo temperature, and the exponent s is taken to be 0.225 [51,52,53,54]. The T-dependence of resistivity described by Equation (1) is typical for metallic systems with magnetic impurities which, aside from electron scattering on crystal defects (residual resistivity) and electron–electron scattering (FL term), exhibit additional and very specific electron scattering on magnetic impurities, known as the Kondo effect [55]. Here, for the Kondo contribution in Equation (1), we adopt the empirical relation ρ K ( T = 0 ) [ T K 2 / ( T 2 + T K 2 ) ] s used in several transport studies of various Kondo systems [51,52,53,54]. Here, the Kondo temperature T K is defined as the temperature at which the Kondo resistivity is half relative to its zero-temperature value ρ K ( T = 0 ) [51,55].
Fitting of the measured resistivity of the Ca 1 x Gd x MnO 3 samples to the Kondo model described by Equation (1) is shown in Figure 5. Let us first focus on the most instructive fitting of the ρ ( T ) curve for x = 0.10 in the maximum field of 16 T (Figure 5a). As we can see, the logarithmic T-scale reveals another feature of the resistivity curve, namely a tendency towards saturation at the lowest T value (not visible on the linear T-scale in Figure 3a). Furthermore, we can see that the Kondo model covers this feature as well as the local resistivity minimum around ≈20 K and successfully fits the measured resistivity up to ≈80 K, beyond which there are deviations from the Kondo model due to the proximity of the IM crossover. Here, it is worth pointing out that the increase of resistivity below T m i n 20 K , where d ρ / d T < 0 (see Figure 4a), does not imply insulator-like behavior but rather metallic behavior with a Kondo-like temperature dependence.
As shown in the inset of Figure 5a, the Kondo model successfully fits the ρ ( T ) curves of the x = 0.10 sample at all applied fields down to 5 T. Here, note that the T-interval, in which the fits apply, shrinks with decreasing fields and for 5 T covers the range up to only ≈60 K due to the proximity of the IM crossover. The fitting of the resistivity curves at 0 and 1 T to Equation (1) gives a negative value for the FL parameter q, which is not physical. This negative value of q is due to the fact that the resistivity for x = 0.10 at 0 and 1 T does not follow the true metallic behavior ( d ρ / d T > 0 ) in any T-range (see inset of Figure 4a). Nevertheless, if we constrain the fitting procedure in such a way that the parameter q is forced to be zero, it turns out that the Kondo model can reasonably fit even the ρ ( T ) curves in 0 and 1 T, though in a very limited T-range up to only ≈15 K (see inset of Figure 5a). Such behavior implies that the conduction mechanism at low fields is still metallic, but due to the limited T-range, only the contribution of Kondo scattering ( d ρ / d T < 0 ) can be clearly discerned. The FL T 2 term, which is expected to appear at higher T values, is likely masked by the proximity of the IM crossover. At 0 and 1 T, this crossover can be associated with the knee-like feature in ρ ( T ) in Figure 3a or equivalently with the wide maximum in d ρ / d T in the inset of Figure 4a.
Let us now focus on the fitting of the measured resistivity for the Ca 1 x Gd x MnO 3 sample with x = 0.05 in 16 T (Figure 5b). For the same reason as for x = 0.10 at low fields, here, the fitting procedure was constrained in such a way that the parameter q was forced to be zero. As can be seen, such constrained Kondo fitting works reasonably well at low T values up to ≈15 K, beyond which there are deviations, likely related to the proximity of the IM crossover, which by analogy with x = 0.10 may be associated with the wide maximum in d ρ / d T at T i n f in Figure 4b. The same is true for ρ ( T ) curves of x = 0.05 for all fields down to 0 T, as indicated by the inset of Figure 5b. Here, the T-interval, in which the fits apply, shrinks very little with decreasing fields due to the IM crossover that barely shifts with decreasing field and for 0 T is plausibly associated with the knee-like feature in d ρ / d T at ≈20 K (see inset of Figure 4b).
A further confirmation of the viability of the Kondo model can be found in the obtained fitting parameters, which are for both Ca 1 x Gd x MnO 3 samples listed in Table 1. The Kondo temperature T K for both samples and for all fields turns out to be around 20 K, in very good agreement with the typical values in the literature, which span the range of 1–100 K [56] and are commonly around 10–20 K [55]. More importantly, the extracted Kondo temperature agrees very well with those obtained in other rare-earth-doped manganites, such as Nd 0.67 Sr 0.33 MnO 3 [57] and Pr 0.67 x Bi x Ba 0.33 MnO 3 [58], where T K was found to be in the range of 25–30 K. Furthermore, the obtained residual resistivity ρ 0 and the zero-temperature Kondo resistivity ρ K ( T = 0 ) for both samples exhibit a monotonous change with magnetic field, which corroborates that the nature of the conduction mechanism does not change going from high to low fields. Why ρ 0 and ρ K ( T = 0 ) significantly change with magnetic field, while T K and q are nearly constant with field, will be discussed in the next section.

4. Discussion

The results of our magnetotransport study on sintered nanocrystalline samples of Ca 1 x Gd x MnO 3 with x = 0.05 and x = 0.10 , described in detail in the previous section, clearly show that Gd doping progressively enhances the electric conductivity compared to the parent compound CaMnO 3 , which is a well-known G-type AFM insulator [2,3] (Figure 2). Similar behavior has also been found by other groups for both polycrystalline [46] as well as thin film [47] Ca 1 x Gd x MnO 3 samples. The obtained decrease in resistivity is easy to understand because the substitution of Ca 2 + with Gd 3 + introduces the mixed valence state Mn 4 + /Mn 3 + , which leads to increasing numbers of conduction e g electrons (Figure 1). Based on the standard double exchange model, the hopping of these e g electrons favors the parallel arrangement of Mn magnetic moments, which also leads to an increase in the FM interaction in the system. Indeed, our previous magnetic study on the very same Ca 1 x Gd x MnO 3 samples x = 0.05 and x = 0.10 indicated the presence of FM clusters in the G-type AFM matrix [45], with a higher fraction in x = 0.10 , in line with the widely accepted scenario of phase separation in manganites [2,5,26,27,28].
The external magnetic field is shown to have a significant impact on the conducting properties of the studied Ca 1 x Gd x MnO 3 samples, leading to a progressive decrease of resistivity in the field, i.e., to a large negative magnetoresistance (Figure 3). Similar negative magnetoresistance has also been previously found in other polycrystalline [46] and thin film [47] Ca 1 x Gd x MnO 3 samples at low x values. Such behavior is not surprising in manganites, especially in the presence of the FM clusters in the AFM matrix, and is usually ascribed to easy alignment and growth of these FM clusters which intensifies the hopping of conduction e g electrons, in turn leading to a significant decrease in resistivity. The same mechanism is at play also in the famous CMR effect, although there are additional effects that need to be considered [43].
The effect of the field on resistivity is shown to be strong only at low temperatures and weakens significantly with heating, becoming negligible at high temperatures (Figure 3). Such behavior implies that the FM clusters appear only at lower temperatures (below T N ) and that their fraction increases with further cooling, in line with previous magnetic measurements for these samples [45]. Based on the metallic behavior found at low T values, which is especially pronounced for the sample x = 0.10 at high fields (Figure 3a), it seams plausible that the FM fraction at low T values becomes high enough for some of the FM clusters to bind together and form a percolative metallic path through the sample. Such a percolative nature of the metallic state has been proposed for various manganite systems: Pr 0.63 Ca 0.37 MnO 3 [59], La 0.33 Pr 0.34 Ca 0.33 MnO 3 [60], La 0.9 Te 0.1 MnO 3 [61], La 0.5 Ca 0.5 Mn 1 x Al x O 3 δ [62], and La 1 x Ca x MnO 3 [36,43,63], implying a more general feature of the electric conduction in manganites.
Within the percolation picture of manganites, the metallic conduction appears only below a certain characteristic temperature at which the FM (metallic) fraction reaches the percolation threshold. In our case, the percolation threshold can be readily associated with the IM crossover, which is most obvious for the Ca 1 x Gd x MnO 3 sample x = 0.10 at fields 5 T , where ρ ( T ) passes through a local maximum (Figure 3a). For the x = 0.10 sample at fields < 5 T , as well as for the x = 0.05 sample at all fields, where ρ ( T ) does not pass through a local maximum, the percolation threshold might be associated with the maximum or the knee-like feature in d ρ / d T (Figure 4).
The most important finding of this study is that the metallic conduction for x = 0.05 and x = 0.10 beyond the percolation threshold exhibits a peculiar resistivity upturn at the lowest T value, which can be described by the simple Kondo model (Figure 5). Namely, the Kondo effect is a very specific scattering mechanism, typical for metallic systems with magnetic impurities, resulting from a virtual exchange of spin state between the conduction electron and magnetic impurity [55,56]. Kondo scattering has been rarely considered in manganites because most of the studies on manganites R 1 x A x MnO 3 have been conducted on compounds in which the randomly distributed rare-earth cation R 3 + is non-magnetic La 3 + [2,3,4,5,64]. In our study, however, the randomly distributed rare-earth cation is Gd 3 + , which has a large magnetic moment 7 μ B . The Kondo-like behavior in our Ca 1 x Gd x MnO 3 samples might therefore be ascribed precisely to the scattering resulting from an exchange of spin between the conduction electron and magnetic Gd 3 + cation. A direct confirmation of this mechanism could be provided via specific heat or magnetic susceptibility measurements, which is left for a future study.
Here, it is worth mentioning that Kondo scattering is usually found in systems with magnetic impurities, the magnetic moment of which is much smaller than that of Gd 3 + . Especially interesting are studies that associate Kondo scattering with the peculiar resistivity upturn at low T values in less common manganites: ( La 0.1 Ce 0.4 Sr 0.5 ) MnO 3 [65], Nd 0.67 Sr 0.33 MnO 3 [57], Pr 0.67 x Bi x Ba 0.33 MnO 3 [58], La 0.75 Sr 0.20 Mn 1 x Co x O 3 [66], and La 1 x Hf x MnO 3 [67]. In this regard, the Kondo scattering in the system containing Gd 3 + with a large magnetic moment is somewhat surprising. According to the theory, however, the magnitude of the impurity’s magnetic moment is not crucial for the Kondo effect. Instead, its coupling to the conduction electrons is important, which must be antiferromagnetic [55]. Moreover, according to the theory, the Kondo contribution to the resistivity increases with the impurity’s magnetic moment, implying that the systems with Gd 3 + might exhibit even stronger Kondo scattering than systems with other rare-earth elements which have smaller moments, of course, assuming similar antiferromagnetic coupling strength.
In all of the aforementioned previous studies [57,58,65,66,67], the Kondo contribution has been modeled by a conventional theoretically predicted Kondo relation that implies a logarithmic T-dependence ρ ( T ) ln T [55,56]. Our study, however, clearly shows that the resistivity of both studied Ca 1 x Gd x MnO 3 samples x = 0.05 and x = 0.10 strongly deviates from such a logarithmic T-dependence and instead tends towards saturation at the lowest T value (Figure 5). As shown in the previous section, this Kondo contribution can be modeled only by using the empirical relation for the Kondo term ρ ( T ) = ρ K ( T = 0 ) [ T K 2 / ( T 2 + T K 2 ) ] s in Equation (1), put forward by several studies of various Kondo systems [51,52,53,54]. This empirical relation removes the non-physical singular behavior of the ln T term at zero-temperature and accurately predicts saturation of the Kondo contribution towards T = 0 , thus better reflecting the behavior of real systems [55,68]. The exponent s in the empirical relation is usually taken to be 0.225 in order to closely match the theoretical result obtained by the numerical renormalization group [51,52]. Although this numerical prediction was calculated for an impurity with the magnetic moment 1 μ B , here, we show that the same result can also reasonably fit our data for the Gd impurity with the magnetic moment 7 μ B . It is worth mentioning that we have considered alternative quantum correction models as well, such as weak-localization, Coulomb interactions, and variable-range hopping [54,58,66,67]. However, none of these have been able to capture the obtained resistivity upturn, especially its saturation as T 0 .
In contrast to the aforementioned less common manganites [53,54,57,58,66,67], our study shows that the metallic behavior in Ca 1 x Gd x MnO 3 can be modeled by a simple expression that, aside from the Kondo term, contains only the additional conventional FL term ρ 0 + q T 2 (Equation (1)). We further show that some of the fitting parameters exhibit a pronounced dependence on external magnetic field (Table 1), which is not surprising given the obtained large negative magnetoresistance (Figure 3). Nevertheless, the field dependence of the fitting parameters deserves closer attention.
Let us now focus on the residual resistivity ρ 0 , which describes a resistivity contribution that comes from elastic scattering of conduction electrons on crystal defects. First of all, it is worth mentioning that the residual resistivity of our Ca 1 x Gd x MnO 3 samples with x = 0.05 and x = 0.10 is significantly larger than the residual resistivity of the corresponding samples in the study by Maignan et al. [69]. Such behavior might be related to the presence of very small crystallites in our samples of only 30 nm, which gives rise to additional elastic scattering of conduction electrons on crystallite boundaries. Furthermore, as can be seen in Table 1, ρ 0 for both Ca 1 x Gd x MnO 3 samples was found to strongly decrease with the magnetic field. Since the scattering on crystal defects is not expected to change in a magnetic field, the only plausible explanation for the obtained decrease in ρ 0 in the field can be offered by applying the classical percolation picture in the regime just beyond the electrical percolation threshold. According to such percolation picture:
ρ = ρ f V f V c p ,
where ρ is the electrical resistivity of the composite (mixture of a conducting and a non-conducting component), ρ f is the electrical resistivity of the filler (the conducting component), V f is the filler volume fraction, V c is the filler volume fraction at the percolation threshold, and p is the critical percolation exponent, which for a 3D filler network lies between 1.6 and 2 [70,71,72,73,74,75,76]. In the case of our Ca 1 x Gd x MnO 3 samples, the composite refers to the phase-separated state of FM metallic regions within an insulating AFM matrix, and the filler refers to the FM metallic regions. Within this percolation picture, the obtained negative magnetoresistance, i.e., the decrease in measured resistivity in the magnetic field (Figure 3), should be attributed entirely to the field-induced increase in the FM metallic fraction V f without any changes in the intrinsic resistivity of the FM metallic component ρ f . A similar consideration also applies to the obtained decrease in residual resistivity ρ 0 in the magnetic field (Table 1), which should be attributed to the growth of the FM metallic fraction V f rather than to changes in the intrinsic scattering mechanism.
The same percolation model can also nicely explain the field dependence of the Kondo-like contribution in both Ca 1 x Gd x MnO 3 samples, specifically why the obtained zero-temperature Kondo resistivity ρ K ( T = 0 ) strongly decreases with magnetic field while the corresponding Kondo temperature T K stays constant (Table 1). Namely, by comparing Equations (1) and (2), it becomes apparent that in the Kondo term ρ K ( T = 0 ) [ T K 2 / ( T 2 + T K 2 ) ] s , only ρ K ( T = 0 ) should be affected by the field-induced increase in the FM metallic fraction V f , while T K = T K / ( 2 1 / s 1 ) 1 / 2 and consequently T K should stay unchanged. The insensitivity of T K to the field has a physical explanation as well, since within the Kondo model, T K depends on the energy and the width of the impurity’s atomic level, as well as on the Coulomb repulsion energy between two electrons at the site of the impurity [56], which are all atomic quantities that are not expected to significantly change in magnetic field.
Finally, let us focus on the last fitting parameter, the FL parameter q, which is related to inelastic electron–electron scattering and was successfully extracted only for the Ca 1 x Gd x MnO 3 x = 0.10 sample and only for fields 5 T (Table 1). According to the percolation model, the parameter q should follow the same evolution with magnetic field as the parameters ρ 0 and ρ K ( T = 0 ) , which follows naturally from the comparison between Equations (1) and (2). However, despite the limited field range, it can be clearly seen that q does not change with field, in contradiction to the proposed percolation scenario.
The percolation scenario in our Ca 1 x Gd x MnO 3 samples with x = 0.05 and x = 0.10 has at least two additional challenges. The first one is related to the FM metallic fraction V f , which at least in high magnetic fields is expected to be proportional to the magnetization of the sample [36,43]. Based on the obtained field evolution of the fitting parameters ρ 0 and ρ K ( T = 0 ) (Table 1), our magnetotransport study implies that the FM metallic fraction V f for both Ca 1 x Gd x MnO 3 samples continues to significantly increase even at the highest fields of up to 16 T. However, our previous magnetic study conducted on the very same Ca 1 x Gd x MnO 3 samples indicates that the magnetization and therefore V f tends towards saturation already in fields up to only 5 T [45]. The second challenge is related to the quantitative change in the fitting parameters ρ 0 and ρ K ( T = 0 ) , which should within the percolation scenario both change with field by the same factor determined solely by V f . However, our study clearly implies that the zero-temperature Kondo resistivity ρ K ( T = 0 ) changes faster with field than the residual resistivity ρ 0 , which is especially pronounced for the sample x = 0.05 (Table 1).
These deviations from the percolation scenario could point towards the presence of additional contributions to the obtained negative magnetoresistance in our Ca 1 x Gd x MnO 3 samples. One contribution might be associated with the Kondo effect itself, as several studies have reported negative magnetoresistance related solely to Kondo scattering in a magnetic field [53,54,77,78,79]. Another contribution might be associated with a magnetic structure that is more complex than the proposed G-type AFM matrix with FM islands, due to which the relation between the FM metallic fraction and the magnetization would not be as trivial as the one assumed here. Finally, there might be a contribution from tunnel magnetoresistance effect across boundaries between FM metallic and AFM insulating regions, considered in some previous studies of manganites [80,81]. In order to resolve these issues, further experimental and theoretical studies are needed, especially ones based on magnetic measurements at high fields and low temperatures, preferably with spatial resolution.

5. Conclusions

Our magnetotransport study on sintered nanocrystalline Ca 1 x Gd x MnO 3 samples shows that light Gd doping x = 0.05 and x = 0.10 not only drastically reduces the resistivity of the parent antiferromagnetic insulator CaMnO 3 but even leads to metallic conduction at low temperatures. This metallic conduction is found to exhibit a peculiar resistivity upturn at the lowest temperatures, which can be well fitted by the simple Kondo model and possibly be ascribed to Kondo-like scattering of conduction electrons on Gd impurities. We also show a large negative magnetoresistance that can be most easily understood within a classical electrical percolation picture where the conducting component refers to ferromagnetic metallic islands and the non-conducting component refers to the antiferromagnetic insulating matrix. The complex evolution of the metallic conduction in a magnetic field, however, points towards the presence of additional contributions to the negative magnetoresistance that might be related to the field-induced suppression of Kondo scattering and/or tunnel magnetoreistance effect. The identification of these additional contributions and their relation to the percolation picture requires further experimental and theoretical studies.

Author Contributions

Conceptualization, M.Č. (Matija Čulo), N.N., T.I., and M.R.; sample synthesis, M.Č. (Maria Čebela), U.Č., B.L., and M.R.; sample characterization, M.Č. (Maria Čebela), U.Č., B.L., and M.R.; magnetotransport measurements, T.I., M.Č. (Matija Čulo), and N.N.; data analysis, M.Č. (Matija Čulo); interpretation, M.Č. (Matija Čulo), T.I., and N.N.; visualization, M.Č. (Matija Čulo); writing—original draft preparation, M.Č. (Matija Čulo); writing—review and editing, T.I., N.N., M.R., and M.Č. (Matija Čulo). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the project ground states in competition (strong correlations, frustration, and disorder) FrustKor financed by the Croatian Government and the European Union through the National Recovery and Resilience Plan 2021–2026 (NPOO) and by the Ministry of Science, Technological Development and Innovation of the Republic of Serbia, based on Contract no. 451-03-136/2025-03/200017 through the realization of research theme 1702514 (“Vinča” Institute of Nuclear Sciences—National Institute of the Republic of Serbia, University of Belgrade); 451-03-137/2025-03/200123 (Faculty of Sciences and Mathematics, University of Priština in Kosovska Mitrovica); 451-03-136/2025-03/200161, and number 451-03-137/2025-03/200161 (Faculty of Pharmacy, University of Belgrade). We acknowledge project support provided by the Cryogenic Centre at the Institute of Physics—KaCIF co-financed by the Croatian Government and the European Union through the European Regional Development Fund-Competitiveness and Cohesion Operational Programme (Grant No. KK.01.1.1.02.0012).

Data Availability Statement

The data are contained within the article.

Acknowledgments

We thank Mirta Herak and Damir Dominko for elucidating discussions and Đuro Drobac for technical support.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Jin, S.; Tiefel, T.H.; McCormack, M.; Fastnacht, R.A.; Ramesh, R.; Chen, L.H. Thousandfold Change in Resistivity in Magnetoresistive La-Ca-Mn-O Films. Science 1994, 264, 413–415. [Google Scholar] [CrossRef]
  2. Dagotto, E.; Hotta, T.; Moreo, A. Colossal magnetoresistant materials: The key role of phase separation. Phys. Rep. 2001, 344, 1–153. [Google Scholar] [CrossRef]
  3. Salamon, M.B.; Jaime, M. The physics of manganites: Structure and transport. Rev. Mod. Phys. 2001, 73, 583–628. [Google Scholar] [CrossRef]
  4. Nagaev, E. Colossal-magnetoresistance materials: Manganites and conventional ferromagnetic semiconductors. Phys. Rep. 2001, 346, 387–531. [Google Scholar] [CrossRef]
  5. Dagotto, E. Nanoscale Phase Separation and Colossal Magnetoresistance; Springer: Berlin/Heidelberg, Germany, 2003. [Google Scholar]
  6. Zener, C. Interaction between the d -Shells in the Transition Metals. II. Ferromagnetic Compounds of Manganese with Perovskite Structure. Phys. Rev. 1951, 82, 403–405. [Google Scholar] [CrossRef]
  7. Anderson, P.W.; Hasegawa, H. Considerations on Double Exchange. Phys. Rev. 1955, 100, 675–681. [Google Scholar] [CrossRef]
  8. Kubo, K.; Ohata, N. A Quantum Theory of Double Exchange. I. J. Phys. Soc. Jpn. 1972, 33, 21–32. [Google Scholar] [CrossRef]
  9. Goodenough, J.B. Theory of the Role of Covalence in the Perovskite-Type Manganites [La, M(II)]MnO3. Phys. Rev. 1955, 100, 564–573. [Google Scholar] [CrossRef]
  10. Bîrsan, E. The superexchange interaction influence on the magnetic ordering in manganites. J. Magn. Magn. Mater. 2008, 320, 646–650. [Google Scholar] [CrossRef]
  11. Yi, H.; Yu, J.; Lee, S.I. Suppression of ferromagnetic ordering in doped manganites: Effects of the superexchange interaction. Phys. Rev. B 2000, 61, 428–431. [Google Scholar] [CrossRef]
  12. Bao, W.; Axe, J.; Chen, C.; Cheong, S.W.; Schiffer, P.; Roy, M. From double exchange to superexchange in charge-ordering perovskite manganites. Phys. B Condens. Matter 1997, 241–243, 418–420. [Google Scholar] [CrossRef]
  13. Ohtaki, M.; Koga, H.; Tokunaga, T.; Eguchi, K.; Arai, H. Electrical Transport Properties and High-Temperature Thermoelectric Performance of (Ca0.9M0.1)MnO3 (M = Y, La, Ce, Sm, In, Sn, Sb, Pb, Bi). J. Solid State Chem. 1995, 120, 105–111. [Google Scholar] [CrossRef]
  14. Jaime, M.; Salamon, M.B.; Pettit, K.; Rubinstein, M.; Treece, R.E.; Horwitz, J.S.; Chrisey, D.B. Magnetothermopower in La0.67Ca0.33MnO3 thin films. Appl. Phys. Lett. 1996, 68, 1576–1578. [Google Scholar] [CrossRef]
  15. Jaime, M.; Hardner, H.T.; Salamon, M.B.; Rubinstein, M.; Dorsey, P.; Emin, D. Hall-Effect Sign Anomaly and Small-Polaron Conduction in (La1-xGdx)0.67Ca0.33MnO3. Phys. Rev. Lett. 1997, 78, 951–954. [Google Scholar] [CrossRef]
  16. Palstra, T.T.M.; Ramirez, A.P.; Cheong, S.W.; Zegarski, B.R.; Schiffer, P.; Zaanen, J. Transport mechanisms in doped LaMnO3: Evidence for polaron formation. Phys. Rev. B 1997, 56, 5104–5107. [Google Scholar] [CrossRef]
  17. Chun, S.H.; Salamon, M.B.; Han, P.D. Hall effect of La2/3(Ca,Pb)1/3MnO3 single crystals. J. Appl. Phys. 1999, 85, 5573–5575. [Google Scholar] [CrossRef]
  18. Millis, A.J. Lattice effects in magnetoresistive manganese perovskites. Nature 1998, 392, 147–150. [Google Scholar] [CrossRef]
  19. Wollan, E.O.; Koehler, W.C. Neutron Diffraction Study of the Magnetic Properties of the Series of Perovskite-Type Compounds [(1-x)La,xCa]MnO3. Phys. Rev. 1955, 100, 545–563. [Google Scholar] [CrossRef]
  20. Ramirez, A.P.; Schiffer, P.; Cheong, S.W.; Chen, C.H.; Bao, W.; Palstra, T.T.M.; Gammel, P.L.; Bishop, D.J.; Zegarski, B. Thermodynamic and Electron Diffraction Signatures of Charge and Spin Ordering in La1-xCaxMnO3. Phys. Rev. Lett. 1996, 76, 3188–3191. [Google Scholar] [CrossRef]
  21. Chen, C.H.; Cheong, S.W. Commensurate to Incommensurate Charge Ordering and Its Real-Space Images in La0.5Ca0.5MnO3. Phys. Rev. Lett. 1996, 76, 4042–4045. [Google Scholar] [CrossRef]
  22. Mori, S.; Chen, C.H.; Cheong, S.W. Pairing of charge-ordered stripes in (La,Ca)MnO3. Nature 1998, 392, 473–476. [Google Scholar] [CrossRef]
  23. Zang, J.; Bishop, A.R.; Röder, H. Double degeneracy and Jahn-Teller effects in colossal-magnetoresistance perovskites. Phys. Rev. B 1996, 53, R8840–R8843. [Google Scholar] [CrossRef]
  24. Mishra, S.K.; Pandit, R.; Satpathy, S. Mean-field theory of charge ordering and phase transitions in the colossal-magnetoresistive manganites. J. Physics: Condens. Matter 1999, 11, 8561–8578. [Google Scholar] [CrossRef]
  25. Hotta, T.; Malvezzi, A.L.; Dagotto, E. Charge-orbital ordering and phase separation in the two-orbital model for manganites: Roles of Jahn-Teller phononic and Coulombic interactions. Phys. Rev. B 2000, 62, 9432–9452. [Google Scholar] [CrossRef]
  26. Dagotto, E. Complexity in Strongly Correlated Electronic Systems. Science 2005, 309, 257–262. [Google Scholar] [CrossRef]
  27. Nagaev, E.L. Lanthanum manganites and other giant-magnetoresistance magnetic conductors. Physics-Uspekhi 1996, 39, 781–805. [Google Scholar] [CrossRef]
  28. Kagan, M.; Kugel, K.; Rakhmanov, A. Electronic phase separation: Recent progress in the old problem. Phys. Rep. 2021, 916, 1–105. [Google Scholar] [CrossRef]
  29. Freitas, R.; Ghivelder, L.; Levy, P.; Parisi, F. Magnetization studies of phase separation in La0.5Ca0.5MnO3. Phys. Rev. B 2002, 65, 104403. [Google Scholar] [CrossRef]
  30. Giri, S.K.; Nath, T.K. Suppression of Charge and Antiferromagnetic Ordering in La0.5Ca0.5MnO3 Nanoparticles. J. Nanosci. Nanotechnol. 2011, 11, 4806–4814. [Google Scholar] [CrossRef]
  31. Dhieb, S.; Krichene, A.; Boudjada, N.C.; Boujelben, W. Suppression of Metamagnetic Transitions of Martensitic Type by Particle Size Reduction in Charge-Ordered La0.5Ca0.5MnO3. J. Phys. Chem. C 2020, 124, 17762–17771. [Google Scholar] [CrossRef]
  32. Levy, P.; Parisi, F.; Polla, G.; Vega, D.; Leyva, G.; Lanza, H.; Freitas, R.; Ghivelder, L. Controlled phase separation in La0.5Ca0.5MnO3. Phys. Rev. B 2000, 62, 6437–6441. [Google Scholar] [CrossRef]
  33. Quintero, M.; Passanante, S.; Irurzun, I.; Goijman, D.; Polla, G. Grain size modification in the magnetocaloric and non-magnetocaloric transitions in La0.5Ca0.5MnO3 probed by direct and indirect methods. Appl. Phys. Lett. 2014, 105, 152411. [Google Scholar] [CrossRef]
  34. Gamzatov, A.G.; Batdalov, A.B.; Aliev, A.M.; Amirzadeh, P.; Kameli, P.; Ahmadvand, H.; Salamati, H. Influence of the granule size on the magnetocaloric properties of manganite La0.5Ca0.5MnO3. Phys. Solid State 2013, 55, 502–507. [Google Scholar] [CrossRef]
  35. Sarkar, T.; Ghosh, B.; Raychaudhuri, A.K.; Chatterji, T. Crystal structure and physical properties of half-doped manganite nanocrystals of less than 100-nm size. Phys. Rev. B 2008, 77, 235112. [Google Scholar] [CrossRef]
  36. Novosel, N.; Rivas Góngora, D.; Jagličić, Z.; Tafra, E.; Basletić, M.; Hamzić, A.; Klaser, T.; Skoko, Ž.; Salamon, K.; Kavre Piltaver, I.; et al. Grain-Size-Induced Collapse of Variable Range Hopping and Promotion of Ferromagnetism in Manganite La0.5Ca0.5MnO3. Crystals 2022, 12, 724. [Google Scholar] [CrossRef]
  37. Aydogdu, G.; Habermeier, H. Half Ca-doped LaMnO3 films on (001) SLAO and STO substrate studied by magnetization and transport measurements. J. Magn. Magn. Mater. 2009, 321, 1731–1734. [Google Scholar] [CrossRef]
  38. Egilmez, M.; Ma, R.; Chow, K.H.; Jung, J. The anisotropic magnetoresistance in epitaxial thin films and polycrystalline samples of La0.65Ca0.35MnO3. J. Appl. Phys. 2009, 105, 07D706. [Google Scholar] [CrossRef]
  39. Zarifi, M.; Kameli, P.; Ehsani, M.; Ahmadvand, H.; Salamati, H. Effects of strain on the magnetic and transport properties of the epitaxial La0.5Ca0.5MnO3 thin films. J. Magn. Magn. Mater. 2016, 420, 33–38. [Google Scholar] [CrossRef]
  40. Gutiérrez, D.; Radaelli, G.; Sánchez, F.; Bertacco, R.; Fontcuberta, J. Bandwidth-limited control of orbital and magnetic orders in half-doped manganites by epitaxial strain. Phys. Rev. B 2014, 89, 075107. [Google Scholar] [CrossRef]
  41. Zhang, A.M.; Zhang, X.; Zhou, J.; Zhang, W.C.; Lin, J.G.; Wu, X.S. Strain Dependence of Magnetic and Transport Properties in Half-Doped La0.5Ca0.5MnO3 Films. IEEE Trans. Magn. 2015, 51, 1–4. [Google Scholar] [CrossRef]
  42. Xiong, Y.M.; Wang, G.Y.; Luo, X.G.; Wang, C.H.; Chen, X.H.; Chen, X.; Chen, C.L. Magnetotransport properties in La1-xCaxMnO3 (x=0.33,0.5) thin films deposited on different substrates. J. Appl. Phys. 2005, 97, 083909. [Google Scholar] [CrossRef]
  43. Tafra, E.; Novosel, N.; Skoko, Ž.; Ivek, T.; Basletić, M.; Mihaljević, B.; Jagličić, Z.; Góngora, D.R.; Tomić, S.; Hamzić, A.; et al. Colossal magnetoresistance effect and spin-dependent variable-range hopping in the charge ordered phase of overdoped (La, Ca)MnO3 manganites. Phys. Rev. B 2025, 111, 115107. [Google Scholar] [CrossRef]
  44. Dagotto, E. Open questions in CMR manganites, relevance of clustered states and analogies with other compounds including the cuprates. New J. Phys. 2005, 7, 67. [Google Scholar] [CrossRef]
  45. Rosić, M.; Kljaljević, L.; Jordanov, D.; Stoiljković, M.; Kusigerski, V.; Spasojević, V.; Matović, B. Effects of sintering on the structural, microstructural and magnetic properties of nanoparticle manganite Ca1-xGdxMnO3 (x=0.05,0.1,0.15,0.2). Ceram. Int. 2015, 41, 14964–14972. [Google Scholar] [CrossRef]
  46. Beiranvand, A.; Tikkanen, J.; Huhtinen, H.; Paturi, P. Electronic and magnetic phase diagram of polycrystalline Gd1-xCaxMnO3 manganites. J. Alloy. Compd. 2017, 720, 126–130. [Google Scholar] [CrossRef]
  47. Schulman, A.; Beiranvand, A.; Lähteenlahti, V.; Huhtinen, H.; Paturi, P. Appearance of glassy ferromagnetic behavior in Gd1-xCaxMnO3 (0≤x≤1) thin films: A revised phase diagram. J. Magn. Magn. Mater. 2020, 498, 166149. [Google Scholar] [CrossRef]
  48. Rosić, M.; Logar, M.; Devečerski, A.; Prekajski, M.; Radosavljević-Mihajlović, A.; Kusigerski, V.; Spasojević, V.; Matović, B. Synthesis, structural and magnetic properties of nanostructured Ca0.9Gd0.1MnO3 obtained by modified glycine nitrate procedure (MGNP). Ceram. Int. 2011, 37, 1313–1319. [Google Scholar] [CrossRef]
  49. Rosić, M.; Logar, M.; Zagorac, J.; Devečerski, A.; Egelja, A.; Kusigerski, V.; Spasojević, V.; Matović, B. Investigation of the structure and the magnetic behavior of nanostructure Ca1-xGdxMnO3 (x=0.05;0.1;0.15;0.2) obtained by modified glycine nitrate procedure. Ceram. Int. 2013, 39, 1853–1861. [Google Scholar] [CrossRef]
  50. Zeng, Z.; Greenblatt, M.; Croft, M. Large magnetoresistance in antiferromagnetic CaMnO3-δ. Phys. Rev. B 1999, 59, 8784–8788. [Google Scholar] [CrossRef]
  51. Costi, T.A.; Hewson, A.C.; Zlatic, V. Transport coefficients of the Anderson model via the numerical renormalization group. J. Phys. Condens. Matter 1994, 6, 2519–2558. [Google Scholar] [CrossRef]
  52. Goldhaber-Gordon, D.; Göres, J.; Kastner, M.A.; Shtrikman, H.; Mahalu, D.; Meirav, U. From the Kondo Regime to the Mixed-Valence Regime in a Single-Electron Transistor. Phys. Rev. Lett. 1998, 81, 5225–5228. [Google Scholar] [CrossRef]
  53. Lee, M.; Williams, J.R.; Zhang, S.; Frisbie, C.D.; Goldhaber-Gordon, D. Electrolyte Gate-Controlled Kondo Effect in SrTiO3. Phys. Rev. Lett. 2011, 107, 256601. [Google Scholar] [CrossRef] [PubMed]
  54. Cai, X.; Yue, J.; Xu, P.; Jalan, B.; Pribiag, V.S. From weak antilocalization to Kondo scattering in a magnetic complex oxide interface. Phys. Rev. B 2021, 103, 115434. [Google Scholar] [CrossRef]
  55. Kondo, J. Resistance Minimum in Dilute Magnetic Alloys. Prog. Theor. Phys. 1964, 32, 37–49. [Google Scholar] [CrossRef]
  56. Kouwenhoven, L.; Glazman, L. Revival of the Kondo effect. Phys. World 2001, 14, 33–38. [Google Scholar] [CrossRef]
  57. Arun, B.; Suneesh, M.; Vasundhara, M. Comparative Study of Magnetic Ordering and Electrical Transport in Bulk and Nano-Grained Nd0.67Sr0.33MnO3 Manganites. J. Magn. Magn. Mater. 2016, 418, 265–272. [Google Scholar] [CrossRef]
  58. Asmira, N.A.; Ibrahim, N.; Mohamed, Z. Effect of Bi substitution on the collapse of double metal–insulator peaks and induction of a Kondo-like effect in Pr0.67-xBixBa0.33MnO3 manganite. Appl. Phys. A 2023, 129, 741. [Google Scholar] [CrossRef]
  59. Hardy, V.; Wahl, A.; Martin, C. Percolation transitions tuned by temperature, magnetic field, and time in a phase-separated manganite. Phys. Rev. B 2001, 64, 064402. [Google Scholar] [CrossRef]
  60. Zhang, L.; Israel, C.; Biswas, A.; Greene, R.L.; De Lozanne, A. Direct Observation of Percolation in a Manganite Thin Film. Science 2002, 298, 805–807. [Google Scholar] [CrossRef]
  61. Zhou, Q.l.; Jin, K.j.; Zhao, K.; Guan, D.y.; Lu, H.b.; Chen, Z.h.; Yang, G.z. Transport study of percolation modes in the mixed-phase manganites. Phys. Rev. B 2005, 72, 224439. [Google Scholar] [CrossRef]
  62. Shang, C.; Xia, Z.; Zhai, X.; Liu, D.; Wang, Y. Percolation like transitions in phase separated manganites La0.5Ca0.5Mn1-xAlxO3-δ. Ceram. Int. 2019, 45, 18632–18639. [Google Scholar] [CrossRef]
  63. Ziese, M. Critical scaling and percolation in manganite films. J. Phys. Condens. Matter 2001, 13, 2919–2934. [Google Scholar] [CrossRef]
  64. Salazar-Muñoz, V.; Lobo Guerrero, A.; Palomares-Sánchez, S. Review of magnetocaloric properties in lanthanum manganites. J. Magn. Magn. Mater. 2022, 562, 169787. [Google Scholar] [CrossRef]
  65. Eto, T.; Oomi, G.; Honda, F.; Sundaresan, A. High-pressure studies of Kondo-like perovskite (La0.1Ce0.4Sr0.5)MnO3. J. Magn. Magn. Mater. 2001, 226–230, 879–881. [Google Scholar] [CrossRef]
  66. Syskakis, E.; Choudalakis, G.; Papastaikoudis, C. Crossover between Kondo and electron–electron interaction effects in La0.75Sr0.20MnO3 manganite doped with Co impurities? J. Phys. Condens. Matter 2003, 15, 7735–7749. [Google Scholar] [CrossRef]
  67. Guo, E.J.; Wang, L.; Wu, Z.P.; Wang, L.; Lu, H.B.; Jin, K.J.; Gao, J. Magnetic field mediated low-temperature resistivity upturn in electron-doped La1-xHfxMnO3 manganite oxides. J. Appl. Phys. 2012, 112, 123710. [Google Scholar] [CrossRef]
  68. Wilson, K.G. The renormalization group: Critical phenomena and the Kondo problem. Rev. Mod. Phys. 1975, 47, 773–840. [Google Scholar] [CrossRef]
  69. Maignan, A.; Martin, C.; Damay, F.; Raveau, B. Factors Governing the Magnetoresistance Properties of the Electron-Doped Manganites Ca1-xAxMnO3 (A = Ln, Th). Chem. Mater. 1998, 10, 950–954. [Google Scholar] [CrossRef]
  70. Kirkpatrick, S. Percolation and Conduction. Rev. Mod. Phys. 1973, 45, 574–588. [Google Scholar] [CrossRef]
  71. Rahaman, M.; Aldalbahi, A.; Govindasami, P.; Khanam, N.; Bhandari, S.; Feng, P.; Altalhi, T. A New Insight in Determining the Percolation Threshold of Electrical Conductivity for Extrinsically Conducting Polymer Composites through Different Sigmoidal Models. Polymers 2017, 9, 527. [Google Scholar] [CrossRef]
  72. Barrau, S.; Demont, P.; Peigney, A.; Laurent, C.; Lacabanne, C. DC and AC Conductivity of Carbon Nanotubes-Polyepoxy Composites. Macromolecules 2003, 36, 5187–5194. [Google Scholar] [CrossRef]
  73. Brigandi, P.J.; Cogen, J.M.; Pearson, R.A. Electrically conductive multiphase polymer blend carbon-based composites. Polym. Eng. Sci. 2014, 54, 1–16. [Google Scholar] [CrossRef]
  74. Potschke, P.; Arnaldo, M.H.; Radusch, H.J. Percolation behavior and mehanical properties of polycarbonate composites filled with carbon black/carbon nanotube systems. Polimery 2012, 57, 204–211. [Google Scholar] [CrossRef]
  75. Zhang, H.B.; Zheng, W.G.; Yan, Q.; Jiang, Z.G.; Yu, Z.Z. The effect of surface chemistry of graphene on rheological and electrical properties of polymethylmethacrylate composites. Carbon 2012, 50, 5117–5125. [Google Scholar] [CrossRef]
  76. Coelho, P.H.D.S.L.; Marchesin, M.S.; Morales, A.R.; Bartoli, J.R. Electrical percolation, morphological and dispersion properties of MWCNT/PMMA nanocomposites. Mater. Res. 2014, 17, 127–132. [Google Scholar] [CrossRef]
  77. Andrei, N.; Furuya, K.; Lowenstein, J.H. Solution of the Kondo problem. Rev. Mod. Phys. 1983, 55, 331–402. [Google Scholar] [CrossRef]
  78. Joshua, A.; Ruhman, J.; Pecker, S.; Altman, E.; Ilani, S. Gate-tunable polarized phase of two-dimensional electrons at the LaAlO3/SrTiO3 interface. Proc. Natl. Acad. Sci. USA 2013, 110, 9633–9638. [Google Scholar] [CrossRef] [PubMed]
  79. Hien-Hoang, V.; Chung, N.K.; Kim, H.J. Electrical transport properties and Kondo effect in La1-xPrxNiO3-δ thin films. Sci. Rep. 2021, 11, 5391. [Google Scholar] [CrossRef]
  80. Li, H.; Li, L.; Li, L.; Liang, H.; Cheng, L.; Zhai, X.; Zeng, C. Giant intrinsic tunnel magnetoresistance in manganite thin films etched with antidot arrays. Appl. Phys. Lett. 2014, 104, 082414. [Google Scholar] [CrossRef]
  81. Sboychakov, A.O.; Rakhmanov, A.L.; Kugel’, K.I.; Kagan, M.Y.; Brodsky, I.V. Tunneling magnetoresistance of phase-separated manganites. J. Exp. Theor. Phys. 2002, 95, 753–761. [Google Scholar] [CrossRef]
Figure 1. Splitting of the atomic Mn 3d levels into high-energy e g and low-energy t 2 g orbitals in the crystal field created by octahedral surrounding of O atoms. In the case of Mn 3 + , there is additional splitting caused by Jahn–Teller elongation distortion of the MnO 6 octahedron [5] (see text). Valence electrons are shown as red arrows, Mn atoms as red circles, and O atoms as green circles.
Figure 1. Splitting of the atomic Mn 3d levels into high-energy e g and low-energy t 2 g orbitals in the crystal field created by octahedral surrounding of O atoms. In the case of Mn 3 + , there is additional splitting caused by Jahn–Teller elongation distortion of the MnO 6 octahedron [5] (see text). Valence electrons are shown as red arrows, Mn atoms as red circles, and O atoms as green circles.
Nanomaterials 15 00784 g001
Figure 2. Temperature dependence of resistivity for the Ca 1 x Gd x MnO 3 samples with x = 0.05 (cyan line) and x = 0.10 (pink line). For comparison, the resistivity curve for the parent compound CaMnO 3 from ref. [50] is also shown (black line).
Figure 2. Temperature dependence of resistivity for the Ca 1 x Gd x MnO 3 samples with x = 0.05 (cyan line) and x = 0.10 (pink line). For comparison, the resistivity curve for the parent compound CaMnO 3 from ref. [50] is also shown (black line).
Nanomaterials 15 00784 g002
Figure 3. Temperature dependence of resistivity in applied magnetic field for the Ca 1 x Gd x MnO 3 samples with (a) x = 0.10 and (b) x = 0.05 . Discrete color gradation from light pink (cyan) to dark pink (cyan) corresponds to the increasing field strengths: 0, 1, 5, 8, 10, 12, and 16 T for x = 0.10 ( x = 0.05 ), respectively.
Figure 3. Temperature dependence of resistivity in applied magnetic field for the Ca 1 x Gd x MnO 3 samples with (a) x = 0.10 and (b) x = 0.05 . Discrete color gradation from light pink (cyan) to dark pink (cyan) corresponds to the increasing field strengths: 0, 1, 5, 8, 10, 12, and 16 T for x = 0.10 ( x = 0.05 ), respectively.
Nanomaterials 15 00784 g003
Figure 4. Temperature dependence of resistivity derivative d ρ / d T in the maximum magnetic field μ 0 H = 16 T for the Ca 1 x Gd x MnO 3 samples with (a) x = 0.10 and (b) x = 0.05 . Here, μ 0 is the vacuum permeability. Characteristic temperatures of T N , T I M , T i n f , and T m i n are indicated by black arrows. The T-dependence of d ρ / d T at all fields is shown in the insets, where different field strengths of 0, 1, 5, 8, 10, 12, and 16 T correspond to the discrete color gradation from light pink (cyan) to dark pink (cyan) for x = 0.10 ( x = 0.05 ), respectively.
Figure 4. Temperature dependence of resistivity derivative d ρ / d T in the maximum magnetic field μ 0 H = 16 T for the Ca 1 x Gd x MnO 3 samples with (a) x = 0.10 and (b) x = 0.05 . Here, μ 0 is the vacuum permeability. Characteristic temperatures of T N , T I M , T i n f , and T m i n are indicated by black arrows. The T-dependence of d ρ / d T at all fields is shown in the insets, where different field strengths of 0, 1, 5, 8, 10, 12, and 16 T correspond to the discrete color gradation from light pink (cyan) to dark pink (cyan) for x = 0.10 ( x = 0.05 ), respectively.
Nanomaterials 15 00784 g004
Figure 5. Fitting of the measured resistivity of the Ca 1 x Gd x MnO 3 samples with (a) x = 0.10 and (b) x = 0.05 to the simple Kondo model described by Equation (1) in the maximum magnetic field μ 0 H = 16 T . The fitting procedure at all fields is shown in the insets, where different field strengths of 0, 1, 5, 8, 10, 12, and 16 T correspond to the discrete color gradation from light pink (cyan) to dark pink (cyan) for x = 0.10 ( x = 0.05 ), respectively. The measured data are shown as empty symbols, and the corresponding fits are given as black lines. To avoid non-physical behavior, the fits of ρ ( T ) for x = 0.10 at 0 and 1 T, as well as the fits of ρ ( T ) for x = 0.05 at all fields, were performed in such a way that the parameter q in Equation (1) was forced to be zero (see text).
Figure 5. Fitting of the measured resistivity of the Ca 1 x Gd x MnO 3 samples with (a) x = 0.10 and (b) x = 0.05 to the simple Kondo model described by Equation (1) in the maximum magnetic field μ 0 H = 16 T . The fitting procedure at all fields is shown in the insets, where different field strengths of 0, 1, 5, 8, 10, 12, and 16 T correspond to the discrete color gradation from light pink (cyan) to dark pink (cyan) for x = 0.10 ( x = 0.05 ), respectively. The measured data are shown as empty symbols, and the corresponding fits are given as black lines. To avoid non-physical behavior, the fits of ρ ( T ) for x = 0.10 at 0 and 1 T, as well as the fits of ρ ( T ) for x = 0.05 at all fields, were performed in such a way that the parameter q in Equation (1) was forced to be zero (see text).
Nanomaterials 15 00784 g005
Table 1. Extracted parameters from the fitting of the measured resistivity of the Ca 1 x Gd x MnO 3 samples with x = 0.10 and x = 0.05 to the Kondo model described by Equation (1) and shown in Figure 5. The error bars of the fitting parameters ρ 0 , q, and ρ K ( T = 0 ) reflect 20% uncertainty in sample geometry, and the fitting parameter T K reflects the uncertainty in the fitting procedure.
Table 1. Extracted parameters from the fitting of the measured resistivity of the Ca 1 x Gd x MnO 3 samples with x = 0.10 and x = 0.05 to the Kondo model described by Equation (1) and shown in Figure 5. The error bars of the fitting parameters ρ 0 , q, and ρ K ( T = 0 ) reflect 20% uncertainty in sample geometry, and the fitting parameter T K reflects the uncertainty in the fitting procedure.
x = 0.10 x = 0.05
μ 0 H
(T)
ρ 0
( 10 2 Ω cm)
q
( 10 7 Ω cm K 2 )
ρ K ( T = 0 )
( 10 3 Ω cm)
T K
(K)
ρ 0
( Ω cm)
q
( 10 7 Ω cm K 2 )
ρ K ( T = 0 )
( Ω cm)
T K
(K)
0 7.8 ± 1.5 - 14.8 ± 3.0 20 ± 7 2.0 ± 0.4 - 5.6 ± 1.1 19 ± 7
1 6.3 ± 1.3 - 12.3 ± 2.5 19 ± 7 1.9 ± 0.4 - 3.5 ± 0.7 19 ± 7
5 4.3 ± 0.9 7.9 ± 1.6 10.0 ± 2.0 21 ± 5 1.3 ± 0.3 - 1.3 ± 0.3 19 ± 7
8 3.5 ± 0.7 8.3 ± 1.7 7.7 ± 1.5 18 ± 5 1.0 ± 0.2 - 0.8 ± 0.2 22 ± 7
10 3.2 ± 0.6 8.3 ± 1.7 6.5 ± 1.3 19 ± 5 0.9 ± 0.2 - 0.7 ± 0.1 22 ± 7
12 2.9 ± 0.6 8.3 ± 1.7 5.6 ± 1.1 20 ± 5 0.8 ± 0.2 - 0.5 ± 0.1 21 ± 7
16 2.4 ± 0.5 7.4 ± 1.5 3.5 ± 0.7 16 ± 5 0.6 ± 0.1 - 0.4 ± 0.1 22 ± 7
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ivek, T.; Čulo, M.; Novosel, N.; Čebela, M.; Laban, B.; Čakar, U.; Rosić, M. Kondo-like Behavior in Lightly Gd-Doped Manganite CaMnO3. Nanomaterials 2025, 15, 784. https://doi.org/10.3390/nano15110784

AMA Style

Ivek T, Čulo M, Novosel N, Čebela M, Laban B, Čakar U, Rosić M. Kondo-like Behavior in Lightly Gd-Doped Manganite CaMnO3. Nanomaterials. 2025; 15(11):784. https://doi.org/10.3390/nano15110784

Chicago/Turabian Style

Ivek, Tomislav, Matija Čulo, Nikolina Novosel, Maria Čebela, Bojana Laban, Uroš Čakar, and Milena Rosić. 2025. "Kondo-like Behavior in Lightly Gd-Doped Manganite CaMnO3" Nanomaterials 15, no. 11: 784. https://doi.org/10.3390/nano15110784

APA Style

Ivek, T., Čulo, M., Novosel, N., Čebela, M., Laban, B., Čakar, U., & Rosić, M. (2025). Kondo-like Behavior in Lightly Gd-Doped Manganite CaMnO3. Nanomaterials, 15(11), 784. https://doi.org/10.3390/nano15110784

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