1 Introduction

Graphene aerogels have emerged as a highly promising class of materials due to their exceptional properties and diverse applications. Their lightweight and highly porous structures which are composed of interconnected graphene sheets exhibit remarkable mechanical [1], electrical [2], thermal [3], and photo-thermal [7,8,9]. Highly aligned graphene aerogels are characterized by the preferential orientation of individual graphene sheets [10], resulting in a high aspect ratio and a common directionality. This controlled alignment imparts several notable benefits over their randomly oriented counterparts. Firstly, the aligned structure of graphene aerogels significantly enhances their mechanical strength and stiffness, making them capable of withstanding higher loads and exhibiting remarkable structural integrity [11, 12]. This enhanced mechanical robustness is particularly advantageous for applications requiring materials with high strength-to-weight ratios, such as lightweight structural components, flexible electronics, and aerospace materials. Secondly, alignment greatly influences the electrical conductivity of graphene aerogels [13]. The ordered arrangement of graphene sheets allows for efficient charge transport pathways along the aligned direction, resulting in anisotropic electrical conductivity. This can be applied in electronic devices, such as field-effect transistors, where aligned graphene aerogels enable improved charge mobility and device performance. Moreover, the anisotropic conductivity of aligned graphene aerogels facilitates the development of directional electronic devices and sensors [14], opening up new avenues for electronic and sensing applications. Thirdly, alignment plays a vital role in facilitating efficient heat and mass transfer within graphene aerogels [15]. The highly ordered structure enables preferential pathways for thermal conduction, enabling rapid heat dissipation and effective thermal management. Furthermore, the aligned channels within the aerogel structure provide efficient transport pathways for gases and liquids, making aligned graphene aerogels suitable for applications such as thermal steam and organic absorption.

Significant progress has been made in the synthesis and characterization of highly aligned graphene aerogels. Researchers have employed various techniques, including directional freeze casting [4a) [70]. The resultant freeze-dried monolith exhibited well-aligned frameworks along the immersing direction (Fig. 4d) and cellular channel-like architectures in the cross-section view (Fig. 4e). The aligned graphene aerogels was achieved by immersing GO suspensions through a continuous injection into liquid nitrogen [55]. This inventive technique combined spinning technology with ice-templating alignment, facilitating the generation of aligned pore structures. By situating a container filled with a GO suspension atop a thermally conductive base and encompassing it with thermally insulating surroundings on a chilled platform, a unidirectional temperature gradient is generated, emanating from the bottom towards the top (as depicted in Fig. 4b). This thermal configuration prompts the vertical growth of ice crystals. Consequently, the resultant graphene aerogel exhibits vertically aligned graphene walls (Fig. 4f), alongside tubular structures displaying comparatively less alignment within the horizontal plane (Fig. 4g) [74, 83]. Affixing the mold to the side of a metallic box filled with liquid nitrogen generates a horizontal temperature gradient (Fig. 4c), fostering the lateral growth of ice crystals and corresponding alignment [84].

2.1.2 Bidirectional Freeze Casting

Bidirectional freeze casting is a versatile and promising fabrication technique employed for the production of materials with aligned and ordered microstructures in two perpendicular directions, possible in constructing centimeter-scale long-range ordered 3D structures [27, 86]. Different from conventional or unidirectional freeze casting, bidirectional freeze casting involves temperature gradients simultaneously along horizontal (ΔTH) and vertical (ΔTV) directions (Fig. 5a–c) [27, 87, 88]. During the freezing process, ice crystals gradually grow in both horizontal and vertical orientations, resulting in the formation of microstructures exhibiting aligned features in the two orthogonal directions [27, 89, 90].

Fig. 5
figure 5

Copyright 2018, American Chemical Society. Copyright 2021, Wiley–VCH. Copyright 2020, Elsevier. Schematic and morphology comparisons of aerogels prepared using d, e bidirectional freezing, f, g unidirectional freezing and h, i common freezing. Reproduced with permission [88]. Copyright 2020, Elsevier

a–c Schematic illustrations of three typical bidirectional freeze casting apparatus and processes for the fabrication of highly aligned graphene aerogels. Reproduced with permission [27, 87, 88].

An approach involves the use of a thermally conductive copper bridge, partially submerged in liquid nitrogen at one end and exposed to a higher-temperature environment at the other, effectively generating a dual-temperature bridge. Placing GO suspensions onto this bridge results in the creation of a horizontal temperature gradient extending from the liquid-nitrogen side to the other, along with a vertical gradient from the copper bridge to the upper suspension, as shown in Fig. 5a [87]. This thermal configuration leads to the initiation of ice crystal nucleation at the edge adjacent to the liquid nitrogen, followed by growth in both horizontal and vertical orientations. The low-temperature end of the metal plate can be effectively cooled by utilizing substances like liquid nitrogen, dry ice, or chilled ethanol, whereas the high-temperature end can be exposed to air or water for efficient temperature control [59, 89, 91]. The temperature difference between the high- and low-temperature ends, as well as the placement of the mold, determine the temperature gradients in both horizontal and vertical orientations, influencing the inter-lamella spacing within aerogels. After freeze drying and reduction, graphene aerogels with aligned lamella are resulted.

The incorporation of a polymer wedge at the bottom of the mold containing GO suspensions has been proven to be an effective method to precisely adjust the temperature gradient [92, 93]. This bottom-wedged container is situated typically on a metallic platform that is partially immersed in liquid nitrogen or other cold media to attain the necessary freezing temperatures, as illustrated in Fig. 5b [27, 94]. Ice crystals initiate at the lowest edge of the wedge, growing along both the wedge and the vertical direction. The rate of cooling and the slope angle of the wedge collaboratively determine the alignment observed in the final structures. For instance, Bai and colleagues achieved a monodomain structure—a consistent single orientation across the entire sample—with a wedge slope angle of 20° and a cooling rate of 5 or 10 °C min−1 [95]. At cooling rates of 5 and 10 °C min−1, the alignment was enhanced with increasing wedge angle up to 20°. However, when the cooling rate was low (1 °C min−1), no long-range alignment was obtained, regardless of the wedge angle. An open-top cubic mold, comprising a thermally conductive copper bottom, a copper side, and three thermally insulating polymeric foam sides, can also be employed to create a dual-temperature gradient for bidirectional freeze casting, as shown in Fig. 5c [88]. When the mold is in contact with cold sources, the copper bottom and copper side experience notably rapid temperature reduction compared to the three polymeric foam sides, consequently establishing dual temperature gradients extending from the bottom to the top and from the copper to the polymeric foam side.

The comparison of schematic diagrams and morphologies of aerogels fabricated using bidirectional, unidirectional, and conventional freeze casting methods is illustrated in Fig. 5d–i [88]. Among these approaches, the bidirectional freeze casting method featuring dual temperature gradients (Fig. 5d) stands out for its capability to produce a highly organized large-size single-domain lamellar architecture. This architecture displays straight or undulating microstructures that are parallel to the plane of ice crystal growth (the yz plane) [96], as evidenced by the alignments observed in the cross section (the xy plane) and the longitudinal section (the xz plane) (Fig. 5e). Differently, unidirectional freeze casting yields organized microstructures solely along the direction of ice growth (the xz plane), while the plane perpendicular to the ice growth (the xy plane) exhibits multidomain alignments or honeycomb patterns, as shown in Fig. 5f, g. By comparison, the common freeze casting method (Fig. 5h) results in a less ordered microstructure in both the xy and xz planes (Fig. 5i). These distinct microstructural variations across the aerogels carry significant implications for their mechanical flexibility and multifunctional characteristics. The highly aligned graphene aerogel fabricated by bidirectional freeze casting holds the potential to possess outstanding properties, attributing to its precisely defined and well-ordered microstructure.

2.1.3 Radial Freeze Casting

Radial freeze casting involves a freezing process characterized by a radial temperature gradient typically extending from the outer-container to the center, which prompts ice crystals to grow radially inward [97]. This technique allows researchers to manipulate the freezing dynamics, leading to the creation of porous scaffolds with a distinct radial microstructure reminiscent of the natural cellular tracheid patterns. Drawing inspiration from the remarkably efficient capillary transport of water observed in trees, the concept of radial freeze casting has been harnessed to fabricate porous biomimetic structures that replicate the intricate cellular tracheid arrangements found in conifer trees [6a) creates a dynamic thermal landscape characterized by dual temperature gradients along both the vertical and the radial directions (Fig. 6b). With a significantly lower thermal conductivity of the polymer bottom than the surrounding metallic container, the ice grows predominantly in the radial direction, while both longitudinal and radial growth of ice crystals happen (Fig. 6c) [99]. After ice sublimation, highly aligned graphene aerogels with radially aligned cross sections and vertically aligned walls are obtained (Fig. 6d).

Fig. 6
figure 6

Copyright 2018, American Chemical Society. e Schematic diagram of radial freeze casting of rGO hydrogel and f microstructure of corresponding aerogels. Reproduced under the terms of the Creative Commons CC BY license [102]. Copyright 2021, The Authors, published by Springer Nature. g Fabrication of radiating graphene aerogel spheres by immersing GO droplets into liquid nitrogen. Reproduced with permission [103]. Copyright 2017, American Chemical Society; and h, i their cross-sectional microstructure. Reproduced with permission [104]. Copyright 2016, American Chemical Society

Illustration of a freezing setup for fabricating graphene aerogels with radial orientation, b the temperature gradient, c the freezing-drying process, and d SEM morphologies of a radiating graphene aerogel. Reproduced with permission [99].

The radial freezing can also be employed for the rearrangement of building blocks in hydrogels. Lin et al. applied a radial freeze casting approach to convert reduced graphene oxide (rGO) hydrogels characterized by concentric rings into a captivating 3D interconnected network, resembling a spider-web structure [102]. This transformative process unfolded within a specialized metal mold, featuring a thermally insulated polymeric base immersed in liquid nitrogen, as shown in Fig. 6e. During freeze casting, the outer circumference of the mold makes direct contact with the frigid liquid nitrogen, creating a radial temperature gradient that triggers the growth of ice crystals from the mold's periphery towards its center. This promotes the reconfiguration of the previously separate rGO concentric-ring walls, originally formed through hydrothermal self-assembly, into a complex spider-web-like network after freeze-drying, as shown in Fig. 6f.

As elucidated in the “Unidirectional freeze casting” section, the gradual immersion of the GO suspension in liquid nitrogen results in a unidirectional ice growth. However, a captivating shift occurs when plunging the suspension in a cylindrical mold into liquid nitrogen. This swift immersion engenders an instantaneous and uniform exposure of the entire suspension to the frigid milieu, instigating the inception of temperature gradients rippling from the peripheral edges to the core of the cylinder [69]. Consequently, ice crystals grow radially, resulting in radial alignment of graphene sheets in the resultant aerogels.

Similarly, the fabrication of freestanding porous graphene aerogel beads, characterized by radially oriented channels extending from the sphere's surface to its center, was developed through the high-voltage-assisted fast casting of spherical GO suspensions into liquid nitrogen, as shown in Fig. 6g [103,104,105]. As these GO dispersion droplets were immersed in the frigid liquid nitrogen environment, an immediate and substantial temperature gradient ensued, triggering the rapid nucleation of ice crystals at the sphere's surface. The ice crystals then grew progressively from the surface towards the center of the sphere, as shown in Fig. 6g. A cross-sectional analysis unveils the presence of these radial channels, characterized by meticulously aligned graphene walls that span from the sphere's center to its surface, as showcased in Fig. 6h, i [104]. This distinct microscopic architecture imparts remarkable supercapacitive properties and exceptional water contaminant absorption capabilities to these graphene aerogel spheres, signifying the far-reaching potential of this technique in tailoring advanced material functionalities [103, 104].

2.2 Self-Assembly Induced Alignment

Self-assembly induced alignment in graphene aerogels refers to the process where GO sheets align themselves during the formation of aerogel structure, attributing to unique properties and the inherent ordering tendencies of GO LCs. At lower concentrations, colloids composed of anisotropic particles (here the GO sheets) exhibit an isotropic phase, yet they undergo a transition into a biphasic combination, characterized by the coexistence of both isotropic and nematic phases, as the particle concentration increases [106]. When the concentration of GO sheets reaches the critical transition concentration, a fully formed and distinct nematic phase becomes established, as shown in Fig. 7a [107]. The empirical critical LC concentration, CLC, which is deduced from the well-known Onsager's theory, is determined by [107]:

$${C}_{LC}\approx \frac{4T}{W}$$
(1)

where T and W are the thickness and width of GO sheets, respectively. According to the empirical equation, we can conclude that the critical transition concentration of the GO suspensions is highly related to the aspect ratio of GO sheets.

Fig. 7
figure 7

Copyright 2016, Wiley–VCH. b Nematic phase fraction in GO suspensions against the concentration. Reproduced with permission [109]. Copyright 2014, American Chemical Society. c, d Typical nematic texture and freeze-dried morphology of GO LCs. Reproduced with permission [110]. Copyright 2011, Wiley–VCH. Polarized optical microscopy and Cryo-SEM images of GO LCs in the e, f lateral and g, h center regions, showing helical structures. Reproduced with permission [113]. Copyright 2011, The Authors, published by Springer Nature

GO LC which is the prerequisite of self-assembly induced graphene alignment. a The transition of GO suspensions from isotropic to nematic LCs with increasing concentration. Reproduced with permission [107].

The influence of aspect ratio on the critical concentration for the transition to the nematic phase has been experimentally confirmed through an investigation of the nematic phase fraction in GO suspensions with different sizes of GO sheets [108,109,110]. At a sufficiently low concentration of 0.1 mg mL−1, the large GO (LGO) dispersion with lateral sizes ranging from 0–20 μm remained isotropic, while higher concentrations led to the coexistence of isotropic and nematic phases, resulting in observable macroscopic phase separation after standing for two weeks of self-size separation, as shown in Fig. 7b [109]. The critical concentration for LC transition was determined to be 1.0 mg mL−1 for LGO dispersions. In contrast, the nematic phase in small GO (SGO with lateral sizes ranging from 0–5 μm) suspensions emerged at 2.0 mg mL−1, with the complete transition to an all-liquid–crystal phase occurring at 8 mg mL−1, notably higher concentrations compared to those observed for LGO suspensions.

These distinct textures observed within GO LCs serve as the essential foundation for the fabrication of highly aligned graphene aerogels through self-assembly techniques. By varying concentrations above the CLC and employing different aspect ratios of GO sheets, a diverse array of textures can be achieved within the LC, directly influencing the eventual microscopic orientation of the self-assembled building blocks in the resultant aerogel [111]. Notably, GO LCs with a mean diameter of 1.65 μm and concentrations ranging from 0.3 to 0.5 wt% exhibited Schlieren textures characterized by disclinations of varying signs and strengths (Fig. 7c), effectively reflecting the local orientation of the GO sheets [112]. Correspondingly, SEM images of freeze-dried textured GO LCs revealed a consistent alignment of graphene around a few ± 1/2 disclinations (Fig. 7d), which is consistent with those in LCs. An intriguing long-range helical structure was also observed by Xu et al. within GO LCs [113]. Specifically, at a GO concentration of 0.98 vol%, surpassing the CLC of 0.23 vol%, the GO LC displayed a distinct fingerprint texture, as shown in Fig. 7e, f. This unique texture was further corroborated by freeze-fracture cross-sectional SEM morphologies, revealing a well-ordered circular arrangement of GO sheets (Fig. 7g, h).

Typical self-assembly process of utilizing the in-situ gelation of the ordered GO LCs by hydrothermal treatment to maintain the alignment characteristics. This step involves sealing the GO dispersion within a high-pressure vessel and heating it under controlled conditions, typically at temperatures ranging from 80 to 200 °C [114]. The elevated temperature and pressure create an environment that promotes the reconfiguration and reduction of oxygen functional groups, as well as partial restoration of the sp2 carbon–carbon bonding network in the GO sheets [115]. The hydrothermal treatment also facilitates the self-assembly of rGO sheets into three-dimensional structures due to the reformation of π-π stacking interactions, resulting in the formation of rGO hydrogels (Fig. 8a) [11, 116]. The inherent alignment of LCs remains preserved throughout the hydrothermal process due to the absence of any disruptive factors such as vibration or stirring during the self-assembly process [117, 118]. This natural preservation of order translates into graphene alignment within the resultant aerogels after freeze drying, as shown in Fig. 8b, c [119, 120].

Fig. 8
figure 8

Copyright 2015, American Chemical Society. e–i Self-assembly of base-induced highly ordered GO LCs and microscopic structures of the LC and aerogel. Reproduced with permission [28]. Copyright 2015, Wiley–VCH. j–l Vacuum filtration- and crosslinking-assisted self-assembly. Reproduced with permission [123]. Copyright 2017, Wiley–VCH

Typical self-assembly strategies for the fabrication of highly aligned graphene aerogels. a Hydrothermal treatment induced self-assembly of GO sheets for the fabrication of graphene aerogels and b–d corresponding aerogel morphologies fabricating using different GO concentrations. Reproduced with permission [11].

The alignment achieved in self-assembled aerogels is intricately linked to the inherent orientation of GO LC precursors [116]. As such, factors influencing the LC order, such as the previously discussed concentration and lateral size of GO sheets, play a critical role in determining the eventual alignment of graphene sheets within the aerogels. Appropriate concentration of GO dispersion is crucial for fostering aligned microstructures during the hydrothermal self-assembly process. Insufficient concentrations lead to the absence of LC characteristics, while excessively high concentrations result in limited GO sheet mobility, hampering ordered arrangements (Fig. 8b–d) [11, 121]. Greater lateral dimensions or larger aspect ratios of GO sheets significantly enhance their anisotropic attributes, thereby being advantageous for achieving self-assembled alignment.

When employing conventional-sized GO sheets with an average size < 10 µm, the establishment of a greatly oriented LC needs a large concentration of ~ 10 mg mL−1 [112]. In the pursuit of lightweight materials, or more precisely, materials with elevated specific properties when normalized by density, the creation of ordered structures within GO suspensions at lower concentrations becomes imperative. A base-induced rearrangement of GO sheets within suspensions to form highly oriented LC structures at lower concentrations has been developed. Yao et al. potassium hydroxide (KOH)-induced evolution of GO liquid crystals (3.5 mg mL−1) with notably augmented ordering upon the addition of specific amounts of KOH, as shown in Fig. 8e [28]. Besides the KOH, the strong base sodium hydroxide was also proved to be effective in creating ordered laminar textures in GO LCs [122]. This evolution is a result of two synergistic factors: firstly, partial reduction of GO sheets by KOH extends their rigid domains, facilitating the creation of highly ordered microstructures; secondly, increased electrostatic repulsion between GO sheets enhances suspension fluidity, allowing GO sheets to migrate to regions of lowest Gibbs free energy [28, 122]. The impressive alignment achieved by KOH-induced ordered LCs (Fig. 8f) is seamlessly transferred to the self-assembly process, yielding graphene aerogels with analogous orientational morphologies (Fig. 8g–i).

The incorporation of electrolytes to disrupt the mutual balance of GO sheets within a solution has emerged as an effective strategy for achieving the self-assembly of graphene aerogels by overcoming electrostatic repulsion [124]. The introduction of specific amounts of electrolytes, such as acids [125], salts [126], and organics [124], triggers the formation of GO hydrogels. The electrolyte-induced assembly shares a common characteristic with other self-assembly methods, namely, that the alignment within the resulting graphene aerogels is profoundly influenced by the properties of GO LCs. In order to further enhance the alignment based on conventional GO LCs, the integration of vacuum filtration, which generates shear forces, with the electrolyte-induced assembly has been devised, as illustrated in Fig. 8j [123]. This innovative approach disrupts the electrostatic equilibrium between GO sheets through the introduction of electrolytes, leading to the formation of GO microclusters. Simultaneously, the application of vacuum filtration-driven external forces is harnessed to intensify alignment, thereby fostering the development of a finely tuned and ordered structure (Fig. 8k, l).

2.3 Shear-Induced Alignment

GO dispersions exhibit shear-thinning behavior, characterized by a reduction in viscosity when subjected to shear stress [127]. This phenomenon arises from the realignment and reorientation of GO sheets, causing the disruption of interactions between the sheets. Initially, the presence of randomly dispersed GO sheets leads to a high viscosity due to their mutual interactions. However, the shear stress prompts the GO sheets to align and reorient, decreasing the resistance to flow and subsequently reducing viscosity [127]. The network-like structure formed by GO sheets, combined with their high aspect ratio and functional groups, contributes to the shear-thinning behavior. Upon shearing, GO suspensions can transition from a colloidal isotropic state to a nematic liquid crystal phase [128]. When the concentration of GO dispersion falls below the threshold for biphasic-to-nematic transition, the shear rate-dependent GO structure manifests non-Newtonian properties. At low shear rates, the low-concentration GO suspension displays weak shear-thinning behavior, while a shear-thickening behavior becomes apparent at high shear rates, as exemplified by the Taylor vortex flow in Fig. 9a [129]. In cases of high GO concentration, the more distinguished shear thinning results in the emergence of distinct stripe patterns as shear rates increase, ultimately leading to the formation of highly oriented GO liquid crystals (Fig. 9a).

Fig. 9
figure 9

Copyright 2021, American Chemical Society. b Relaxation of GO LCs after scratching and c-e corresponding mechanisms revealing; and f patterned GO LCs constructed by shearing microlithography. Reproduced under the terms of the Creative Commons CC BY license [131]. Copyright 2019, The Authors, published by Springer Nature

Shearing-induced GO rearrangement in LCs. a Macroscopic observation of GO suspensions under shearing. Reproduced with permission [129].

The shear-thinning properties of GO liquid crystals enable the reconfiguration of GO sheets within the liquid crystal phase. Utilizing mechanical shearing by means of a rod or wire to scratch GO liquid crystals induces the reorientation of GO sheets in the direction of the shear [130]. Over a certain period, this reorientation may gradually subside, with the relaxation time directly correlated to the concentration of GO, as shown in Fig. 9b [131]. In the isotropic phase at lower GO concentrations, GO sheets exhibit a random distribution and engage in independent oscillations near their equilibrium positions, following a Brownian rotational motion. Following a single shearing event, relaxation occurs within tens of seconds. Conversely, the alignment of GO sheets achieved through shearing within the nematic liquid crystal phase remains stable for over 30 days, which suggests that a concentration larger than the CLC is required for the fabrication of highly aligned graphene aerogels via shearing-induced alignment.

The mechanism underlying the shear-induced reorientation of GO sheets is elucidated in Fig. 9c–e [131]. As a rod moves within GO suspensions, it engenders a flow-around-pole phenomenon [132], giving rise to a shear field surrounding the in-motion rod, as shown in Fig. 9c. The width (W) of these localized Stokes flows is related to the rod’s radius and its moving velocity (v), and this relationship can be estimated using:

$$W=2\left(1+\sqrt{\frac{v}{{v}_{cr}}}\right)R$$
(2)

where the vcr is the critical velocity to mobilize GO sheets. As the rod radius increases or the moving speed intensifies, the width of the reoriented GO suspensions expands (Fig. 9d, e). A GO LC possessing a viscosity of ~ 100 Pa·s and a density of ~ 103 kg m−3 was employed to visualize the intriguing phenomenon of shear-induced GO sheet rearrangement. Through scratching the tape-casted GO LCs, where the GO sheets are predominantly aligned parallel to the plane, vertical alignment of GO sheets formed. This intriguing observation, depicted by the highlighted regions in Fig. 9f, holds the potential for practical applications in engineering the conformational aspects within LCs.

On the basis of shear-thinning phenomenon and shear induced alignment of GO suspensions, two typical aligning techniques have been extended for the fabrication of highly aligned graphene aerogels, namely the flow induced alignment (Fig. 10a–d) and the shearing microlithography (Fig. 10e–h). The technique of uniaxial flow has been widely embraced to induce the alignment of GO sheets along a specific flow direction. This approach, when combined with a subsequent freeze-casting process, unveils a promising avenue for crafting graphene aerogels with exceptionally high alignment. The general procedure involves loading GO LCs into a syringe, followed by controlled extrusion through a nozzle at a carefully regulated speed into a frigid bath. Notably, the aligned GO sheets within the LCs come to the fore at the nozzle region and steadfastly maintain their alignment as they freeze in the cold bath, as vividly illustrated in Fig. 10a [133]. Following the subsequent freeze-drying step, this resulting aerogel manifests a morphology akin to that achieved through unidirectional freeze casting techniques. As shown in Fig. 10b, c [55], the material showcases aligned graphene sheets running parallel to the flow direction, while the cross-section reveals a multi-domain tubular structure.

Fig. 10
figure 10

Copyright 2012, 2014, American Chemical Society; and d Layer-by-layer 3D printing through a slit extrusion head Reproduced with permission [29]. Copyright 2022, Wiley–VCH. Shearing microlithography: e microwire shearing. Reproduced with permission [137]. Copyright 2023, Wiley–VCH; and f–h scratching the GO LC using an immersed microprobe Reproduced with permission [131]. Copyright 2019, The Authors, published by Springer Nature

Typical shear-induced alignment for fabricating aligned graphene aerogels. Flow induced alignment: a–c Alignment during ejecting from a nozzle Reproduced with permission [55, 133].

The flow-induced alignment can also be combined with 3D printing to fabricate carefully designed hierarchical graphene aerogels. A case in point is the work by Wang et al., who achieved the fabrication of a high-density graphene aerogel microlattice with an ordered structure through the direct ink writing of glycerol-functionalized GO LCs [134]. They yielded a striking outcome with highly aligned filaments within the structure. Furthermore, when employing a slit extrusion head, the potential of producing planar graphene aerogels with horizontally aligned building blocks becomes feasible. This was achieved through a layer-by-layer deposition of GO dispersions on a cooled platform, followed by freeze-drying, as illustrated in Fig. 10d [29]. It's noteworthy that the specific morphology of the printed graphene aerogel is intricately intertwined with the printing speed and the diameter of the slit extrusion, imparting a tunable quality to the process.

The method of shearing microlithography entails the movement of a stick within the GO LCs, a process that facilitates the emergence of macroscopically aligned two-dimensional (2D) GO nanosheets. The outcome of this procedure is visually striking, revealing distinct dark and bright textures that represent the anisotropic optical responses of the meticulously ordered GO nanoflakes [135]. The larger GO sheets, boasting a greater aspect ratio, exhibit a more pronounced moment of force under the influence of the shearing field [108]. This engenders a larger driving force, enabling them to overcome the resistive hydrodynamic drag within the viscous suspension [136]. In a similar vein, the orientation of graphite oxide suspensions can also be meticulously controlled through shearing, serving as a valuable reference for aligning GO. A boundary-free vertically-moving microwire shearing technique has been implemented to achieve a precise vertical orientation of colossal graphite oxide sheets, as vividly depicted in Fig. 10e [137]. Large-scale vertical arrays of giant graphite oxide flakes were fabricated by narrowing the interval spacing (S) between adjacent shearing fields. The initial cycle of this process resulted in a π wall, signifying a distinct nonuniform structure marked by vertical alignment within the central domain transitioning to horizontal alignment at the boundaries, where the width (W) spanned 50 µm [137]. With a favorable S/W ratio of 0.1, a remarkably well-aligned porous microstructure emerged, displaying an intriguing anisotropic skeleton with bidirectional sheet ordering.

Vertical alignment of GO sheets can be effectively achieved through the horizontal sliding of a probe immersed within the GO suspension. This technique allows for the digital programming of GO LCs, offering the capability to program the alignment or texture of GO sheets. Different moving routines of the microprobe can be employed to achieve precisely programmed alignment patterns [138]. In cases where the probe is guided towards a specific direction within a sufficiently small interval, a uniform and distinctly vertical alignment of GO sheets is realized, as schematically illustrated in Fig. 10f, g [131]. Through a meticulous freeze-drying process, this vertical alignment phenomenon becomes an inherent trait of the resultant graphene aerogel, as exemplified in Fig. 10h.

2.4 Further Enhancement of Alignment

While GO presents a promising foundation for the construction of aligned graphene aerogels, its extensive oxygen-containing groups, particularly hydroxy groups, situated on the basal plane of GO, lead to the creation of more hydrogen bonds with ice crystals than with liquid water. This intricate interaction inhibits the growth of ice crystals [139], consequently impeding the alignment in resulting aerogels even under the conditions of unidirectional or bidirectional freeze casting. Efforts have been made to mitigate the interactions between GO sheets and ice crystals that form in situ, aiming to enhance alignment within GO architectures, including, (i) antifreeze-assisted freezing [30], (ii) cation-assisted freezing [35], and (iii) additional thawing and freezing [140] (Fig. 11).

Fig. 11
figure 11

Copyright 2017, American Chemical Society. e–g Cation-facilitated alignment. Reproduced with permission [35]. Copyright 2021, Elsevier

Further enhancement of alignment in graphene aerogels. a–d The antifreeze-assisted aligning. Reproduced with permission [30].

Organic solvents, widely recognized as potent antifreeze agents, exert a notable influence on the freezing dynamics, altering the crystallization patterns of ice crystals via their interaction with hydrogen bonds [150]. The architectures of rGO aerogels originating from partially reduced GO microgels are subsequently modulated during the ensuing freeze-thaw phase. Central to this is the fine-tuning of inter-sheet π-π interactions during the supplementary thawing and reduction phases. This optimization confers a robust structural framework that withstands the additional ice growth process [151]. Through subsequent cycles of thawing and freezing, the arrangement of rGO sheets is further refined, culminating in heightened alignment of constituent building blocks and the realization of highly oriented graphene aerogels [149].

2.5 Comparison of the Aligning Techniques

Based on the above discussion regarding the fabrication of aligned graphene aerogels, a comparative overview is summarized in Table 1. Driving forces of the directional freeze casting, self-assembly and shear-induced alignment are the growth of ice crystals, reduction-induced gelation, and shearing, respectively. Morphologies of graphene aerogel fabricated by directional freeze casting is highly related to the direction of temperature gradient, whereas self-assembly and shearing induced alignment is dependent on the orientation of GO LCs and the shearing direction, respectively (Fig. 2). Lamellar structure can be obtained from the bidirectional freeze casting and shearing forced aligning. Long-range alignment is easy to be obtained by techniques except self-assembly induced aligning. Owing to their highly aligned lamellar structure and remarkable superelasticity, bidirectionally freeze-cast graphene aerogels find particular applicability in pressure sensing and thermal management applications. Meanwhile, the radial alignment in radially freeze-cast aerogels affords ample surrounding channels for the movement of organic solvents and oils, making them advantageous for organic absorption applications.

Table 1 Comparisons of different aligning techniques for graphene aerogels

The costs of different fabrication techniques provide important information for industries manufacturing. The relative fabrication cost of the aligning techniques discussed in this review are summarized in Table 1. The directional freeze casting method involves a freezing process using cold sources, such as liquid nitrogen or chilly ethanol. The freezing period is relatively short because of the ultralow-temperature cold source, taking about tens of minutes. The design of the freezing mold, generally made by polymers and metals, and the temperature gradient are the key to control the direction of ice growth. With the mold and cold source being main expenses of the directional freeze-casting (no complicated equipment), the easy-processing freeze casting technique are relatively cheap and time-efficient, being the most promising aligning technique for large-scale applications [26, 152]. However, careful optimization of freezing parameters is required for long-range alignment. The self-assembly aligning technique involves reduction induced gelation of GO LCs, generally a hydrothermal process at temperature ranges of 60–200 °C for about a few hours to over 1 day [28, 53, 126]. Taking the electronic energy and time cost into account, the cost for self-assembly method lies in-between the directional freeze casting and shear induced aligning. The shear-induced aligning techniques, including flow induced aligning and shearing microlithography of GO LCs, are promising for large-scale fabrication of graphene aerogels with long-range alignment, but the cost of equipment for large aerogels are relatively high. Besides, large-size, high-concentration GO LCs are required, increasing the cost of raw material and make this technique the most expensive among the three aligning techniques [29, 137, 153]. While the freeze casting method demonstrates scalability and a cost advantage compared to other aligning techniques, the overall cost of the graphene aerogel fabrication process remains relatively high for large-scale industrial applications. This mainly includes expenses associated with the costly GO suspension, freeze casting, freeze drying, and post-treatments (such as chemical reduction and high-temperature annealing), limiting practical applications of graphene aerogels.

2.6 Fabrication of Highly Aligned Graphene Aerogel/Polymer Composites

One approach for fabricating aligned graphene aerogel/polymer composites involves constructing highly aligned graphene aerogels followed by polymer infiltration, as shown in Fig. 12a [154]. In this methodology, the foundation lies in the fabrication of highly aligned graphene aerogels, a task achieved through diverse means such as template-based methodologies, self-assembly, or precision alignment techniques like shearing. These strategies confer meticulous command over the spatial orientation of graphene within the aerogel's intricate architecture. Once the coveted highly aligned graphene aerogels are realized, they are subsequently infused with a polymer solution or precursor, facilitating the complete impregnation of the polymer matrix into the porous aerogel framework. This meticulous infiltration ensures an intimate interlocking between the polymer and the pre-existing aligned graphene scaffold. Subsequently, the solvent is extracted or the polymer is cured, obtaining an aligned graphene/polymer composite [143]. Importantly, post-polymer infiltration, the profound orientation of graphene walls within the aerogel remains impeccably preserved, as shown in Fig. 12b, c. The resultant composite material showcases significantly enhanced mechanical, electrical, and thermal attributes, rendering it a versatile contender for a spectrum of applications [155].

Fig. 12
figure 12

Copyright 2019, Elsevier; and d directional freeze-drying of GO/polymer suspensions. Reproduced under the terms of the Creative Commons CC BY license [156]. Copyright 2016, The Authors, published by Springer Nature. e Nanoscopic alignment of graphene sheets in the aligned composite skeleton. Reproduced with permission [157]. Copyright 2017, Elsevier

Fabrication of aligned graphene aerogel-based composites via a–c post-infiltration of polymer matrix in the preconstructed aligned graphene aerogels. Reproduced with permission [154].

An alternative avenue for crafting aligned graphene /polymer composites aerogel involves the technique of directional casting of graphene/polymer solutions, as shown in Fig. 12d. This method entails the preparation of a graphene/polymer solution, achieved by uniformly dispersing graphene within a polymer solution to obtain a homogenous solution or suspension, typically an aqueous medium. The ensuing step encompasses the casting or pouring of this solution into a designated mold, subsequently subjecting it to the aligning procedures elucidated earlier. Following the freeze-drying phase, the resultant composite aerogel embodies the framework of aligned graphene/polymer composite walls [156]. Notably, these aligned graphene/polymer composite aerogels can also be subjected to external compressive forces for the fabrication of composites, ultimately culminating in the formation of solid composites furnished with oriented graphene fillers (Fig. 12e) [157].

Overall, both methodologies present potent avenues for the creation of aligned graphene aerogel/polymer composites, each proffering distinct merit. The first approach centers on the meticulous fabrication of highly aligned graphene aerogels, followed by polymer impregnation, obtaining aligned 3D interconnected graphene aerogel networks within composites [143]. Conversely, the second technique employs the directed casting of graphene/polymer solutions, promoting the simplicity and scalability. Uniformly dispersed and aligned graphene sheets in the aligned skeletons are generally obtained. The ultimate choice between these methods relies on factors such as the desired level of alignment control, scalability, and the specific requisites of the intended application [158].

3 Fundamental Properties

3.1 Anisotropic Properties

The carbon atoms constituting the graphene lattice are intricately linked through robust covalent bonds, forming a resilient 2D structural framework. Due to the atomic structure and bonding, graphene exhibits intrinsically anisotropic properties, which lead to anisotropy in highly aligned graphene aerogels. The directional dependence of graphene's properties extends to the aerogel structure when graphene sheets are aligned within it. This directional dependency inherent to graphene's properties becomes entwined with the aerogel's characteristics, thereby imparting direction-specific qualities. As a consequence of the alignment, the aerogel's mechanical, electrical, thermal, transport, EMI shielding, and other properties are significantly influenced along different orientations. For the sake of clarity in our discussion, the directional aspects of properties are graphically represented in Fig. 13a, b.

Fig. 13
figure 13

Copyright 2018, Elsevier. f, g Electrical. Reproduced with permission [5]. Copyright 2016, American Chemical Society. h, i Thermal. Reproduced with permission [34]. Copyright 2022, American Chemical Society. Reproduced under the terms of the Creative Commons CC BY license [166]. Copyright 2020, The Authors, published by Springer Nature. j, k Mass transport. l, m EMI shielding properties. Reproduced with permission [171]. Copyright 2016, American Chemical Society

Anisotropic properties of highly aligned graphene aerogels and their composites. a, b Illustrations of the coordinate directions. Anisotropic properties and relevant mechanism: c–e mechanical. Reproduced with permission [33].

The anisotropy of mechanical stiffness and hysteresis of aligned graphene aerogel and their composites have been demonstrated by researchers [58, 89]. In bidirectionally freeze-dried graphene aerogel, the alignment of graphene creates a continuous load-bearing network, enabling efficient stress transfer and resulting in higher stiffness along the aligned direction (the y direction), as shown in Fig. 13c [33]. Conversely, the absence of a continuous alignment perpendicular to the aligned direction (the x direction) leads to lower stiffness. As a comparison, the unidirectionally freeze-dried graphene aerogels show similar stiffnesses in the x and y directions (Fig. 13d), both being subordinate to that observed in the z direction [33]. This is due to the isotropic porous structure within the xy plane and the alignment of skeleton along the y direction [159]. The comparison of mechanical properties between bidirectionally and unidirectionally freeze-dried graphene aerogel confirms the contributions of graphene alignment to macroscopic properties. Hysteresis, indicative of energy dissipation during cyclic loading and unloading, generally exhibits lower values perpendicular to the alignment, predominantly because of weaker frictions among graphene sheets and gentler airflow within the porous structure during deformation [37, 58]. Consequently, highly aligned graphene aerogels show super-elasticity exceptional and resilience upon cyclic compression [160,161,162]. An example is that the bidirectional freeze-casting graphene aerogels a 96% recovery of height after a compression to 90%, as shown in Fig. 13e [33]. The anisotropic mechanical properties of aligned graphene aerogel can be inherited their composites, showing different stress–strain characteristics along different directions [154].

The delocalized π-electrons within the carbon hexagonal lattice enable electrons to move unrestrictedly along the graphene plane, rendering high electrical conductivity along its in-plane direction. This characteristic extends to graphene aerogels and their composites, where aligned graphene conductive frameworks give rise to anisotropic electrical conductivity, marked by enhanced conductivity along the aligned axis and diminished conductivity perpendicular to it [34]. An illustrative example of this phenomenon is observed in the research by Wang et al. who investigated anisotropic electrical conductivities in graphene aerogel freeze-dried from GO suspensions of varying concentrations, as shown in Fig. 13f [5]. The anisotropy of electrical conductivity in graphene aerogels becomes more pronounced with higher concentrations of GO, attributing to increased alignment of graphene sheets induced by the elevated GO concentration. Similarly, Gao et al. reported on highly aligned graphene aerogel/PDMS composites displaying electrical conductivity notably higher by 1–2 orders of magnitude in the z direction compared to the x direction [154]. The anisotropy in electrical conductivity is attributed to the aligned graphene sheets within the aerogel structure, which establish a continuous conductive network along the alignment direction and an inconsecutive bridging structure in the transverse direction, thus enabling more efficient electron transport and higher conductivity along the direction of alignment (Fig. 13g).

The interplay of convection, conduction, and radiation mechanisms significantly influences the anisotropic thermal conductivity observed in aligned graphene aerogels. The arrangement of graphene sheets within the aerogel creates an uninterrupted conductive network, facilitating efficient electron and phonon transport and consequently enhancing thermal conduction [34], as illustrated by the solid blue arrows in Fig. 13h. Notably, the thermal conduction occurring between the aligned graphene skeletons is markedly higher along the alignment direction [34], owing to the relatively lower thermal conductivity of air present among aligned graphene walls. Additionally, the occurrence of thermal convection perpendicular to the alignment is negligible due to the limited pore size among the graphene walls, which restricts the onset of natural thermal convection [163]. The distinctive anisotropic structure also enhances heat dissipation along the z direction, preventing heat accumulation [34]. Consequently, aligned graphene aerogels exhibit anisotropic thermal conductivities, showing remarkably higher thermal conductivity along the alignment direction compared to the perpendicular direction [37, 64, 159]. When graphene aerogel-based composites are formed by infiltrating polymers into the pre-constructed porous graphene architectures, the anisotropic thermal conducting traits of aligned graphene aerogels extend to these composites as well [164, 165]. An illustrative example by Liu et al. involves the fabrication of lamellar-like graphene aerogels through bidirectional freeze-casting, succeeded by compaction perpendicular to the alignment direction and vacuum-assisted impregnation with epoxy to produce composites featuring a high alignment of graphene fillers [166]. These composites demonstrate a notable anisotropy in thermal conductivity, exhibiting around 20 and 17.5 W m−1 K−1 along the z- and y- direction, respectively. However, the thermal conductivity decreases substantially to about 1.2 W m−1 K−1 along the x-direction, which is perpendicular to the alignment, as shown in Fig. 13i.

Graphene sheets possess exceptional barrier properties that effectively block the transport of matter through their plane, owing to the tightly packed atomic structure strong bonding energy, and unique electronic configuration of graphene [167]. These factors collectively create a robust barrier that prevents the passage of molecules or atoms through the graphene sheet [168]. Aligned graphene aerogels capitalize on this inherent barrier capability by featuring porous channels oriented along the alignment direction and stacked graphene walls in the transverse orientation. This structural arrangement gives rise to intriguing anisotropic transport characteristics, as illustrated in Fig. 13j, k. The aligned pores establish preferential pathways for the transport of molecules, ions, and substances, offering a conduit for efficient movement. On the contrary, the walls, composed of densely stacked graphene building blocks, exhibit significantly greater resistance, limiting the transport of substances perpendicular to the alignment. This leads to the distinct anisotropic transport properties observed in aligned graphene aerogels, where molecular or atomic movement becomes directionally dependent.

The EMI shielding of materials is attributed to reflection, absorption, and multiple reflection of electromagnetic waves [169]. The reflection of electromagnetic radiation is fundamentally a consequence of the interaction between waves and free charges present on the material's surface, which suggest that the presence of conductive networks characterized by a substantial concentration of charge carriers significantly contributes to the efficacy of reflection-based EMI shielding effectiveness (SE). The SE associated with absorption quantifies a material's capacity to attenuate electromagnetic radiations into thermal or internal energies, which is achieved through processes such as the generation of localized currents within conductive networks, the polarization or relaxation of dipoles and charges, and the charge delocalization [170]. An integral aspect that emerges from these mechanisms is the evident correlation between the EMI shielding properties of materials and their electrical conductivities. Therefore, the anisotropic electrical conductivity intrinsic to aligned graphene and its composites naturally gives rise to anisotropic EMI shielding properties (Fig. 13l) [82, 154, 171]. As electromagnetic radiation penetrates graphene aerogels from a direction perpendicular to the alignment, the waves experience a series of reflections and scatterings at the numerous interfaces presented by the oriented graphene walls. This continuous back-and-forth interaction contributes to the gradual dissipation of electromagnetic waves into thermal energy (Fig. 13m) [171]. But when the waves approach the aerogels parallel to the aligned graphene frameworks, their penetration is comparatively unhindered, leading to a reduction in the dissipation of incident radiations [34, 172].

3.2 Contributions of Alignment to Physical Properties of Composites

3.2.1 Electrical Conductivity

Percolation Theory Electrically conductive materials have diverse applications in electronics, energy storage, sensors, EMI shielding, wearables, automotive systems, and aerospace, emphasizing their essential role in modern technology. When conductive additives are introduced into insulating polymers, the electrical conductivity of the resulting composites undergoes a distinct transition from an insulating to a conducting state as the concentration of conductive components or pathways within the material surpasses a specific threshold, which is well-known as the percolation phenomenon [173]. In the region below this threshold, the composite material behaves as an insulator, displaying minimal to negligible electrical conductivity. However, once the threshold is exceeded, the material makes a pivotal shift into a conductive state, with electrical characteristics becoming exceptionally responsive to the concentration and arrangement of the conductive constituents, as shown in Fig. 14a [174].

Fig. 14
figure 14

Electrical percolation and conductivity of graphene/polymer composites. a Schematic illustrations of the percolation theory. Reproduced with permission [193]. Copyright 2016, Society of Chemical Industry. b Percolation threshold of polymer composites reinforced with different fillers: CVD-grown graphene [176, 177], aligned graphene aerogel [5, 11, 65], random graphene aerogel [178, 179], dispersed graphene [180,181,182,183,184,185], and segregated graphene [186,187,188,189]. c Electrical conductivity of polymer composites reinforced with different fillers: aligned graphene aerogel [5, 65, 171, 190], random graphene aerogel [178, 191, 192], CVD-grown graphene network [177, 194, 195], dispersed graphene [180, 196,197,198,199], and segregated graphene [187,188,189, 200]

The transition is often described as a critical connectivity point where a continuous conductive pathway forms across the material, allowing the flow of electric current. The percolation threshold is a specific concentration value above which the material's conductivity increases dramatically, which can be estimated based on the power law equation [175]:

$$\sigma \propto {\sigma }_{f}{(V-{V}_{p})}^{a}$$
(3)

where σ and σf are the electrical conductivity of composite and filler, V and Vp represent the filler volume fraction and the percolation threshold, and a is a critical exponent. In essence, the percolation phenomenon elucidates the intriguing interplay between the arrangement and concentration of conductive elements within a composite material, fundamentally influencing its electrical properties.

The percolation threshold of various graphene/polymer composites containing different networks of graphene are plotted in Fig. 14b. The composites reinforced with large-aspect-ratio chemical vapor deposition (CVD) grown graphene sheets characterized by a high aspect ratio exhibited remarkably low percolation thresholds, attributing to the pristine lattice structure and the extensive aspect ratio of these graphene sheets [176, 177]. Interestingly, pre-fabricated composites incorporating aligned graphene aerogels showcased exceptionally low percolation thresholds of approximately 0.007 vol% [5, 65]. This strikingly low value is primarily a result of the formation of conductive networks within the well-ordered walls of the aerogel architecture, achieved at an impressively modest filler content. However, an aligned graphene aerogel created through a hydrothermal-induced self-assembly process exhibited a comparatively higher percolation threshold of 0.12 vol% [11], resulting from the stacking of GO building blocks during the hydrothermal process. Comparatively, composites containing randomly oriented graphene aerogels exhibited notably higher percolation thresholds than those composed of aligned graphene aerogels [178, 179]. Composites consisting of random graphene aerogels showed relatively lower or comparable percolation thresholds than those with dispersed graphene sheets [180,181,182,183,184,185]. The segregated graphene reinforced composites [186,187,188,189], generally fabricated by hot compression of graphene coated polymer spheres, displayed percolation thresholds comparable to the random graphene aerogel reinforced composites. In a comprehensive assessment of various graphene network types within polymer composites, it is evident that the aligned aerogel architecture emerges as a promising option.

Electrical Conductivity Upon surpassing the percolation threshold, the electrical conductivity of graphene/polymer composites exhibited a distinct transition from insulating to conducting, showing a rapid surge in electrical conductivity near the percolation threshold followed by a more gradual increase (Fig. 14a). A comprehensive comparison of graphene/polymer composites featuring different graphene network configurations is illustrated in Fig. 14c. Composites reinforced with graphene aerogel, dispersed graphene and segregated graphene commonly employ GO as the graphene precursor, as the precursor for graphene, owing to its ease of processing and favorable compatibility with polymer matrices. The attainment of electrical conductivity is realized through the partial reduction of the GO precursor, resulting in the effective conducting component of rGO within the composites. According to Fig. 14c, it is evident that aligned graphene aerogels represent the most effective rGO fillers among these categories [5, 65, 171, 190], because the arrangement of interconnected rGO building blocks in the oriented aerogel framework serves as efficient conductive pathways. Noteworthy examples include unidirectional freeze-cast graphene aerogel/epoxy composites, where the oriented aerogel frameworks yielded a commendable conductivity of 13.5 S m−1 at a minimal filler content of 0.11 vol% [65]. Furthermore, through a combination of hydrothermal and unidirectional freeze-drying processes, aligned graphene aerogels with enhanced wall thicknesses were achieved, leading to composites exhibiting an electrical conductivity of 980 S/cm at a filler content of approximately 0.4 vol% [171]. In contrast, the randomly oriented graphene aerogel, despite possessing a 3D network similar to that of highly aligned graphene aerogel, exhibited a comparatively less effective conductive behavior due to its more convoluted network structure [178, 191, 192].

Dispersed graphene reinforced polymer composites are commonly prepared using solution processing. With good dispersion and sufficiently large content of graphene, conductive networks can be formed. However, due to the inherent challenge in achieving uniform dispersion of graphene and rGO within polymer matrices, composites containing dispersed graphene sheets often exhibit a certain degree of agglomeration of the filler. This, in turn, results in relatively diminished electrical conductivities, even at higher filler contents [180, 196,197,198,199]. In the process of fabricating composites via the hot compression of graphene-coated polymer particles, segregated graphene-rich regions spontaneously form, engendering electrically conductive pathways. The inherent segregation of graphene within the polymer matrix gives rise to continuous three-dimensional networks, fostering electrical conduction and leading to relatively higher electrical conductivity compared to composites with dispersed graphene configurations [187,188,189, 200]. Composites incorporating networks of graphene grown via CVD, including graphene foam (GF) [177], graphene woven fabric (GWF) [194], and multilayer graphene web (MGW) [195], have demonstrated the most exceptional electrical conductivities within polymer composites. These exceptional conductivities of CVD-grown graphene reinforced composites are attributed to the 3D continuous network and the flawless graphitic structure of the graphene framework.

In summary, from the above comprehensive analysis of the electrical conductivity of graphene/polymer composites featuring different building blocks and structural arrangements, we can conclude that the contribution of these materials to electrical conductivities of composites follows a general trend: CVD-grown graphene network > highly-aligned graphene aerogel > randomly oriented graphene aerogel ≈ segregated graphene > dispersed graphene. Considering that the intrinsic electrical conductivity of rGO building blocks in graphene aerogels is significantly lower than that of graphene building blocks in CVD-grown networks, the pronounced contribution of the aligned structural configuration to enhanced electrical conductivity is evident.

3.2.2 Fracture Toughness

Fracture toughness is a material property that quantifies a material's resistance to fracture or failure when subjected to a crack or other forms of stress concentration. It indicates the amount of energy a material can absorb before fracturing, which is a critical parameter for assessing a material's ability to withstand mechanical loads and resist crack propagation. Indeed, fracture resistance holds even greater significance and importance in practical engineering applications compared to modulus and strength. To gain insights into the contribution of highly-aligned graphene aerogels on composite toughness, a comparison of fracture toughness enhancement with the inclusion of various graphene additives is depicted in Fig. 15a.

Fig. 15
figure 15

Copyright 2015, American Chemical Society

Fracture toughness of graphene/epoxy composites. a Increment in fracture toughness of polymer composites reinforced with different fillers: CVD-grown graphene [177, 194, 195], aligned graphene aerogel [11, 42, 65, 143, 166, 201, 202], random graphene aerogel [65] and dispersed graphene [203,204,205]. b–d Fractured cross-sectional SEM images and corresponding schematic crack prorogation process of aligned graphene aerogel/epoxy composite. Reproduced with permission [11].

The utilization of highly aligned graphene aerogels as reinforcement in epoxy composites results in significant enhancements in fracture toughness. An increment of 98% was witnessed in unidirectional freeze-cast graphene aerogel-reinforced epoxy composites, where the filler content was 2.4 vol% [201]. Epoxy composites incorporating highly aligned graphene aerogels, manufactured through the directional casting of large-size GO suspensions, experienced a notable enhancement of approximately 70% at a meager filler content of 0.11 vol% [65]. Another remarkable case is observed in bidirectional freeze-cast graphene aerogel-reinforced epoxy composites, featuring a bio-inspired nacre structure, which achieves a substantial increment of about 85% at a filler content of 0.65 vol% [202].

As comparisons, with equivalent rGO building blocks, it is evident that the aligned graphene aerogel [11, 42, 65, 143, 166, 201, 202] outperforms the randomly oriented graphene aerogel [65] and dispersed graphene [203,204,205] in terms of fracture toughness enhancement. CVD-grown graphene networks exhibit substantial impacts on the enhancement of the fracture toughness of epoxy-based composites. Notably, an exceptional enhancement of 50–70% can be achieved with a low graphene loading ranging from 0.005 to 0.28 vol% [177, 194]. At higher graphene contents, for instance, 4.1 vol% in the case of MGW, the enhancement reached approximately 108% [195]. With comparable enhancements in fracture toughness between composites reinforced by aligned graphene aerogel and CVD-grown graphene, we can see the significant role of the aligned structure in improving fracture toughness when taking into account the exceptional mechanical properties of nearly flawless and seamlessly continuous CVD-grown graphene.

The distinctive toughening mechanisms responsible for the substantial enhancement of fracture toughness in aligned graphene aerogel/epoxy composites are mainly attributed to the graphene framework induced crack reorientation. As shown in Fig. 15b–d, aligned graphene aerogel/epoxy composites exhibited fracture surfaces with a rough, stair-like appearance, as cracks were deflected by the aligned layers of GO [11]. This deflection resulted in the dissipation of a significant amount of energy, consequently leading to enhanced fracture toughness. The aligned graphene aerogels effectively blunted and diverted crack tips along the interfaces between rGO and epoxy, which effect was visually evident from the fracture surfaces, consisting of smooth epoxy regions measuring around 20–30 µm (subject to the aerogel fabrication process) in size, interspersed by the rGO networks.

3.2.3 Thermal Conductivity

Thermally conductive polymer composites enhance thermal management in industries ranging from consumer electronics to automotive and aerospace. They find crucial applications in efficient heat dissipation for high-power electronic devices and batteries, enabling improved device performance and reliability. Graphene, with its exceptionally high in-plane thermal conductivity of 5300 W m−1 K−1 [206], holds great promise for enhancing the typically low thermal conductivity of polymers of 0.1–0.3 W m−1 K−1. To maximize the contribution of graphene additives, a high quality of graphene additives is important. Highly-aligned graphene aerogels with well-crystalline structures induced by thermal annealing at 2800 °C (pink area in Fig. 19a) [64, 74, 137, 166, 207] significantly outperform partially reduced and aligned counterparts (orange area) [102, 201, 208]. Composites reinforced with aligned graphene aerogel reduced at 2800 °C achieved a remarkable thermal conductivity of approximately 6.6 W m−1 K−1 at a mere 0.77 vol% graphene content (solid stars) [64], while a much higher filler loading of around 2.3 vol% rGO was necessary for an even lower thermal conductivity of 2.6 W m−1 K−1 (red hollow hexagons) [102]. The superior thermal conductivity of high-quality graphene arises from reduced defects and minimized scattering of phonons or electrons, as shown in Fig. 16b, c [15].

Fig. 16
figure 16

Copyright 2021, Elsevier. d Schematic mechanisms of thermal conduction in highly aligned graphene aerogel/polymer composites. The solid and hollow symbols in a represent the graphene and rGO building blocks in the thermally conductive graphene network, respectively

Thermal conductivity of graphene/polymer composites. a Thermal conductivity of graphene/polymer composites reinforced with different fillers: aligned graphene aerogel [64, 74, 102, 137, 166, 201, 207, 208, 221], random graphene aerogel [208,209,210], dispersed graphene [211,212,213], segregated graphene [214,215,216,217], and CVD-grown graphene [195, 218,219,220]. Schematic illustrations of thermal conduction in b defective rGO and c well-crystalline graphene sheets. Reproduced with permission [15].

The arrangement and alignment of graphene play a pivotal role in enhancing the thermal conductivity of graphene/polymer composites, as illustrated in Fig. 16a. Highly-aligned graphene aerogels, thermally reduced at 2800 °C [64, 74, 137, 166, 207], outperform random graphene aerogels [208,209,210], dispersed [211,212,213] and segregated rGO or graphene nanoplatelet (GNP) [214,215,216,217] in enhancing thermal conductivities of composites. They can even surpass the thermal conductivity of CVD-grown graphene/polymer composites [195, 218,219,220]. A polymer composite reinforced with 2800 °C-reduced bidirectionally freeze-cast graphene aerogel, possessing lamellar alignment and high-quality graphene building blocks, achieved an impressive thermal conductivity of approximately 20 W m−1 K−1 along the alignment direction at a graphene content of 2.2 vol% [166]. A high filler content of 14.8 vol% was achieved by 2800 °C-annealed shear-aligned graphite oxide array, contributing to an exceptionally high thermal conductivity of 93 W m−1 K−1 along the alignment [137]. In composites reinforced with CVD-grown graphene, the presence of vertically oriented graphene nanowalls [219] results in higher thermal conductivity compared to composites with random graphene networks [195, 218, 220], providing additional evidence for the beneficial impact of alignment.

The strategies devised to achieve elevated thermal conductivities in composites across various length scales are illustrated in Fig. 16d [195]. At the nanoscale, the 2800 °C annealed graphene building blocks showcase almost flawless crystallinity, leading to a scarcity of phonon scattering sites and yielding graphene sheets with remarkably high intrinsic thermal conductivities. On the microscale, the stacking and interconnection of graphene sheets within aerogel walls facilitate efficient phonon transfer across interfaces due to closely aligned phonon vibrational frequencies. On a larger microscopic scale, the organized alignment of graphene sheets constructs aerogel walls that substantially decrease phonon scattering at filler/polymer interfaces, thereby promoting extended-range phonon transport.

4 Quantitative Characterization Techniques of the Graphene Alignment

4.1 SEM Image-Based Orientation Distribution Analysis

SEM and transmission electron microscopic observations are a highly visualized techniques for alignment characterization, elucidating micro- and nano-scale morphologies across different orientations, respectively [222]. On the basis of SEM images, the extent of alignment finds quantitative assessment through orientation distribution analysis [223, 224]. This method conventionally involves the selection of a statistically substantial graphene sheet pool from SEM captures. Subsequently, reference lines are inscribed on these images, and the acute angles (η) between the reference line and individual graphene sheets are recorded, enabling rigorous quantitative analysis [5]. The orientation distribution function, N(η), is suitably modeled using the GaussAmp peak function:

$$N\left(\eta \right)={N}_{0}+A{e}^{-\frac{{(\eta -{\eta }_{c})}^{2}}{2{w}^{2}}}$$
(4)

where N0, A and ηc are the baseline height, peak amplitude, and the angle corresponding to the peak center, respectively. The quantification of the angular displacement from the peak center, denoted as w, is ascertained through the full width at half maximum (FWHM), as outlined below:

$$w=\frac{FWHM}{2\sqrt{ln4}}$$
(5)

The degree of graphene alignment is quantitatively assessed using the average orientation angle, ⟨cos2 η⟩, with respect to the preferred orientation direction. The ⟨cos2 η⟩ defines the proportion of graphene sheets within the angular element, dη, as indicated by [225]:

$$\langle {cos}^{2}\eta \rangle =\int_{0}^{\pi /2}N\left(\eta \right){cos}^{2}\eta sin\eta d\eta$$
(6)

where N(η) determines the portion of graphene falling within the angular interval dη and satisfies the normalization function:

$$\int_{0}^{\pi /2}N\left(\eta \right) sin\eta d\eta =1$$
(7)

Wang et al. quantitatively describe the alignment degree in unidirectionally cast graphene aerogels by the average orientation angle [5], a concept that also extends to the assessment of filler alignment in composites [175]. The parameter ⟨cos2 η⟩ serves as a metric for the alignment, which effectively characterizes the degree of orientation. The ⟨cos2 η⟩ values of 0 and 1 indicate elements that are respectively perpendicular to or aligned along the preferred orientation, while a ⟨cos2 η⟩ value of 1/3 signifies a 3D random distribution of graphene sheets. Application of the orientation distribution analysis to a typical unidirectional freezing graphene aerogel reveals a distribution of graphene sheets spanning from − 90° to 90°, as shown in Fig. 17a, d. This distribution exhibits a pronounced peak at around 0°, signifying the alignment of graphene sheets along the unidirectional casting direction (Group 1), alongside a less prominent peak centered at approximately ± 90°, corresponding to bridging elements amidst alignments (Group 2), as shown in Fig. 17g. Evaluating ⟨cos2 η1⟩, which pertains to the orientation angles of the predominant group (Group 1), elucidates that graphene sheets in aerogels fabricated using 1 mg mL−1 GO dispersions (Fig. 17b, c) achieved a higher degree of alignment compared to those prepared with 0.5 mg mL−1 GO (Fig. 17e, f).

Fig. 17
figure 17

Copyright 2016, American Chemical Society

Quantitative analysis of the alignment in graphene aerogels using filler orientation distributions. Reproduced with permission [5].

4.2 Polarized Raman Spectroscopy

Polarized Raman spectroscopy is a highly effective and widely embraced technique for characterizing the graphene alignment within diverse materials, which offers an efficient and non-destructive approach for assessing alignment quality, orientation, and the anisotropic attributes of graphene-based materials. A schematic representation of a typical polarized micro-Raman setup is presented in Fig. 18a. This apparatus enables precise control over the polarization of incident laser beams and facilitates the analysis of Raman-scattered light through the utilization of linear polarizers and wave plates [226]. The Raman scattering process within graphene is resonant, rendering it indifferent to the energy of the excitation laser. A wide spectrum of laser wavelengths, ranging from the infrared to the near ultraviolet, can be employed. Through an analysis of the intensity variations in the G- or 2D-band within the Raman spectrum, researchers can extract quantitative insights into the alignment characteristics of graphene within diverse materials. This analytical capability empowers researchers to fine-tune and regulate alignment processes, thereby achieving the desired properties tailored for specific applications.

Fig. 18
figure 18

Copyright 2016, The Authors, published by Springer Nature. b–e Qualitative analysis of graphene alignment. Reproduced with permission [180]. Copyright 2013, Elsevier. Quantitative analysis of graphene alignment: f The Cartesian coordinate system with the sample geometries. Reproduced under the terms of the Creative Commons CC-BY license [230]. Copyright 2015, The Authors, published by Elsevier. g The local orientation of graphene in specimens, h coordinates of the specimen relative to the experimental measurement parameters. Reproduced with permission. Reproduced under the terms of the Creative Commons CC-BY license.[231]. Copyright 2015, The Authors, published by Elsevier. 3D Raman intensity for i VV and j VH polarization configurations, and k relative G band intensity. Reproduced with permission [31]. Copyright 2011, American Chemical Society. l Intensity variation for laser beam propagation in X and Z directions of graphene paper, and m orientation distribution function of HOPG and graphene paper. Reproduced under the terms of the Creative Commons CC-BY license [231]. Copyright 2015, The Authors, published by Elsevier. The designations VV and VH in i–k refer to the parallel and the perpendicular polarization of incident and scattered rays, respectively

Qualitative and Quantitative analysis of graphene alignment by polarized Raman spectroscopy. a Schematic illustration of a typical polarized Raman spectroscopy. Reproduced under the terms of the Creative Commons CC BY license [226].

The Raman spectra of graphene and its derivatives typically possess a D-band (~ 1350 cm−1) arising from vibrations at defects like vacancies, grain boundaries, and heteroatoms within graphene sheets, a G-band (~ 1580 cm−1) associating with in-plane lattice vibration, a shoulder peak D′ (~ 1620 cm−1) which signifies the presence of defects within the graphitic carbon–carbon structure, and second-order peaks of 2D (~ 2700 cm−1) and 2D′ (~ 3240 cm−1) [227, 228]. The absence of the D-band, which denotes the sp3 hybridization, is a common characteristic in chemically vapor deposition-grown or mechanically exfoliated graphene sheets due to the near-flawless nature of the graphitic structure with little defect [170].

The principles underlying Polarized Raman spectroscopy for assessing graphene alignment stem from the intricate interplay between incident light and the lattice structure of graphene, coupled with the resulting vibrational modes [222]. The intensity of Raman signals from graphene is notably influenced by the angle of polarization, demonstrating a robust Raman resonance along the alignment direction parallel to the polarization vector [229]. Polarized Raman spectroscopy enables the excitation of diverse polarization modes of incident light, encompassing those that are parallel (longitudinal) and perpendicular (transverse) to the alignment direction. The interplay between the light polarization and the oriented graphene lattice significantly influences the Raman scattering efficiency, leading to discernible changes in the Raman spectrum. By assessing the Raman signal under various polarization configurations, researchers can extract valuable information about the alignment direction and assess the extent of alignment. Notably, the G-band within the Raman spectrum of graphene serves as a pivotal marker, resonating with the in-plane vibrational motion of carbon atoms in the hexagonal lattice. As graphene sheets align, distinct alterations in the characteristics of the G-band peak can occur, thereby providing valuable information about the alignment state [180]. As an example, the alignment of graphene sheets in their polymer-based composites was qualitatively investigated by Kim and coworkers using polarized Raman spectroscopy [181, 222]. The Raman spectra acquired along different directions showed slight differences in composites characterized by randomly dispersed graphene, as shown in Fig. 18b, c. However, significantly higher G-band intensities were observed when the incident laser aligned with the preferred orientation compared to the configuration perpendicular to the alignment (Fig. 18d, e).

Polarized Raman spectroscopy is also an effective tool for the quantitative assessment of graphene's spatial orientation, as shown in Fig. 18f–m. In Fig. 18f, a Cartesian coordinate system is established, with the X, Y, and Z axes correspondingly defined with respect to the specimen [230]. The laser beam operates along the Z-axis (perpendicular to the specimen's top surface) as well as the X and Y axes (parallel to the surface). Employing fixed polarization configurations, spectra are commonly acquired by systematically rotating specimens to various angles, namely ΦX, ΦY, and ΦZ, to align the laser beams parallel to the corresponding axis. The local direction of graphene sheets within a specimen is defined by denoting the surface normal vector as the z direction and the surface as xy plane, as shown in Fig. 18g [231]. The interrelation between the sample coordinates (X, Y, Z) and the graphene coordinates (x, y, z) is articulated through the Euler angles (θ, φ, ξ), where θ and φ represent the polar angles of the graphene surface normal vector in the sample Cartesian coordinates, and ξ designates the rotation angle of graphene. Illustrating the laboratory coordinates, Fig. 18h elucidates the manner in which specimens are systematically rotated by an angle Φ concerning the laser polarization directions [231].

The presented Fig. 18i, j showcases typical three-dimensional polarized Raman spectra of graphene composites, acquired across various θ values ranging from 0 to 360° in both VV and VH configurations [31]. The G-band intensity within the VV configuration displays cyclic fluctuations, exhibiting peaks and valleys every 90°, while this periodic trend is less discernible within the VH configuration. Consequently, to assess graphene's orientation, it becomes feasible to scrutinize intensities within a representative span of 0–90°, as shown in Fig. 18k. By using the G-band intensity at θ = 0° in the VV configuration as a baseline, the relative Raman intensity (IR) of the G-band within the VV configuration demonstrates a marked reduction as θ increases, signifying a prevalent alignment of graphene in composites along the direction of θ = 0° [31]. In contrast, the relative G-band intensity within the VH showed slight variations against θ, suggesting the absence of preferential orientation.

In cases where the orientation of graphene sheets uniformly changes across a shared plane, the orientation distribution function can be expressed as fN(θ) [230]:

$${f}_{N}\left(\theta \right)=\sum_{i=0}^{\infty }\frac{2i+1}{2}\langle {P}_{i}\left(cos\theta \right)\rangle {P}_{i}\left(cos\theta \right)$$
(8)

where Pi(cosθ) represents the Legendre polynomial of the degree i and \(\langle {P}_{i}({\text{cos}}\theta )\rangle\) denotes the mean value which can be expressed as follows:

$$\langle {P}_{i}\left(cos\theta \right)\rangle =\frac{\int_{\theta =0}^{\theta =\pi }{P}_{i}\left(cos\theta \right){f}_{N}\left(\theta \right)sin\theta d\theta }{\int_{\theta =0}^{\theta =\pi }{f}_{N}\left(\theta \right)sin\theta d\theta }$$
(9)

The Raman scattering intensity of the sample, Isample(Φ), can be determined by the polarization angle Φ in relation to the specimen. This is under the premise of a uniform distribution of the surface normal vectors around the Z-axis:

$${I}_{{\text{sample}}}\left(\Phi \right)={I}_{0}\left\{\frac{8}{15}+\langle {P}_{2}\left(cos\theta \right)\rangle \left(-\frac{16}{21}+{\frac{8}{7}{\text{cos}}}^{2}\Phi \right)+\langle {P}_{4}\left(cos\theta \right)\rangle \left(\frac{8}{35}-\frac{8}{7}{{\text{cos}}}^{2}\Phi {+{\text{cos}}}^{4}\Phi \right)\right\}$$
(10)

where I0 is the amplitude of incident light.For graphene, a non-polar material, the \(\langle {P}_{i}({\text{cos}}\theta )\rangle\) is nonzero only for even I and the polarized Raman spectroscopy can determine \(\langle {P}_{2}({\text{cos}}\theta )\rangle\) and \(\langle {P}_{4}({\text{cos}}\theta )\rangle\). Despite the presence of oxygen-containing groups, the crystal structure of GO and rGO remains largely similar to that of graphene, resulting in analogous vibration modes and Raman spectra [232]. Therefore, the Eq. (9) can also be extended to quantitatively analyze the spatial orientation of both GO and rGO sheets. By fitting the experimental Ramana spectra to Eq. (10), \(\langle {P}_{2}({\text{cos}}\theta )\rangle\) and \(\langle {P}_{4}({\text{cos}}\theta )\rangle\) values can be obtained to quantificationally describe the spatial alignment of graphene in aerogels or composites. It is worth noting that the Eq. (10) is applicable to both the G band and the 2D band of the Raman spectra of materials, yielding consistent outcomes regardless of the polarizability tensors used in the computation [231]. The \(\langle {P}_{2}({\text{cos}}\theta )\rangle\) serves as the primary parameter characterizing the average orientation angle of graphene, while \(\langle {P}_{4}({\text{cos}}\theta )\rangle\) holds less significance [233, 234]. Generally, larger \(\langle {P}_{2}({\text{cos}}\theta )\rangle\) values indicate better alignment of graphene flakes. For instance, \(\langle {P}_{2}({\text{cos}}\theta )\rangle =1\) means a perfect alignment, while \(\langle {P}_{2}({\text{cos}}\theta )\rangle =(3\langle {cos}^{2}\theta \rangle -1)/2=0\) indicates randomly oriented graphene sheets, except in an extremely unique scenario where all graphene sheets are aligned along the Z axis at an angle θ that gives \(\langle {{\text{cos}}}^{2}\theta \rangle =1/3\). In this case, consideration of \(\langle {P}_{4}({\text{cos}}\theta )\rangle\) becomes relevant. To illustrate, a \(\langle {P}_{2}({\text{cos}}\theta )\rangle\) = 0.17 and \(\langle {P}_{4}({\text{cos}}\theta )\rangle\) = 0.05 is reported for graphene paper (Fig. 18l) [231], indicating a subtle long-range nanoscale ordering in graphene paper from a quantitative standpoint.

Utilizing the values of \(\langle {P}_{2}({\text{cos}}\theta )\rangle\) and \(\langle {P}_{4}({\text{cos}}\theta )\rangle\) values obtained using experimental polarized Raman spectra, the orientation distribution function, fN, can be mathematically expressed as follows [235]:

$${f}_{N}\left(\theta \right)=Aexp\left[-\left({\lambda }_{2}{P}_{2}\left(cos\theta \right)+{\lambda }_{4}{P}_{4}\left(cos\theta \right)\right)\right]$$
(11)

where the coefficients A, λ2 and λ4 within the equation can be determined through numerical solutions [235]. A consistent fN(θ) suggests an absence of specific orientation in the graphene structure, while it rapidly diminishes in cases of significant alignment. Applying this principle, Li et al. studied the alignment of carbon flakes in investigated the alignment of carbon flakes in materials like graphene paper and highly-ordered pyrolytic graphite (HOPG), as shown in Fig. 18m [231]. The fN(θ) of HOPG shows a more pronounced decrease with increasing θ compared to that of graphene paper, indicating a more effective alignment of graphene sheets in HOPG. This alignment is further corroborated by SEM observations, as shown in the insets of Fig. 18m [231].

4.3 X-ray Scattering

The X-ray scattering embodies the phenomenon arising from the interaction of X-ray photons with matter, giving rise to the scattering of these photons (Fig. 19a, b) [236, 237]. This process finds its explanation in two fundamental mechanisms: elastic scattering (Rayleigh scattering) and inelastic scattering (Compton scattering and X-ray fluorescence). Elastic scattering, also known as Rayleigh scattering, occurs when X-ray photons interact with the electrons of the atoms in the material without transferring any energy to them. This type of scattering preserves the energy and wavelength of the incident X-rays, but changes their direction. In the realm of elastic X-ray scattering, two distinct techniques stand out: small-angle X-ray scattering (SAXS) and wide-angle X-ray scattering (WAXS). The SAXS examines the scattering of X-rays at small angles (typically 0.1° to 10°) and offers insights into the nanoscale structure, including particle size, shape, and molecular arrangement. WAXS focuses on the scattering of X-rays at larger angles (generally larger than 10°) and provides detailed information about the atomic and molecular arrangement in crystalline materials [238]. The X-rays are scattered in all directions due to the presence of nano structures, resulting in a scattered X-ray pattern. This pattern, a consequence of the interference between scattered X-rays, engenders constructive and destructive interference, thus generating a diffraction pattern amenable to excavate valuable insights regarding the dimensions, geometry, and layout of nanoscale structures within the material [239].

Fig. 19
figure 19

Copyright 2013, Elsevier. Copyright 2016, AIP publishing. Assessment of graphene alignment through SAXS: c 2D SAXS patterns, d the azimuthal dependence of scattering intensity, e orientational distribution coefficient of graphene-based materials fabricated under different magnetic field strengths; and f SAXS analysis of composites featuring diverse graphene alignment. Reproduced under the terms of the PNAS License [245]. Copyright 2017, National Academy of Sciences. Evaluation of graphene alignment using WAXS: g azimuthal scan profiles, h SEM images, and i 2D WAXS patterns of graphene fibers with different stretching ratios. Reproduced with permission [32]. Copyright 2020, Wiley–VCH

a, b Schematic illustrations of SAXS and WAXS techniques. Reproduced with permission [236, 237].

In the case of graphene, SAXS is effective in deciphering the alignment of graphene sheets within the material. By analyzing the scattering patterns at small angles, SAXS can reveal the degree of ordering and orientation of graphene sheets, indicating the presence of aligned structures [240, 241]. Although graphene is not a traditional crystal, it possesses a highly ordered arrangement of carbon atoms within its two-dimensional lattice. On the other hand, WAXS can provide information about the crystal structure, lattice parameters, and interatomic distances in graphene. By examining the scattering patterns at wider angles, WAXS can reveal the orientation and alignment of graphene layers, providing insights into the crystallographic phases and orientation of graphene-based materials [32]. Overall, SAXS provides information about the overall nanoscale structure and alignment of graphene sheets, while WAXS offers insights into the atomic-level arrangement and orientation of graphene layers [241, 242]. By combining SAXS and WAXS techniques, researchers can gain a comprehensive understanding of the alignment of graphene. Considering the spot size of X-ray scattering, SAXS and WAXS provides average information of samples from macro to nano scale.

The quantitative analysis of graphene alignment is facilitated through the utilization of 2D SAXS and WAXS patterns. The extent of alignment of graphene sheets is precisely assessed by establishing a correlation between the azimuthal dependency of the scattering intensity, denoted as I(θ), and an orientational distribution coefficient, O. This correlation is achieved through the application of Hermans’ distribution function [243]:

$$O=\langle \frac{1}{2}\left(3{{\text{cos}}}^{2}\theta -1\right)\rangle$$
(12)

where θ represents azimuthal angle. The orientation distribution coefficient can be correlated with the I(θ) by:

$$\cos^{2} \theta = \frac{{\mathop \smallint \nolimits_{0}^{\pi /2} I\left( \theta \right)\cos^{2} \theta \sin \theta d\theta }}{{\mathop \smallint \nolimits_{0}^{\pi /2} I\left( \theta \right)\sin \theta d\theta }}$$
(13)

By combining Eqs. (12) and (13), the orientational distribution coefficient, O, can be quantified. The O value is ranging from 0 to 1, where O = 1 indicates a perfect alignment of graphene sheets while O = 0 corresponds to a completely random orientation [244].

Lu and coworkers effectively demonstrated the manipulation of graphene alignment through the influence of an external magnetic field, as substantiated by comprehensive SAXS characterizations [245]. At a relatively modest external magnetic field strength of 1 T, the 2D SAXS pattern showcased an elliptical diffusive configuration, while the eccentricity of this elliptical pattern progressively increased at higher field strengths, as shown in Fig. 19c. To quantify this alignment, the orientational distribution coefficient was deduced through a Gaussian approximation fitting of the scattering intensity against azimuthal angle (Fig. 19d). Remarkably, the obtained results in Fig. 19e deliver a consistent raising of the orientational distribution coefficient, evolving from 0.45 at 0.5 T to a heightened value of 0.85 at 6 T. This unequivocally underscores the magnetic field's significant contribution towards enhancing the degree of graphene alignment. To delve deeper, different alignment characteristics of graphene within composites were further investigated utilizing 2D SAXS, as shown in Fig. 19f [245]. The scattering intensity of the pristine polymer exhibited a negligible profile due to the absence of graphene constituents. Conversely, when randomly dispersed, graphene induced a distinct enhancement in scattering intensity, manifesting as a broad and isotropic halo pattern. The vertical- and planar-graphene show anisotropic equatorial and meridional scattering patterns [246], respectively. These outcomes, accompanied by comparable orientational distribution coefficients approximating 0.85, definitively confirm the specific orientations of the respective graphene entities.

The assessment of graphene alignment in materials can also be carried out through the WAXS techniques. Qian et al. reported an orientational distribution coefficient of O = 0.8 in graphene film based on 2D WAXS patterns [9]. Complementing this analytical approach, alignment visualization was further corroborated via SEM characterizations. WAXS studies have also unveiled the alignment tendencies within graphene fibers subjected to varying stretching ratios, as shown in Fig. 19g–i [32]. As the stretching ratio increased, a noticeable reduction in graphene fiber diameter was observed, coinciding with the gradual diminishment of radial wrinkles along the fiber axis (Fig. 19h). Through a comprehensive analysis of the azimuthal angle-dependent scattering intensities in WAXS patterns (Fig. 19g), orientational distribution coefficients, O, were accurately calculated. For fibers subjected to the preparation process without stretching, an O value of 0.64 was recorded (Fig. 19i). Impressively, this value notably escalated to O = 0.86 in graphene fibers characterized by a stretching ratio of 32%. This signifies the contribution of WAXS to the quantitative study of alignment of graphene.

4.4 Comparison of the Characterization Techniques

The three quantitative characterization methods are compared in Table 2. SEM image-based orientation distribution analysis, reliant on SEM images, enables the visual assessment of graphene sheet orientation and alignment. This approach, operating at the microscale, provides local insights into alignment variations within the aerogel but may lack comprehensive coverage. In contrast, polarized Raman spectroscopy, operating at the same scale, offers orientation information through the analysis of anisotropic vibrational modes in graphene layers. However, it also tends to provide localized information. On a broader scale, X-ray scattering techniques, including WAXS and SAXS, emerge as versatile tools capable of quantitatively assessing alignment, crystallinity, and spacing of graphene sheets, spanning from macroscale to nanoscale. The X-ray scattering techniques, which are more comprehensive and quantitative, are highly adaptable, offering insights into structural properties across a wide range of length scales and are thus highly valuable for studying the alignment degree in graphene aerogels in academic research endeavors.

Table 2 Comparisons of different quantitative characterization techniques of the alignment degree

5 Multifunctional Applications

5.1 Energy Conersion and Storage

5.1.1 Solar-Thermal Energy Conversion

Solar-thermal energy conversion and storage harness renewable solar energy, addressing temporal, intensity, and spatial discrepancies between thermal energy demand and supply, and offering a potent means for solar energy utilization. Solar-thermal materials absorb energy from sunlight, raising temperatures during light exposure, and release energy gradually upon cessation, thus achieving effective energy storage and release (Fig. 20a) [261]. Copyright 2020, Elsevier. c, d Thermally annealed aligned graphene aerogel for Li-batteries to hinder the growth of dendrite. Reproduced with permission [263]. Copyright 2021, Elsevier. e–j Comparison of the Li ion diffusion and dendrite growth in vertically aligned, horizontally aligned, and randomly arranged graphene aerogels. Reproduced with permission [265]. Copyright 2020, Elsevier

Highly aligned graphene aerogel-based composites for Li-ion batteries. a Schematic illustration of the Li-ion battery. Reproduced with permission [259].