Abstract
Functional near-infrared spectroscopy (fNIRS), high-density diffuse optical tomography (HD-DOT), and electroencephalography (EEG) are established, cost-effective, and non-invasive neuroimaging techniques, whose integration represents a promising direction for brain activity decoding with high spatiotemporal resolution in naturalistic scenarios. However, robust machine-learning methods for combining these signals remain challenging. In this review, we focus on multimodal fusion methods, emphasizing data-driven unsupervised symmetric techniques, and study their performance on our own HD-fNIRS–EEG data with synthetic ground truth. To this end, we performed a systematic method-oriented survey on fNIRS/DOT–EEG fusion, categorizing works based on fusion strategies, and identifying common artifact removal techniques and integrated auxiliary signals. Our review indicates that while many studies incorporate robust artifact handling for EEG, confounder correction in fNIRS remains limited to filtering or motion removal. Moreover, short-separation measurements and other auxiliary signals for fNIRS remain underutilized. Fusion methods predominantly rely on data concatenation, model-based, or decision-level strategies, while source-decomposition techniques are underrepresented, despite their potential for revealing more complex latent neurovascular coupling processes. To address the scarcity of multimodal public datasets, we generated a realistic synthetic HD-fNIRS–EEG dataset that simulates a finger tapping motor task, with concurrent suppression of EEG alpha-band power and an increase in hemoglobin in fNIRS from a shared neuronal source. We illustrate a proof-of-concept comparison of some source-decomposition methods on this dataset and provide the full implementations and an example Jupyter notebook to reproduce and extend these results.
Keywords: functional near-infrared spectroscopy (fNIRS), diffuse optical tomography (DOT), electroencephalography (EEG), multimodal fusion, source decomposition
1. Introduction: Motivation and Scope
Human neuroscience and neurotechnology are on the verge of a significant transformation, moving from conventional laboratory settings toward embracing the complexity of natural environments (Park et al., 2018; Pinti et al., 2020; Schmuckler, 2001; Sonkusare et al., 2019; von Lühmann et al., 2021). This shift toward an ecologically valid depiction of human brain function promises to provide new scientific insights and breakthroughs in our understanding of neuronal development, health, and aging, and drive translation in medicine and psychiatry. However, substantial interdisciplinary challenges need to be overcome as these settings elude many of the conventional methods, sensors, and parametric experimental paradigms currently available. To better understand the complex interactions between brain networks, the body, and the environment, it is paramount to develop integrative multimodal platforms that can continuously monitor the embodied brain (Kiverstein & Miller, 2015; Scholkmann et al., 2022) and discover and model the intertwined complex interactions between physiology, behavior, and cognition in everyday settings.
1.1. EEG and fNIRS
Wearable functional near-infrared spectroscopy (fNIRS), high-density diffuse optical tomography (HD-DOT), and electroencephalography (EEG) are technologies that offer such a pathway. fNIRS and EEG are established, safe, and cost-effective non-invasive neuroimaging techniques that allow brain imaging under naturalistic settings. In recent years, wearable fiberless continuous wave (CW) fNIRS systems and wearable EEG have made rapid progress (Ayaz et al., 2012; Casson et al., 2010; Niso et al., 2023; Pinti et al., 2018; Vidal-Rosas et al., 2023; von Lühmann et al., 2021; Zhao & Cooper, 2017). EEG and fNIRS complement each other by combining EEG’s ability to capture synchronous neuro-electrical activity with millisecond time scale and fNIRS’ ability to measure slow changes in cerebral blood flow with better spatial localization. Both measures, while sensitive to different types of physiology and measurement artifacts, are linked via neurovascular coupling (NVC) (T. Huppert et al., 2017; Keles et al., 2016; Logothetis, 2008), which relates transient neural activity to subsequent hemodynamic changes, the latter corresponding to BOLD signals in fMRI. Consequently, multimodal neuroimaging with EEG–fNIRS is highly promising for many research disciplines such as brain–computer interfaces (BCI) (Fazli et al., 2012; Hong et al., 2018; von Lühmann, 2018), clinical neurology (Abtahi et al., 2020; Schneider et al., 2015; Sirpal et al., 2019; Wallois et al., 2010) and neurorehabilitation (Berger et al., 2019; Mihara & Miyai, 2016). Wearable EEG and fNIRS instruments and first combined fNIRS–EEG devices are increasingly available (Safaie et al., 2013; Uchitel et al., 2021; von Lühmann et al., 2017). Furthermore, to realize the full potential of fNIRS, high-density (HD) arrays of multiple source–detector separations (commonly less than 15 mm to 4 cm) with overlapping sensitivity distributions can be employed to enable three-dimensional image reconstruction of functional activation (Arridge & Schotland, 2009; Bluestone et al., 2001; Boas et al., 2001; Wheelock et al., 2019; Zeff et al., 2007), referred to as high-density diffuse optical tomography (HD-DOT). Recent HD-DOT studies have achieved the same spatial resolution of the entire surface of the brain as the latest fMRI-based Human Connectome Parcellation scheme (Habermehl et al., 2012; White & Culver, 2010), enabling connectivity analysis or decoding performance comparable with fMRI (Eggebrecht et al., 2014; Ferradal et al., 2016; Tang et al., 2023). Complemented by recent advances in inferring subcortical from cortical network activity (Pang et al., 2023), fNIRS-based HD-DOT is now a viable extension of fMRI (Anaya et al., 2023; Vidal-Rosas et al., 2023) for the everyday world, which can offer the opportunity for seamless integration with EEG.
1.2. Artifacts
Despite their great potential upon fusion, it is a highly non-trivial challenge to distinguish evoked neuronal from complex systemic physiological activity in fNIRS (T. J. Huppert, 2016; Scholkmann et al., 2022; Yücel et al., 2015), and EEG (Parra et al., 2005; Urigüen & Garcia-Zapirain, 2015) already in well-controlled conventional experiments, a problem that is greatly amplified by natural behavior and motion. fNIRS and EEG are contaminated by hemodynamic and bioelectric signals from systemic, non-neural (extracerebral) physiology, which can mask the underlying neural activation and reduce contrast. While many of these confounder signals, such as cardiac, respiratory, muscle, and blood pressure changes, affect both modalities, they manifest with distinct temporal, spatial, and amplitude characteristics in each modality. More precisely, while indeed the same physiological source can originate artifacts in both modalities, it leads to a very different vascular and bio-electrical manifestation (like cardiac activity leading to electrical ECG spikes and hemodynamic pulse waves). Given their different physical measurement principles, both modalities are “blind” to each other’s case, making their combined analysis complementary. In EEG, contamination is primarily caused by ocular activity (EOG) and head and neck muscle activity (EMG), which increase variance across much of the relevant EEG spectrum. Neuronal and non-neural field potentials (e.g., EOG) traverse the tissue quasi-instantaneously. This is not true for fNIRS. In fNIRS, the contamination comes from the scalp, brain motion in the cerebro-spinal fluid (CSF), and the brain vasculature (T. J. Huppert, 2016; Scholkmann et al., 2022; Yücel et al., 2015, 2021). Hemodynamics and blood flow are affected by physical activity, induced and spontaneous blood pressure changes, breathing, heart rate and partial pressure of CO2, and can have a major impact on task-evoked and resting-state fNIRS signals (Franceschini et al., 2006; Funane et al., 2014; Goodwin et al., 2014; Kirilina et al., 2012; Mesquita et al., 2010). Due to the physiological complexity of hemodynamics in vasculature, channels measure a unique mixture of global and local physiological and neuronal activity, both evoked by and independent from stimuli and behavior. In contrast to EEG, these different hemodynamic signal components are non-instantaneously coupled into measurement channels with spatially and behaviorally (e.g., motion) dependent non-stationary delays and morphology as they disperse along the arteriole system and undergo vasomodulation. The shared latent neurovascular link between fNIRS and EEG, the fact that their instruments can be wearable, and the complementary nature of their physiology, artifacts, and noise make EEG and fNIRS promising candidates for fusion-based multimodal neuroimaging in general, and in the everyday world in particular.
1.3. Data-driven approaches
While model/regression-based methods such as the General Linear Model (GLM) in fNIRS offer well-established approaches for the analysis of conventional parametric fNIRS and EEG experiments, they usually fall short when (1) precise knowledge about stimuli and timing is not available or (2) latent complex physiological relationships and coupling mechanisms are unknown. Data-driven approaches based on machine learning can help alleviate this problem (von Lühmann, Li, Müller, et al., 2020; von Lühmann et al., 2019) and have seen a steep rise in the last decade (Zhuang et al., 2020). However, there is still a significant lack of established data-driven methods that combine fNIRS/DOT and EEG to harness unsupervised symmetric (bidirectional) multimodal sensor fusion to robustly model, identify, and extract co-modulation or reject physiological confounders. As our review of the existing literature on EEG–fNIRS fusion will reveal, most existing methods either require supervision through a priori labeling or modeling or do not fuse information bidirectionally on the sensor level, where physiological interactions and confounders can best be taken into consideration. We consider that this lack of well-established methods for unsupervised fusion, which is independent of labels or stimulus timing, is particularly problematic for (1) the analysis of data from continuous brain imaging in naturalistic environments, (2) the assessment of (dynamic) brain network activity in functional connectivity analysis, and (3) more advanced preprocessing in hybrid EEG–fNIRS single trial analysis, for example, for BCI. This set of problems fully motivates the focus of this review.
1.4. Related reviews
Due to its great promise, multimodal EEG–fNIRS integration has been reviewed already in other occasions from different perspectives. We identified 10 methods-oriented review articles, all except one focusing on specific domains of application. In the context of hybrid BCI, Ahn and Jun (2017) surveyed studies implementing EEG–fNIRS fusion methods for general BCI purposes, J. Chen et al. (2023) reviewed previous works for motor rehabilitation, Hong et al. (2018) covered feature extraction and classification algorithms for locked-in syndrome, H. Khan et al. (2021) analyzed articles focusing on human gait, and AlQahtani et al. (2024) focused on lower prosthetic limbs for rehabilitation. Other method-oriented reviews surveyed articles utilizing EEG–fNIRS fusion methods in the context of robot-assisted gait rehabilitation (Berger et al., 2019), diverse subdomains of neuroergonomics (Bourguignon et al., 2022), neurovascular coupling analysis (Yeung & Chu, 2022), and high-frequency cerebral physiology analysis (Vakitbilir et al., 2024). A general-purpose methodology-focus review was presented in R. Li et al. (2022), where the authors conducted an exhaustive literature search ending up with 92 selected studies. The latter were separated into three categories depending on the fusion strategy: EEG-informed fNIRS analyses, fNIRS-informed EEG analyses, and parallel fNIRS–EEG analyses. These studies, covered in R. Li et al. (2022)’s work, alongside their categorization, were also considered in our survey.
1.5. Motivation and scope
While the reviews mentioned above provide a thorough and valuable overview of many EEG–fNIRS fusion approaches, a particular domain of methods has barely received attention (see categorization in the next section and Fig. 2): Unsupervised data-driven methods for parallel/symmetric fNIRS–EEG fusion. We hypothesize that these methods are particularly valuable for the analysis and decoding of data from continuous brain imaging in naturalistic everyday-world environments, and in the presence of physiological artifacts. To fill this gap, we focus our work on this domain, updating and complementing previous reviews with a perspective on source-decomposition multimodal fusion methods. Furthermore, due to the community’s general interest in physiological interpretability, we restrict our study to classical machine-learning (ML) methods, as interpretation of their deep-learning (DL) counterparts is still in large parts an ongoing effort (Eastmond et al., 2022). With the sole purpose of distinguishing these methods and then excluding the DL ones, we define the latter as neural networks with two or more non-linear layers. We emphasize that such a definition plays a purely practical role, and should not be taken as a rigorous, global definition of the concept. For a comprehensive review of DL methods for fNIRS/DOT, we refer the reader to our recent article (Dissanayake et al., 2025). On top of our method-oriented survey, we also perform a qualitative and quantitative comparison of what we consider are the most promising data-driven unsupervised symmetric sensor fusion methods for multimodal source decomposition. Their performance is assessed on a realistic HD-fNIRS–EEG data with synthetic ground truth developed for this purpose. We make all compared source-decomposition methods available to the community in our Python-based fNIRS/DOT analysis toolbox Cedalion (IBS-Lab & Cedalion-Developers, 2024),1 and we provide a link to an example notebook2 to enable researchers to reproduce and explore our results based on the same synthetic data (see Supplementary Materials).
Fig. 2.
Taxonomy of EEG–fNIRS fusion strategies. Top: Organization of the nine categories into a vertical stacked bar plot, where bar thickness reflects the percentage of articles from our survey that used each fusion strategy. The right/left arrows on each box represent the directionality of the method, namely whether they are modality symmetric (left and right), EEG-informed (right), fNIRS-informed (left), or non-directional (none). Bottom: Table of category features, relevant for single-trial applications in continuous brain imaging. As indicated by the arrow, in this review, we are particularly focused on source-decomposition methods, since they contain all those desired properties. In both subfigures, colored categories represent the most promising fusion strategies for capturing more complex relationships due to their flexibility and early fusion implementations.
It is worth noting that our emphasis on source-decomposition methods relative to model-based, asymmetric, or non-directional approaches arises mainly from our perspective on continuous naturalistic brain imaging (von Lühmann et al., 2021) and should not be construed as a general endorsement of these techniques as the optimal machine-learning solutions for all neuroscience applications.
1.6. Review structure
The rest of the paper is organized as follows: In Section 2, we provide a taxonomy and categorization of fNIRS–EEG fusion methods. In Section 3, we present the results of our method-oriented survey of fusion approaches with a particular focus on data-driven unsupervised symmetric methods. In Section 4, we introduce our realistic simulated HD-fNIRS–EEG data and use it to compare the performance of some source-decomposition methods. In Section 5, we conclude the review with a discussion of current challenges, limitations, and future directions.
2. Taxonomy and Overview of Fusion Strategies
2.1. EEG–fNIRS fusion
Multimodal data offer a significant advantage over single modalities when it provides complementary information of the same underlying phenomena. Fusion methods play a crucial role in enhancing the interpretation and utility of such multimodal data by leveraging individual strengths, while mitigating the effect of modality-specific confounding signals. In particular, EEG–fNIRS fusion capitalizes on the temporal accuracy of EEG in detecting electrical signals and the spatial accuracy of fNIRS in locating these signals within specific brain regions. Additionally, due to the different nature of the acquisition principles, most artifacts are not shared between modalities. Consequently, EEG–fNIRS fusion methods are promising data analysis techniques for NVC investigation, and brain activity decoding with improved spatiotemporal and signal-to-noise ratio (SNR) when compared with their individual application.
2.2. Fusion-level taxonomy
Fusion methods can be classified by their location along the data analysis pipeline, leading to the following categories: Data-Level Fusion (Fig. 1a) happens at the beginning of the pipeline, combining raw, or very low-processed data.3 They tend to capture a more detailed picture of brain activity by exploiting modality features that might be lost in later stages of processing. Typical challenges include handling the inherent differences between the nature of the datasets, synchronizing and aligning data, and dealing with high-dimensionality problems. Feature-Level Fusion (Fig. 1b) takes place after single-modality features are generated individually. The unimodal prior steps can alleviate high-dimensionality and data-incompatibility problems, by bringing datasets into a common (low-dimensional) feature space. However, loss of relevant information may occur at this step. In Decision-Level Fusion (Fig. 1c), independent decisions derived from single modalities are aggregated to produce a final output. They can be useful when each modality assesses an independent aspect of brain activity, but they can also fail to exploit complementary information due to late fusion. On top of these three cases, in practice, we often encounter hybrid strategies, where each modality enters the fusion at a different stage (Fig. 1d left), or cases in which modalities are combined at multiple stages (Fig. 1d middle and right).
Fig. 1.
Stages or levels at which fusion methods can be integrated into the data analysis pipeline: X and Y represent EEG and fNIRS-derived multivariate time series, respectively, including raw sensor signals. The U, F, and D nodes represent unimodal processes, the proper fusion methods, and decision algorithms, respectively. The gray texts provide some concrete examples of these processes.
2.3. Model-based and data-driven
In terms of functionality, EEG–fNIRS fusion methods can be further distinguished between model-based versus data-driven, and asymmetrical versus symmetrical, leading to the nine categories depicted in Figure 2. Model-based algorithms utilize theoretical models and a priori knowledge to parameterize an underlying phenomenon while data-driven methods leverage data to uncover patterns, primarily through statistical techniques. When an accurate theoretical representation is used, model-based methods can outperform data-driven ones in terms of interpretability and precision, while for incomplete or inapplicable models, data-driven methods offer more flexibility and adaptability to learn complex patterns from the data (von Lühmann, Li, Müller, et al., 2020; von Lühmann et al., 2019).
2.4. Asymmetric, symmetric, and non-directional
Asymmetric methods make use of one of the modalities to enhance the signal quality of the other (R. Li et al., 2022). In these unilateral approaches, the former modality does not enjoy the benefits of multimodality. EEG-informed algorithms exploit EEG high temporal resolution to guide fNIRS processing, improving the quality of the reconstructed hemodynamic response (Y. Gao, Jia, et al., 2023; Peng et al., 2014; Pouliot et al., 2014). fNIRS-informed strategies, in contrast, leverage fNIRS higher spatial resolution to restrict EEG analysis to specific regions of interest in channel (channel selection) (R. Li et al., 2017; Moslehi & Davies, 2021) or volume space (source reconstruction) (Cao et al., 2021; R. Li et al., 2019). The latter typically makes use of head models for better source localization (Aihara et al., 2012), which increases computational and memory demands. From our survey, we found that most EEG-informed fusion methods were model-based and supervised (labeled data), via GLM-based fNIRS analysis, while fNIRS-informed algorithms were typically unsupervised (unlabeled data).
In EEG–fNIRS symmetric methods, both modalities inform and are informed by the other. They can be divided into model-based or data-driven, depending on the amount of prior knowledge incorporated during fusion. The former category is further split depending on whether the method contains parameters that must be learned during training (learnable) (Y. Gao, Liu, et al., 2023; Sood et al., 2016) or whether they are static, “handcrafted” functions (a priori) (Aghajani et al., 2017; Giacometti & Diamond, 2014; Othman et al., 2021). Such a priori knowledge is typically derived from human expertise rather than data, making the model less flexible and hence less capable of capturing single-trial variability. This is the case of many NVC-like features, which combine modalities via a pre-defined function, such as correlation (L.-C. Chen et al., 2015), that models the relationship. Symmetric data-driven approaches are further separated into data selection and source-decomposition methods. The former focuses on selecting (without modifying) a subset of the most informative input data (Deligani et al., 2021; Hasan et al., 2020). Source-decomposition fusion methods (Zhuang et al., 2020), also known as latent component analysis, assume each dataset is generated from a mixture of latent (unobserved) sources that must be identified by the algorithm and are assumed to co-modulate between modalities. Apart from the choice of co-modulation function, which defines the algorithm (Witten et al., 2009), these highly flexible and unsupervised methods are purely data-driven, making them ideal for capturing complex structures for which no good a priori model is known (Dähne et al., 2013; Dashtestani et al., 2022).
On top of these categories, we have data concatenation (Almajidy et al., 2015; Cao et al., 2022) and decision-level fusion (Fazli et al., 2012; Pfurtscheller, 2010; Shin, Kim, et al., 2018), corresponding to non-directional methods that take both modalities as inputs but, unlike directional approaches, they do not use the output to inform any of the modalities during processing. The former strategy concatenates features into a unified vector and then passes it as input of a decision-making algorithm, while the latter makes independent single-modality decisions and then combines them to produce a final output. Both methods are simple to implement and computationally cheap, but they may struggle to capture the underlying complex relationships between modalities.
2.5. Relevance for continuous brain imaging
Such a taxonomy allows us to identify what we consider are the most promising EEG–fNIRS fusion strategies for capturing more complex relationships between modalities (colored in Fig. 2) due to their flexibility and early fusion implementations. On top of these features, source-decomposition methods are also data-driven, and unsupervised (Fig. 2 bottom), making them good candidates for successful single-trial applications in continuous brain imaging data, for example, from the everyday world. Despite the potential benefits of the flexible methods (colored) over the others (gray), the former seem to be underrepresented in current studies according to our survey (Fig. 2 top), where more than half of the articles implement selection, model-based a priori, or non-directional approaches. In particular, most of the BCI-related articles utilize some of these inflexible methods.
3. Survey of Fusion Methods
3.1. Preliminaries
3.1.1. Methodology
The survey was conducted using PubMed and Web of Science (WoS) databases, comprising articles until March 25, 2025, with no lower limit on publishing date (Fig. 3). We searched for English articles containing the combination of keywords (“fNIRS” OR “NIRS” OR “functional near-infrared spectroscopy” OR “near-infrared spectroscopy”) AND (“EEG” OR “electroencephalography”) in the title OR abstract, obtaining 1032 and 1134 results for PubMed and WoS, respectively. Eliminating duplicates, we ended up with 1280 articles. As a preliminary automatic filter, we excluded preprints, comment articles, retractions, and erratum, and separated apart for later inspection 328 reviews, leading us to 883 published articles. From this subset, we initiated a manual filtering via title and abstract inspection, excluding articles that involved non-human subjects, were dataset or instrumentation oriented, or did not use fusion methods but rather unimodal analyses in parallel. This manual screening returned 190 articles, including 41 using DL-based methods. The remaining 149 documents went through a full-text reading procedure, revealing 25 additional nonrelevant articles, according to the criteria mentioned above. The final selection consisted of 124 non-duplicated, journal-published articles involving EEG–fNIRS classical ML fusion methods, to which we restrict all further analyses in this review. Publishing dates range from April 2010 (Pfurtscheller, 2010) up to March 2025 (Saricaoglu et al., 2025), most of them from Q1 and Q2 journals, such as IEEE TNSRE, IEEE TBME, IEEE ACCESS, Nature Scientific Reports, NeuroImage, Frontiers in Neuroscience, and MDPI Sensors. The 328 separated review articles were scrapped via manual title/abstract filtering ending up with 10 methodology-focus EEG–fNIRS fusion-method reviews, which were already introduced in Section 1. During this search, we found no novel EEG–fNIRS fusion methods that had not been included already from the previous 883 surveyed articles.
Fig. 3.

Flow diagram of the conducted survey and distribution of research areas for the selected 124 articles using classical ML-based EEG–fNIRS fusion methods.
3.1.2. DOT-EEG
As diffuse optical tomography (DOT) is a field increasingly on the rise, we also conducted a survey on DOT–EEG fusion methods in a similar fashion, using the same search engines and publishing date ranges. We searched for English articles with the keyword combination (“DOT” OR “diffuse optical tomography” OR “diffuse optical imaging”) AND (“EEG” OR “electroencephalography”) in the title OR abstract. We found 243 and 319 documents in PubMed and Wos, respectively, which, upon deduplication, reduced to 367, including 15 review articles. We then performed a manual filtering via title and abstract inspection, using the same criteria as for the fNIRS–EEG survey, ending up with 5 articles using classical ML-based DOT–EEG fusion methods, and no relevant reviews.4
3.1.3. Research areas
The survey revealed a diverse range of applications for EEG–fNIRS fusion methods, which allowed us to group the 124 articles into 4 categories, whose distribution is shown in Figure 3. From the EEG–DOT fusion methods survey, out of the five selected articles, we found one in the area of clinical assessment, two in NVC investigation, and two in method presentation. Single trial brain decoding encompasses more than half the survey volume, containing studies that investigate brain decoding during experimental tasks such as motor, mental workload, and cognitive load, most of them focusing on hybrid BCI for brain state classification. Articles in the clinical assessment category make use of EEG–fNIRS systems for diagnosing, monitoring, and predicting various neurological and psychiatric conditions such as epilepsy, ADHD, depression, and Alzheimer’s Disease (AD). NVC investigation contains publications that exploit simultaneous EEG and fNIRS recordings to model or perform a quantitative analysis of NVC, sometimes through functional connectivity analysis (FCA). Method presentation groups all studies whose main goal and contribution are the development of a new data analysis method.
3.1.4. Survey results and implications for continuous brain imaging
Before diving into details, we summarize our survey’s main results and their relevance for the main topic of the review, continuous brain imaging. Regarding data preprocessing (Section 3.2), the following conclusions were drawn (Fig. 4):
Fig. 4.
Distribution of artifact removal techniques (top) and auxiliary signals (bottom) used for EEG (left) and fNIRS (right) from surveyed studies. The category “Others” contains non-specified artifact removal methods, as well as algorithms that were found only once during the survey.
Most studies using fNIRS/DOT–EEG fusion methods do not sufficiently address physiological confounders in fNIRS/DOT. Instead, they typically incorporate simple channel pruning and filtering techniques or handle motion artifacts only. In contrast, most of the studies do incorporate confounder removal techniques for EEG, such as EOG, EMG, or ECG regression.
Most studies using fNIRS/DOT–EEG fusion methods do not incorporate auxiliary signals, and when they do, they are mostly used to track EEG-related artifacts such as EOG or EMG, but not for monitoring fNIRS confounders or artifacts. In particular, short-separated (SS) channels are rarely recorded.5
These findings reveal a great potential for the neuroimaging community to improve fNIRS/DOT–EEG fusion methods interpretability and reliability by comprehensively addressing physiological confounders and artifact removal in fNIRS data.
The taxonomy introduced in Section 2 allowed us to group all surveyed EEG–fNIRS/DOT fusion strategies into the nine categories of Figure 2. There, one can also see the fNIRS–EEG methods distribution (while from the DOT–EEG survey, we found two and three articles in the “Model-Based: A priori” and “fNIRS-Informed: Source Reconstruction” categories, respectively). The reason for such a customized taxonomy comes from the need to identify what we consider are the most promising fusion methods for capturing complex relationships between modalities, namely EEG-informed (Section 3.3), fNIRS-informed: Source Reconstruction (Section 3.4), Model-based: Learnable (Section 3.5), and Source Decomposition (Section 3.6). Articles in these categories will be described in detail in the corresponding sections, while the works using nondirectional strategies are given in Table A1 (Concatenation) and Table A2 (Decision-Level) in Appendix A. Here we summarize the survey’s results regarding fusion methods as follows (Fig. 2):
Most of the articles utilize only data concatenation (20%), model-based: a priori (20%), or decision-level (16%) approaches. The remaining articles utilize mostly asymmetrical (16%) or selection (11.2%) algorithms.
The domain of source-decomposition algorithms (13.6%) seems to be underrepresented. The approaches we identified (such as the CCA and jICA methods family, and mSPoC) do not address/remove physiology in fNIRS before or during fusion.
While these findings do not represent a problem for conventional neuroscience applications, we do believe further work is needed to tackle the challenging scenarios that arise from naturalistic brain imaging. From this perspective, the most utilized methods so far may not be enough to uncover complex patterns due to their non-directional nature, lack of learnable parameters, or late fusion. Moreover, while asymmetrical methods have been successfully applied to capture inter-trial variability (Section 3.3.1) or achieve higher spatiotemporal resolution than each individual modality (Section 3.4.1), within them, one of the modalities does not fully profit from fusion, or, in other words, not both modalities inform each other. We believe that such asymmetrical treatment represents a sub-optimal solution that can be cured by moving forward to bi-directional data-driven methods, such as multimodal source-decomposition methods. Consequently, we argue that the latter deserve more attention, due to their great potential to better find shared NVC latent processes, while also simultaneously rejecting physiology and motion artifacts.
In light of these findings and the review’s perspective, we ranked the source-decomposition methods identified in our survey by evaluating their performance using the same synthetic dataset. While a more comprehensive analysis will be conducted in subsequent work, the current results provide first valuable insights into each method’s specific characteristics, strengths, and limitations.
3.2. Artifact removal techniques and auxiliary signals
EEG and fNIRS data are both susceptible to artifacts that can compromise signal quality and contrast-to-noise ratio, with EEG being influenced by electrophysiological noise (e.g., eye blinks, muscle activity) and external electromagnetic interference, and fNIRS being affected by cardiovascular hemodynamics and ambient optical noise. Motion exacerbates artifacts in both modalities, making advanced preprocessing techniques (even prior to fusion) necessary for reliable neuroimaging in naturalistic environments. Single-modality preprocessing typically involves channel pruning, bandpass filtering (targeting brain state frequencies in EEG and hemodynamic responses in fNIRS), re-referencing (EEG), converting light intensity to hemodynamic changes by using the modified Beer-Lambert law (fNIRS), detrending, and normalization. Within such preprocessing pipelines, we are particularly interested in artifact removal techniques and the use of auxiliary physiological signals. For EEG (Islam et al., 2016), regression methods and ICA are typically used to remove EOG, EMG, and ECG interference, while fNIRS implements methods such as wavelet filtering (Chiarelli et al., 2015; Molavi & Dumont, 2012), Principal Component Analysis (PCA) (X. Zhang et al., 2016), and spline interpolation (Scholkmann et al., 2010) to reduce motion artifacts, and more sophisticated algorithms, such as the General Linear Model (GLM) (Diamond et al., 2006; Kirilina et al., 2013) and (when SS channels are available) SS-channel regression (Gagnon et al., 2011; Yücel et al., 2015), to model physiological noise. Additionally, both modalities benefit from spatial filtering, where multi-sensor signals are combined to enhance brain activity components.
Auxiliary signals are often incorporated into noise reduction algorithms to improve contrast. For EEG (Islam et al., 2016), common examples include EOG and EMG to reduce ocular and muscular artifacts, and ECG to track cardiac interference. For fNIRS (Scholkmann et al., 2022; Tachtsidis & Scholkmann, 2016), heart rate is commonly used to track systemic cardiac activity, photoplethysmography (PPG) and blood pressure to monitor vascular dynamics, and slower oscillations such as Mayer waves, respiration to account for breathing-related fluctuations, and SpO2 (oxygen saturation) to assess systemic oxygenation changes. Additionally, SS channels can be incorporated to capture global systemic, non-cerebral hemodynamic activity. In both modalities, accelerometers and gyroscopes help track motion and posture changes, thereby aiding in movement artifact identification and reduction.
EEG–fNIRS survey results
We investigated the use of auxiliary signals and artifact removal techniques on EEG and fNIRS data, beyond channel pruning, filtering, and normalization, arriving at an evident asymmetrical treatment between modalities (Fig. 4). Out of the 124 studies, 81 of them incorporated artifact removal algorithms for EEG data, versus only 43 for fNIRS. Among the latter, 21 employed motion-artifact removal, 12 tackled physiological confounders, and just 10 of studies handled both type of artifacts. EEG artifact removal was heavily dominated by ICA and its variants (60 articles), used for reducing the effect of cardiac, ocular, and muscular confounding signals, followed by other EOG-regression algorithms, Laplacian filtering for muscle contamination (Fitzgibbon et al., 2013), and Artifact Subspace Reconstruction (ASR) (Chang et al., 2020). For fNIRS, in contrast, the subset of articles that did apply noise canceling methods follows a more uniform distribution compared with EEG, signaling a lack of consensus on “best practices” for removing artifacts (Yücel et al., 2021, 2024). For motion artifacts, PCA, wavelet filtering, spline interpolation, and Temporal Derivative Distribution Repair (TDDR) (Fishburn et al., 2019) were the preferred approaches, while GLM and SS channel subtraction were typically utilized to remove physiological noise. The category “Others” encompasses methods used only once, such as Chauvenet’s criterion (T. Talukdar et al., 2013), Savitzky-Golay filtering (Jahani et al., 2018), spatial filtering, and Variational Mode Decomposition (VMD) (Hossain et al., 2022).
Regarding auxiliary signals, we found that 81 articles did not incorporate any of them, 36 included EEG-related ones, such as EOG and EMG, and only 12 recorded fNIRS-specific signals, such as SS channels and SpO2. In some cases, auxiliary signals were incorporated during single-modality preprocessing, and in others they entered as inputs of decision-level algorithms on the same footing as the neuroimaging recordings.
3.2.2. EEG-DOT survey results
From the five surveyed EEG–DOT fusion articles, two of them used simulated data and one did not eliminate artifacts nor included auxiliary signals. Chalia et al. (2016) included ECG monitoring to remove EEG artifacts manually via a clinical neurophysiologist, and DOT motion artifacts were handled by combining visual inspection and the Homer2 toolbox (T. J. Huppert et al., 2009). Cao et al. (2023) used ICA with PCA for EEG denoising, and TDDR with GLM (incorporating recorded gyroscope and accelerometer signals as regressors) to remove motion-induced and physiological noise from DOT data.
3.3. EEG-informed fNIRS analysis
Some EEG–fNIRS fusion techniques treat modalities in asymmetrical ways, leveraging one modality to inform the other, therefore, significantly enhancing signal quality and interpretability compared with non-directional fusion or single-modality approaches. A drawback of these unilateral approaches is that the former modality does not enjoy the benefits of multimodal fusion. In EEG-informed fNIRS analysis, EEG features are utilized to guide or restrict fNIRS processing in order to improve the quality of the reconstructed hemodynamic response. One of the most successful implementations of the EEG-informed methods consists of using EEG features to build regressors for an fNIRS GLM (Fig. 5), where fNIRS hemodynamic signals (HbO and HbR) are treated as the dependent variable (Peng et al., 2014; Pouliot et al., 2014).
Fig. 5.
GLM fNIRS analysis with standard regressors (top) and one of its EEG-informed variants (bottom). In the latter, the peak (A) and delay (T) are derived from frequency-specific EEG bandpower signals and used to scale and shift the standard boxcar function of each trial. Additionally, the canonical HRF is replaced by a function with subject-specific parameters, such as time-to-peak (a) or time-to-undershoot (c). The resulting hemodynamic response design matrix (G) captures trial-to-trial and inter-subject variability.
3.3.1. fNIRS GLM analysis with EEG-derived regressors
fNIRS GLM analysis
The GLM is a widely used technique for fNIRS analysis. Denoting number of fNIRS channels and sample points by and , respectively, GLM expresses the measured oxygenated or deoxygenated hemodynamic signal, , as a linear combination of the predicted hemodynamic responses, additional confounding regressors, and a residual
| (1) |
Here, is the residual error matrix, and are the to-be-determined coefficients of the regressors encoded in the rows of the design matrix . The latter can be divided into task/condition-specific evoked hemodynamic responses, and nuisance regressors utilized for removing physiological and motion artifacts, such as polynomial drifts and features derived from auxiliary signals. In the GLM standard formulation (Diamond et al., 2006; Kirilina et al., 2013), the hemodynamic response is modeled as the convolution between a canonical hemodynamic response function (HRF) and a boxcar function. The HRF parameterizes the hemodynamic shape in response to a short stimulation and is typically modeled as combination of gamma-variant (A. F. Abdelnour & Huppert, 2009; Glover, 1999) or Gaussian functions. The boxcar function encodes the temporal structure of the corresponding task-evoked stimuli.
The optimal coefficients can be estimated in various forms. Assuming uncorrelated residual, and a non-singular matrix (implying, in particular, ), the Ordinary Least Squares (OLS) estimate (1) exists and is unique. Other alternative estimates of include regularization terms, such as Ridge regression to handle collinearity, or iterative re-weighting techniques to deal with outliers and correlated noise, such as Iteratively Reweighted Least Squares (IRLS) (Holland & Welsch, 1977) and Autoregressive IRLS (AR-IRLS) (Barker et al., 2013), which extends IRLS by prewhitening with an AR noise model. Ultimately, statistical tests, such as t-test, can be performed on to identify channels encoding statistically significant differences among conditions.
GLM with EEG-derived regressors
In the simplest GLM formulation, the same HRF shape and boxcar function are assumed across brain regions, subjects, and trials (Fig. 5 top). While this approximation to a stationary brain response makes the algorithm simpler, in reality these variabilities are present (Handwerker et al., 2004; T. Huppert et al., 2006), and so the method may not accurately describe the underlying cerebrovascular activity. Parametric modeling of HRFs and the introduction of EEG-derived boxcar functions have proven to be promising modifications to achieve more favorable performance (Fig. 5 bottom). Features from frequency-specific time-varying EEG bandpower, such as peak and latency with respect to stimulus onset, can be used for the latter purpose (Y. Gao, Jia, et al., 2023). Specifically, the peak delay shifts the start time of the boxcar, and the peak value scales its amplitude. This approach is motivated by the well-established fact that EEG activity systematically changes with specific stimuli, observed either as phase-locked responses (event-related potentials, ERPs) or as induced modulations of oscillatory power (event-related desynchronization/synchronization, ERD/ERS) (Pfurtscheller & Lopes Da Silva, 1999). On top of that, the canonical HRF can be replaced by a parametric double gamma function, optimized with subject-specific parameters, such as time-to-peak or time-to-overshoot (J. Lin et al., 2024).
3.3.2. Survey results
GLM-based
From the DOT-EEG survey, we found no articles following the EEG-informed strategy, while the fNIRS–EEG survey returned 12 works (9.6%) (Fig. 2), in which 9 of them implemented similar GLM-based frameworks to the one presented above (Table 1). The first studies to introduce the EEG-informed idea for fNIRS analysis were Peng et al. (2014, 2016) and Pouliot et al. (2014), with the purpose of studying brain activity during epileptic events. There, onset of seizures or interictal epileptiform discharges (IEDs or spikes) were marked manually from EEG time traces by professionals and then convolved with canonical HRFs to generate the GLM regressors. In Y. Gao, Jia, et al. (2023), the same EEG bandpower peak-and-latency strategy mentioned above was used. After regression, the fitted HbO and HbR signals were passed through a Common Spatial Pattern (CSP) algorithm (Y. Wang et al., 2005) and concatenated with other EEG-derived CSP features to test the performance of a hybrid BCI system on a motor imagery classification task.
Table 1.
Surveyed studies that implemented GLM-based EEG-informed methods.
| Reference | Boxcar stimulus | HRF function | Study objective |
|---|---|---|---|
| Peng et al. (2014) | IEDs (spikes) | Canonical | Epileptic events |
| Pouliot et al. (2014) | |||
| Peng et al. (2016) | |||
| Y. Gao, Jia, et al. (2023) | Bandpower | Motor Imagery | |
| R. Li, Zhao, et al. (2020) | NVC investigation | ||
| Chiarelli et al. (2021) | Bandpower + PCA | FCA in AD | |
| Zama et al. (2019) | ERS/ERD | Response during movement | |
| M. T. Talukdar et al. (2015) | Envelope | Parametric | Median nerve stimulation |
| J. Lin et al. (2024) | Tensor decomposition | NVC investigation |
Another EEG-informed variant
On top of these GLM-based methods, M. J. Khan et al. (2018) and Arif et al. (2022) implemented a novel EEG-informed fNIRS analysis approach for the early detection of hemodynamics responses in hybrid BCI systems by analyzing the trajectory of the (HbO, HbR) vector in a phase diagram. There, two circles are drawn in the phase diagram: a bigger one corresponding to the maximum magnitude of during previous resting-state time and a smaller one corresponding to its maximum magnitude during a 1-second window in which the EEG signal was active. Positive activity is inferred by identifying certain patterns of the vector’s trajectory with respect to the inner and outer circles. With this vector-phase analysis, M. J. Khan et al. (2018) report not only an improved classification accuracy over their single-modality counterparts, but also an impressive reduction of command-generation time to 1.5 seconds on average. Another EEG-informed fusion method based on regularized CSP theory, called R-CSP-E, was proposed in Y. Wang et al. (2022), in which the weighted covariance matrix was transferred from EEG to fNIRS. Fusion occurs when computing the spatial filter for fNIRS in the CSP algorithm, in which the standard fNIRS covariance matrix is replaced by a regularized weighted average between the covariance matrices of each modality. The filter features were then used in a K-nearest neighbors (KNN) classifier.
3.4. fNIRS-informed EEG analysis
Using fNIRS features during EEG data analysis is another asymmetrical fusion technique. These methods leverage fNIRS spatial resolution in order to guide EEG processing, and are typically applied in two contexts: channel selection and source imaging analysis (Fig. 6). The former is commonly found in single-trial analysis, where the experimental paradigm involves different conditions or tasks, and its main goal is to reduce the dimensionality of the problem without compromising performance. This algorithm selects the subset of fNIRS channels that yields the highest contrast between conditions, according to some statistical analysis, such as t tests (R. Li et al., 2017), and then, the closest EEG channels to these fNIRS channels are selected.
Fig. 6.
fNIRS/DOT-informed fusion approaches: channel selection and source reconstruction.
3.4.1. fNIRS-informed EEG source imaging analysis
EEG image reconstruction
EEG source imaging corresponds to the problem of, given some recorded electrical activity on the scalp, reconstructing the cortical sources that generated that signal. This inverse problem is inherently ill-posed due to the fact that several source configurations can lead to the same scalp data. Nevertheless, source reconstruction algorithms can make certain assumptions on the underlying sources in order to mitigate this problem and hence improve spatial resolution. The linear forward model is one of the most popular proposals (Haufe et al., 2014; Parra et al., 2005):
| (2) |
Here, encodes the EEG recordings for the channels, and is a noise term. The underlying (unknown) sources are encoded in , corresponding to dipole cortical currents. is the so-called lead field matrix, which is known, or estimated by a given head model, and parameterizes how strong each dipole signal contributes to each channel. One way of inverting this model (see also Mäkelä et al., 2018; Mosher and Leahy, 1999, for approaches based on the MUSIC algorithms family) is by assuming specific distributions for the noise and currents, such as and , where and are the corresponding covariance matrices and denotes the Gaussian distribution. Under these assumptions, one can obtain the following estimate for the currents:
| (3) |
where is a regularization parameter, representing the trade-off between model accuracy and complexity. The solution (3) can be derived from a Bayesian perspective, where is the maximum a posteriori (MAP) estimate (Sato et al., 2004), corresponding to the current that maximizes the posterior distribution , with the likelihood , the prior and independent of . Prior knowledge of the sources is encoded in , which coincides with the identity matrix when no information is provided. It is worth emphasizing that the same solution (3) can also be reached from a minimum norm approach (Hauk, 2004), where and are not associated with specific distributions. Regardless of the exact approach, such a “blind” model is insufficient to resolve the ill-posed inverse problem fully.
Using fNIRS data as prior
A promising direction to overcome these limitations is to use data from other modalities with better spatial resolution as priors, such as fMRI or fNIRS (Fig. 6). In particular, two proposals created in the context of MRI–MEG/EEG fusion have been adapted for fNIRS–EEG fusion: Variational Bayesian Multimodal EncephaloGraphy (VBMEG) (Sato et al., 2004) and Dynamic Brain Transition Network (DBTN) (Nguyen et al., 2016). In VBMEG, anatomical information is used to determine the locations and orientations of the dipole currents, while functional information, from fMRI or fNIRS, is used to estimate the current amplitudes. More precisely, VBMEG assumes a diagonal covariance matrix for the currents , with the current variances following parametric Gamma distributions. Among the distributions’ free parameters, the mean variance values, which control the intensity of the corresponding currents, are estimated from fMRI or fNIRS features. In contrast, DBTN models as the sum of covariance components , where are -dimensional vector maps of the voxels that show statistically significant activity between experimental conditions (e.g., task vs. baseline). These maps are calculated via a GLM analysis on fMRI or fNIRS data.
3.4.2. Survey results
Channel selection
We encountered 8 articles (6.4%) implementing fNIRS-informed EEG analysis methods, 3 (2.4%) for channel selection, and 5 (4%) for source imaging analysis (Table 2). With the purpose of increasing motor task classification accuracy of hybrid BCI systems with as few channels as possible, R. Li et al. (2017) selected one fNIRS channel per hemisphere based on GLM analysis and t-tests, and Moslehi and Davies (2021) applied the ReliefF algorithm (Urbanowicz et al., 2018) (a filter-based method using KNN to weight channel relevance) to pick either the top five or three fNIRS channels from each frontal and motor regions. In both studies, the adjacent EEG channels to the selected fNIRS ones were chosen.
Table 2.
Surveyed studies that implemented fNIRS/DOT-informed methods.
| Approach | Reference | METHOD | Study objective |
|---|---|---|---|
| Channel Selection (fNIRS) | Al-Shargie et al. (2016) | t-test | Mental stress assessment |
| R. Li et al. (2017) | BCI classification | ||
| Moslehi and Davies (2021) | ReliefF algorithm | ||
| Source Reconstruction (fNIRS) | Aihara et al. (2012) | VBMEG | Reconstruction methods comparison |
| Morioka et al. (2014) | BCI classification | ||
| R. Li et al. (2019) | DBTN | FCA in AD | |
| R. Li, Li, et al. (2020) | FCA in poststroke patients | ||
| X. Li et al. (2022) | |||
| Source Reconstruction (DOT) | Cao et al. (2021) | ReML | Method presentation |
| Cao et al. (2023) | Method validation (visual task) | ||
| Qin et al. (2024) | Method validation (auditory task) |
fNIRS-informed source imaging
For source imaging analysis, Aihara et al. (2012) were the first to study on incorporating the idea of fNIRS-informed EEG source reconstruction, by using the VBMEG algorithm introduced above. There, the authors compared the performance of VBMEG for estimating cortical currents with fMRI prior, fNIRS prior, and no prior, on subjects performing a finger-tapping task. Using HbO peak values for modeling current mean variances, they concluded that the novel Bayesian technique allows to obtain relatively high spatiotemporal, physiologically plausible, brain activity, which cannot be obtained using EEG data alone. The same VBMEG method was applied later in Morioka et al. (2014) for BCI applications, where subjects’ left or right spatially attended direction was classified. This time, the current mean variances were estimated from HbO t-statistics between attention and control conditions. The DBTN method was utilized in a series of articles by R. Li et al. (2019), all of them following a similar approach in which the reconstructed cortical currents were first grouped into Regions of Interest (ROI) and then fed into a weighted Phase Lag Index (wPLI) algorithm (Vinck et al., 2011) to characterize brain connectivity networks in AD (R. Li et al., 2019), and poststroke patients (R. Li, Li, et al., 2020; X. Li et al., 2022).
DOT-informed source imaging
In the context of DOT–EEG fusion, Cao and collaborators proposed an algorithm that utilizes DOT reconstruction as the spatial prior of EEG reconstruction using the restricted maximum likelihood (ReML) framework (F. Abdelnour et al., 2010). There, the same EEG linear model (2) is used, together with a small extension of the current estimates (3), in which additional covariance-dependent terms are added. As for fNIRS-informed EEG source imaging, DOT priors enter through , which in this case is assumed to be diagonal with the non-zero entries related to the reconstructed HbO activities from DOT data. This relationship guarantees that a brain voxel should only have high electrical activity (reconstructed from EEG) when the hemodynamic activity (reconstructed from DOT) is high. Cao et al. (2021) introduced the framework and demonstrated its high spatiotemporal resolution for source recovery on simulated data. More precisely, they showed that neuronal sources that are spatially (2.3–3.3 cm) and temporally (50 ms) close can be clearly distinguished by the fusion method, but not by either modality in isolation. Moreover, they showed that such spatial resolution can be achieved even in DOT systems with a sparser, non-high-density optode configuration. The authors also compared the performance with the fNIRS-informed method of Aihara et al. (2012), concluding that when fNIRS projection (instead of DOT reconstruction) is used as the spatial prior, the algorithm performance drops. Later, Cao et al. (2023) validated the high spatiotemporal resolution of the ReML-based algorithm on a real-world dataset with subjects undergoing a visual stimulus task, in which the stimulus speed was faster than the temporal resolution of DOT. Finally, Qin et al. (2024) utilized the same framework for image reconstruction on simulated auditory cortex data, but this time including a novel data-informed method to select optimal DOT channel lengths for a more accurate source depth reconstruction.
3.5. Model-based EEG–fNIRS symmetric methods
The methods reviewed so far leverage the strengths of one modality to inform the other. While these methods have been successfully applied to capture inter-trial variability (Section 3.3.1) or achieve higher spatiotemporal resolution than each individual modality (Section 3.4.1), the asymmetrical treatment makes one of the modalities lack the benefits of the fusion. Symmetric fNIRS–EEG analysis methods represent an integrative approach where both modalities contribute equally to the fusion process. In particular, model-based symmetric fusion (Fig. 7) relies on knowledge to explicitly model relationships between neural activity and each of the recorded signals, leveraging physiological principles to guide the fusion process. These methods typically include a single-modality preprocessing step before fusion to bring the datasets to a common feature space, such as bandpower time courses, or network connectivity matrices.
Fig. 7.
A priori and learnable model-based symmetric fusion strategies.
Some of the less sophisticated approaches consist of NVC-like feature generation employing handcrafted functions to merge single-modality data, such as correlation, wavelet coherence, or convolution. While these methods utilize knowledge from the NVC theory to build non-linear combinations of modalities, they lack tunable parameters that could improve performance upon training.
3.5.1. Survey results
fNIRS
From the fNIRS–EEG survey, we found 29 articles (23.2%) adopting a model-based symmetric framework (Table 3). Among them, only 4 articles (3.2%) used model-based symmetric methods with learnable parameters. In Sood et al. (2016), the authors developed and tested a Kalman-Filter-based online parameter estimation of an autoregressive (ARX) model to track the transient coupling relation between EEG bandpower changes and fNIRS HbO changes signals during transcranial Direct Current Stimulation (tDCS). There, fusion takes place upon using the EEG and fNIRS features as inputs and outputs of the ARX model, respectively. A novel permutation transfer entropy (PTE) model was proposed in Y. Gao, Liu, et al. (2023) borrowing some ideas from the GLM-based EEG-informed algorithm described in Section 3.3. There, EEG features, corresponding to the peak and latency of the first principal components of EEG frequency bands, are used to modify a typical boxcar function. The latter is then convolved with a canonical HRF to build a design matrix. As reviewed in Section 3.3.1, this matrix is typically used for modeling the hemodynamic response in an fNIRS GLM. However, here PTE is used instead, to consider the directed coupling between the neuronal activities and hemodynamic response. Finally, the generalized relative entropy of PTE time points is computed and used as input of a SVM classifier to quantify mental workload. As opposed to the asymmetrical GLM-based EEG-informed approach, here modalities are treated in a symmetrical fashion.
Table 3.
Surveyed studies that utilized symmetric model-based fusion approaches.
| Approach | Reference | METHOD | Study objective |
|---|---|---|---|
| A priori (fNIRS) | L.-C. Chen et al. (2015) | Correlation | NVC investigation |
| Keles et al. (2016) | |||
| Balconi and Vanutelli (2016) | |||
| Zich et al. (2017) | |||
| Suzuki et al. (2018) | |||
| Y. Chen et al. (2020) | |||
| Saricaoglu et al. (2025) | |||
| J. Lin et al. (2023) | |||
| Aghajani et al. (2017) | BCI classification | ||
| Omurtag et al. (2017) | |||
| C.-T. Lin et al. (2020) | |||
| Chu et al. (2022) | |||
| Borgheai et al. (2019) | NVC in ALS patients | ||
| Kaga et al. (2020) | NVC in ADHD patients | ||
| C. Gao et al. (2024) | NVC of myopic subjects | ||
| Y. Zhang and Zhu (2019) | Resting-state FCA | ||
| Nair et al. (2021) | Anesthesia monitoring | ||
| Si et al. (2024) | Emotion assessment | ||
| Qu et al. (2025) | FCA in stroke patients | ||
| Govindan et al. (2016) | Spectral coherence | NVC investigation | |
| Perpetuini et al. (2020) | Conditional entropy | Working memory in AD | |
| Nourhashemi et al. (2020) | Transfer entropy | NVC in neonates | |
| Othman et al. (2021) | Wavelet coherence | Brain injury | |
| Charly and Onabid (2024) | Convolution | Speech quality | |
| Wan et al. (2024) | Multimodal covariance network | FCA in autistic patients | |
| A priori (DOT) | Giacometti and Diamond (2014) | Correlation | NVC investigation |
| Chalia et al. (2016) | Burst suppression | ||
| Learnable (fNIRS) | Sood et al. (2016) | Kalman filter + ARX model | tDCS |
| Gentile et al. (2020) | Linear regression | Laser-evoked potential | |
| Y. Gao, Liu, et al. (2023) | PTE | Mental workload | |
| T. Gao et al. (2025) | Linear regression | Etomidate use disorder |
DOT
From the DOT-EEG survey, two works used plain correlation as the fusion strategy. In Giacometti and Diamond (2014), the authors studied the relationship between colocalized HD EEG and fNIRS recordings on a single subject by calculating correlation between EEG and fNIRS cortical sensitivity matrices, and tabulated the maximum correlated pairs to be used for future studies on NVC. Correlations between the inverse models were also computed for the whole cortical surface and for ROI. A qualitative analysis between EEG and DOT data was carried out in Chalia et al. (2016) to study the hemodynamic response to burst suppression on infants with hypoxic ischemic encephalopathy. There, the authors managed to identify a temporal correlation between EEG bursts and HbO and HbR changes in channel and image space.
3.6. Data-driven EEG–fNIRS symmetric methods
Fusion methods in this section combine the strengths of symmetric and data-driven approaches to create a flexible framework in which relationships between fNIRS and EEG signals are inferred directly from the data, and used in a bidirectional fashion so each modality receives and provides guidance during processing. We consider these methods to be particularly promising for continuous, naturalistic brain imaging, as their inherent flexibility enables adaptive handling of multimodal data in dynamic and unconstrained settings. Despite these advantages, previous reviews have overlooked these methods. Therefore, this section represents the core contribution of our manuscript, aiming to address this gap and thoroughly explore the potential and characteristics of these underrepresented, yet highly promising, fusion techniques.
3.6.1. Data selection
These algorithms can be further divided into subcategories such as feature/data-selection and feature-extraction. The former category encompasses methods in which a subset of the most relevant EEG and fNIRS features (or raw channel signals) are selected according to some criterion, such as correlation, statistical tests, or mutual information (Table 4). These methods tend to be computationally cheap, and can be used at data/sensor-level (channel selection) or feature-level (feature selection), typically in combination with other feature-extraction algorithms. However, as individual fusion methods, they may fail to uncover complex structures on the data since they do not modify the original signals.
Table 4.
Surveyed studies that implemented data-driven symmetric methods.
| Approach | Reference | METHOD | Study objective |
|---|---|---|---|
| Data Selection | Hasan et al. (2020) | Correlation (channel selection) | BCI classification |
| M. U. Khan and Hasan (2020) | |||
| Ali et al. (2024) | |||
| Zennifa et al. (2018) | Correlation (feature selection) | Engagement recognition | |
| Borgheai et al. (2024) | BCI performance prediction | ||
| Yin et al. (2015) | Mutual information | BCI classification | |
| Deligani et al. (2021) | |||
| Hosni et al. (2022) | LASSO* | ||
| Yi et al. (2023) | Depression assessment | ||
| M. Meng et al. (2021) | Mutual information + LASSO | BCI classification | |
| Ergün and Aydemir (2023) | Cross validation | ||
| Ge et al. (2019) | ReliefF algorithm | NVC in action observation | |
| Cicalese et al. (2020) | Correlation (wrapper-based) | AD stages classification | |
| Ali et al. (2023) | E-WOA | Method presentation | |
| Source Decomposition | Al-Shargie et al. (2017b) | CCA | Mental stress assessment |
| Ghasimi and Shamekhi (2024) | Cognitive workload assessment | ||
| Rezaee et al. (2021) | tCCA | tDCS response in stroke patients | |
| Walia et al. (2022) | Error-related brain states | ||
| Dashtestani et al. (2022) | ssmCCA | Action–observation network | |
| Su et al. (2023) | NVC during motion | ||
| Dähne et al. (2013) | mSPoC | Method presentation | |
| Ieong et al. (2019) | NVC in opiate addiction | ||
| Al-Shargie et al. (2016) | jICA | Mental stress assessment | |
| Hossain et al. (2022) | WPD-CCA | Motion artifact correction | |
| Xu et al. (2023) | PCA | BCI classification | |
| Ran et al. (2025) | TdCCA | FCA in Moyamoya disease | |
| Nia et al. (2025) | CCA + kCCA | Emotion recognition | |
| McLinden et al. (2024) | PAC | NVC investigation | |
| Zhu et al. (2024) | CMVGP-CCM | ||
| Hou et al. (2024) | Hybrid brain network | Emotion recognition | |
| Blanco et al. (2024) | Multiplex network | FCA |
Least Absolute Shrinkage and Selection Operator algorithm (Hastie et al., 2015).
3.6.2. Source-decomposition methods
Data-driven feature extraction combines EEG and fNIRS data at sensor level, and are promising alternatives to their model-based counterparts when theoretical models are insufficient to capture the complex dynamics of the underlying physiological process. Multimodal source decomposition (Zhuang et al., 2020) (Fig. 8) is a popular approach in which the data are assumed to follow a linear forward model:
Fig. 8.
Multimodal source-decomposition methods.
| (4) |
Here, datasets are encoded in and , and their columns and denote multivariate EEG and fNIRS observations. and are the number features and samples/observations, respectively.6 Measurements are decomposed into components or sources , whose strengths are modulated by the corresponding spatial activation patterns, encoded in the columns of . Despite a similar linear form between (4) and (2), it is worth noting the blind nature of the current source separation model, in which not only the sources but also the mixing matrices and are unknown and must be estimated from the data. The noise terms are supposed to capture the activity which is not explained by the sources. We assume both signals contain a subset of components which are related by some unknown relationship. The main objective of source-decomposition methods is to uncover such a relation and find the reconstructed sources via a linear backward model
| (5) |
The filters can then be used to estimate the spatial patterns of the corresponding forward model via (Haufe et al., 2014)
| (6) |
where , are the covariance matrices for each modality. The filters are obtained as the solution of a given inverse optimization problem, which ultimately defines the specific source-decomposition method. In what follows, we describe the most relevant of these methods extracted from the survey. When required, we assume , which can be achieved upon resampling.
3.6.3. CCA and its variants
Standard CCA
Canonical Correlation Analysis (CCA) (Hardoon et al., 2004; Hotelling, 1936) is a data-driven, unsupervised method that seeks to solve the inverse problem (5) by finding linear combinations of the data that are maximally correlated with each other. In its simplest formulation, CCA finds one-dimensional filters and such that the reconstructed sources (a.k.a canonical variates) maximize the correlation coefficient . The optimization problem can be brought to a constrained quadratic form by imposing unit variance on the reconstructed sources
| (7) |
where we denoted the cross-modality covariance matrix . By using the Lagrange-multiplier method (Hardoon et al., 2004), one can reformulate (7) into a generalized eigenvalue problem where eigenvalues coincide with estimated source covariances, , and so the optimal solution corresponds to the one with the largest eigenvalue. It can be shown (L. Sun et al., 2009) that if are whitened, that is, and , then CCA reduces to Partial Least Squares (PLS) (Wold et al., 1984; Worsley et al., 1997). The one-unit algorithm can be extended to find filters (Kettenring, 1971), , by solving the same generalized eigenvalue problem but for the first eigenvectors with the highest eigenvalues. Picking an orthonormal basis, which is always possible via a deflation algorithm (Mackey, 2008), for instance, the resulting filters maximizes such that .
Regularized CCA
In modern neuroimaging datasets, we often encounter situations in which the number of variables is much larger than the number of observations, making CCA to suffer from issues such as high-dimensionality, multicollinearity, and overfitting. This is typically the case in group-level analysis, where variables such as voxels or time points clearly exceed the number of subjects (observations). Regularized CCA adds penalty terms to the objective function (7), controlling model’s complexity, reducing overfitting, and improving generalization. These regularized methods admit a general constrained optimization problem description, in terms of generic convex penalty functions and (Witten et al., 2009; Witten & Tibshirani, 2009)
| (8) |
where , are positive scalars. Many implementations (J. Chen et al., 2013; Dashtestani et al., 2022; Witten et al., 2009) replace the covariance matrices, and , with identity matrices, claiming that there is no loss in performance.
ElasticNet CCA
Two of the most common choices for the penalty terms and are the L2-norm, , which regulates the overall magnitude of the coefficients and leads to Ridge CCA (Bie & De Moor, 2003), and the L1-norm, , which searches for a sparser solution, setting some coefficients to zero, leading to sparse CCA (sCCA) (Parkhomenko et al., 2009). Both penalty terms can be combined to obtain Elastic Net regularization (Hastie et al., 2015).
ssCCA
Elastic Net regularization can be modified to include contextual information about the local structure of the dataset, leading to structured sparse CCA (ssCCA). This is particularly relevant in neuroimaging, where features (e.g., channels or brain regions) follow specific spatial distributions. This can be achieved via Graph-constrained Elastic Net (GraphNet), where the L2 norm penalty function is replaced by with corresponding to the Laplacian matrix of an undirected weighted graph (Chung, 1997). The latter is used as a representation of the dataset, in which features are the nodes, and (weighted) edges indicate which and how features are connected. These edges are encoded in the adjacency matrix with corresponding to the number of features/nodes. With , one can build the degree matrix , which is diagonal with the non-zero entries , and the Laplacian matrix, , which is the canonical matrix representation of a graph. GraphNet encourages filter components to be more similar, if their corresponding features are more connected (He & Niyogi, 2003). The latter is regulated by the weights in , which need to be determined a priori. Popular choices are the binary KNN adjacency matrix, where each feature is connected to its k closest neighbors with a value of 1, or the distance-based weighted approach, where closer features get assigned higher weights using functions such as Gaussian kernels. In the context of wearable neuroimaging, the montage layout/grid can be used to define such connections. The constrained optimization problem defining ssCCA can be also brought into the form of (8) (J. Chen et al., 2013; Witten et al., 2009), with and , in which and are the Laplacian matrices of each dataset.
kCCA
Linear models (4), such as standard CCA, are reasonable first-order approximations for many neuroimaging applications, but they may fall short in certain scenarios where the modalities’ relationships are non-linear in nature. Kernel CCA (kCCA) (Friston et al., 2000; Fukumizu & Gretton, 2007) addresses this limitation by implicitly mapping the original datasets into a high-dimensional feature space, , embedded in kernel functions which are independent of the new number of features. Nonlinearities are encoded in the choice of such kernels, such as polynomials or radial basis (Gaussian) functions, and standard CCA can be recovered by choosing to be the identity function (i.e., ). The kCCA algorithm can be obtained as the non-linear extension of (7)
| (9) |
Interestingly, a similar close-form solution via a generalized eigenvalue problem exists in this case, and the same regularization techniques mentioned above can be applied. By leveraging kernels, kCCA can effectively uncover complex, nonlinear associations between modalities. The need for a pre-defined kernel function to model nonlinearities (model-based) is alleviated in the DL extension of CCA, deep CCA (Andrew et al., 2013), which learns the feature mapping directly from data (data-driven), offering a more flexible and robust approach. However, interpretability in these deep methods is a matter of ongoing research.
tCCA
All the methods described so far assume modalities correlate instantaneously, which is certainly not the case in fNIRS–EEG fusion, for instance, where the common underlying sources are time-shifted due to non-instantaneous and non-constant couplings between modalities. Temporally embedded CCA (tCCA) (Bießmann et al., 2010) captures such temporal offsets. While effective, the idea is rather simple: assuming is the “delayed” modality, is time embedded by concatenating time-shifted copies along the feature dimension via , where is a copy of the original but time shifted by . These time lags must be preselected, and the time dimension size can be preserved upon padding. Standard CCA (or any of the variants explored above) is then applied to and , yielding temporal filters that capture delayed correlations between modalities. In the context of fNIRS, this method has been shown to improve performance in HRF recovery in combination with a multimodal extension of the GLM (von Lühmann, Li, Müller, et al., 2020).
mCCA
Multiset CCA (mCCA) extends CCA to more than two datasets (Correa et al., 2010; Kettenring, 1971). Given datasets, , mCCA aims at finding filters such that the overall correlation between datasets is maximized
| (10) |
Here, is the cross-covariance matrix between datasets and , and the covariance of dataset . is an appropriately chosen cost function of all pair-wise correlation matrices between canonical variates, ( ). In Kettenring (1971), five objective functions were proposed, including the sum of correlations (SUMCOR) and the sum of squares correlations (SSQCOR). The one-unit mCCA algorithm can be extended to find multiple canonical variates by using the iterative deflation method. In the first step, we simply solve (10) to obtain the first set of filters with . In the next iterations, we turn (10) into a constrained optimization problem by imposing orthogonality of the form for all . The maximum number of canonical variates is restricted by .
ssmCCA
Similarly to standard CCA, the GraphNet ssCCA variant presented can be also extended to multiple datasets, leading to the so-called structured sparse multiset CCA (ssmCCA) (Dashtestani et al., 2022), created in the context of EEG–fNIRS fusion for group-level network connectivity analysis. The algorithm is defined as the natural combination of ssCCA and (10). In particular, ssmCCA uses SUMCOR as the multiset objective function, leading to the one-unit algorithm
| (11) |
Here, the cross-covariance matrices were replaced by identity matrices (Witten et al., 2009), are a set of hyperparameters, and the Laplacian matrix of dataset . The extension to multiple filters can be obtained following the same logic as for mCCA (Dashtestani et al., 2022).
3.6.4. Multimodal source power co-modulation (mSPoC)
EEG bandpower is proven to modulate with cognitive tasks, as well as with hemodynamic responses via the NVC, and so it is important to have data-driven fusion methods that capture such a comodulation correctly (Nunez & Srinivasan, 2006; Parra et al., 2005). Conventional methods, such as CCA, however, present some limitations to integrate bandpower signals (Dähne et al., 2013). More precisely, computing bandpower at channel level and then applying a backward linear method (5) are not in line with the assumed linear generative model of EEG (4), which harms the neurophysiological interpretation of the reconstructed sources. mSPoC (Dähne, 2015; Dähne et al., 2013) arises as a method that avoids these pitfalls by inverting the generative model prior to computing bandpower.
The method assumes that both datasets are mean free and temporally aligned with different sample dimensions . EEG recordings, , must be band-pass filtered and divided into non-overlapping time windows, indexed by , and the fNIRS time series should be subsampled so that there is exactly one data point per EEG window. With this resampling, we can assume EEG is approximately stationary within time windows, allowing us to compute summary statistics such as average bandpower. Assuming the backward model of (5), the algorithm aims to uncover components , such that the temporally embedded bandpower dynamics of co-modulates with . To this end, the source bandpower within each time window is approximated by
| (12) |
with denoting the within-window covariance matrix, and time delays between modalities are captured via a finite impulse response filter
| (13) |
parameterized by an -dimensional time filter , and a set of pre-chosen time lags . mSPoC is then defined by the following constrained optimization problem
| (14) |
Via the method of Lagrange multipliers, (14) can be turned into a set of coupled equations that must be solved iteratively, until some convergence criterion is fulfilled. The extension to set of filters can be obtained via a deflation scheme (Dähne, 2015).
3.6.5. Joint ICA
Joint ICA (jICA) (V. Calhoun, Adali, Pearlson, & Kiehl, 2006) is a very straightforward method that was created in the context of unimodal multi-task fMRI analysis (V. D. Calhoun et al., 2006) and rapidly adapted to the fusion of multiple modalities such as fMRI-sMRI (V. Calhoun, Adali, Giuliani, et al., 2006), fMRI-EEG (V. Calhoun, Adali, Pearlson, & Kiehl, 2006), and fMRI-sMRI-EEG (V. Calhoun, Adali, & Liu, 2006). The method is typically used for group-level analysis, and so we now denote as the number of subjects, EEG and fNIRS features, respectively. Single-modality datasets, and , are concatenated along the subject axis to generate a multimodal feature matrix . jICA then assumes the following generative linear model
| (15) |
where is the number of components. Each of these components is characterized by an -dimensional profile and an -dimensional vector that encodes the strength at which the profile is present on each subject. A standard ICA algorithm is then run on (15), assuming statistical independence between , to obtain the reconstructed profiles , which are decomposed into modality-specific sources. Via this approach, jICA finds latent factors that commodulate between modalities and across subjects.
jICA assumes different modulation vectors for each subject, but the same for both modalities, giving the algorithm small flexibility. This assumption is relaxed in parallel ICA (paraICA) (J. Liu & Calhoun, 2007), another ICA extension that makes emphasis on subject-specific multimodal components. However, to the best of our knowledge, paraICA has not been used for fNIRS–EEG fusion yet.
3.6.6. Survey results
We found no data-driven symmetric methods for DOT–EEG fusion. From the fNIRS–EEG survey, we extracted 14 data selection algorithms (11.2%), and 17 feature extraction ones (13.6%), most of them source decomposition (Table 4).
CCA in stress and workload evaluation
The advantages of implementing standard CCA for EEG–fNIRS fusion during mental stress and cognitive workload assessment were demonstrated in Al-Shargie et al. (2017b) and Ghasimi and Shamekhi (2024), respectively. Al-Shargie and colleagues used CCA analysis to reveal that the pair of canonical variates with the highest correlation (0.95) between EEG alpha bandpower suppression and decreased HbO was consistently observed in the right ventrolateral prefrontal cortex (VLPFC), indicating a highly localized area for stress-induced neural changes. This fusion approach significantly enhanced the detection of mental stress, achieving the best classification accuracy (control vs. stress), compared with using either modality alone. Moreover, CCA’s performance was also superior to the one achieved previously with jICA in Al-Shargie et al. (2016), using the same montage and mental arithmetic task. In Ghasimi and Shamekhi (2024), the authors explored the combination of six distinct fusion strategies with several classifiers for cognitive workload assessment. Among them, the combination of CCA with a linear SVM classifier achieved the highest performance for differentiating between various n-back task levels.
CCA and kCCA for emotion recognition
Nia et al. (2025) conducted a comparative study between concatenation, decision-level, CCA, and kCCA fusion strategies to examine the effectiveness of each framework as a hybrid fNIRS–EEG emotion recognition system. The study employs a personalized dataset of an auditory emotion-elicitation paradigm using a sparse multimodal montage over the prefrontal cortex. Power spectral density and differential entropy were used as features for both modalities. The study indicates that concatenation and decision-level fusion were both in par and outperformed the CCA-based strategies across different binary emotion classification tasks, and, on some occasions, the CCA methods performed even worse than single-modality classifiers. Such a poor performance can be attributed to a lack of auxiliary signals and physiological and motion-related noise removal for both modalities. These modality-specific confounders can not only screen shared components, harming CCA performance, but also inflate classification performance in late-fusion methods, such as concatenation and decision level, by allowing them to leverage extracerebral information. Additionally, the absence of robust regularization mechanisms in the CCA methods exacerbates overfitting issues, especially problematic given the high-dimensional input space obtained upon combining the spectral and nonlinear features.
CCA for motion artifact correction
Wavelet Packet Decomposition (WPD) (Mallat, 1999) was combined with CCA (WPD-CCA) in Hossain et al. (2022) for motion artifact correction, decomposing single-channel recordings into multi-channel sub-band signals. These individually generated fNIRS and EEG signals are then used as CCA inputs. The efficacy of the framework is tested on a benchmark dataset using the difference in SNR and percentage reduction in motion artifacts as metrics, concluding that WPD-CCA performs better than single-modality WPD alone.
tCCA in clinical contexts
In Rezaee et al. (2021), tCCA was used to detect continuous transcranial Direct Current Stimulation (ctDCS)-induced cerebral activation changes in chronic stroke survivors. There, resting-state fNIRS–EEG data from left/right prefrontal and sensorimotor cortices were collected before and after ctDCS. tCCA was applied to capture the coupling between the log10-transformed EEG bandpower (1–45 Hz) and total hemoglobin signals (HbO + HbR). Notably, significant canonical correlations were observed only in the vasomotion band (0.07–0.13 Hz), indicating that this frequency window robustly reflects the intermodal relationship, while correlations in the neurovascular coupling band (0.01–0.07 Hz) were not significant. An interesting hybrid data-driven/model-based approach was implemented in Walia et al. (2022) for exploring error-related brain states during complex surgical motor tasks. There, the authors utilized tCCA between HbO data from LS channels and EEG bandpower to extract 15 canonical EEG components. These components were used as hemodynamic regressors (replacing the typical HRF and boxcar function) of a GLM model, which also incorporated SS channels as confounding terms. This fusion technique demonstrated that expert surgeons exhibited a focused pattern of hemodynamic suppression in key cognitive and error-monitoring regions, while novices showed a more widespread activation. Another variation of CCA called temporally driven CCA (TdCCA) was proposed in Ran et al. (2025), where the imaginary part of coherence of the EEG signal and the fNIRS correlation coefficients in the resting and working memory state were used as inputs of standard CCA. This novel method achieved the best performance in brain connectivity classification between patients with Moyamoya disease (MMD) and control groups, compared with other methods, including sCCA and kCCA. This superior performance is attributed to TdCCA’s ability to effectively handle the temporal delays inherent in EEG–fNIRS data via the custom-tailored brain connectivity features.
ssmCCA for action–observation network analysis
Dashtestani et al. (2022) developed ssmCCA (11) to study multimodal, multi-participant fNIRS–EEG data fusion to characterize the human action–observation network during a live-action–observation and execution paradigm. The results from right-handed participants revealed that the left inferior parietal region was significantly activated during both paradigms. ssmCCA offered more refined spatial insights than unimodal analyses alone. Moreover, the authors compared the performance of ssmCCA with sparse mCCA and mCCA, finding that ssmCCA showed the largest correlation magnitudes in both conditions, as a consequence of being the only method among the three that handles multi-collinearity, high dimensionality, and small-sample size at the same time. The same method was later applied for NVC modeling during motor execution, observation, and imagery tasks in Su et al. (2023), validating and extending previous results.
mSPoC
The mSPoC algorithm was developed in Dähne et al. (2013) with the purpose of incorporating a source-decomposition method with a biologically accurate non-linear relation between EEG time courses and hemodynamic responses (as measured by fNIRS or fMRI) via EEG bandpower. The validity of mSPoC was demonstrated for simulated data and in a real-world motor execution paradigm, showing a higher performance than standard CCA in both datasets. mSPoC was used later in Ieong et al. (2019) to investigate NVC on opiate adiction, using functional connectivity-related features as model inputs. The framework allowed the finding of desynchronized alpha rhythms and reduced functional connectivity in abstinent heroin-dependent patients, indicating cerebrovascular injury in the PFC and suggesting a chronic effect from opiate exposure.
Other symmetric data-driven methods
During the survey, we found other novel bi-directional data-driven methods that followed different strategies from the typical source-decomposition framework. For instance, in the realm of NVC investigation, McLinden et al. (2024) and Zhu et al. (2024) developed their own fusion approaches. In the former study, instantaneous EEG power amplitude and fNIRS phase were fed into an integrated, non-linear method called electrovascular phase–amplitude coupling (PAC) with the aim of quantifying NVC during an auditory task. The Collaborative Multi-output Variational Gaussian Process Convergent Cross-Mapping (CMVGP-CCM) method was presented in Zhu et al. (2024) and used for causal analysis of EEG–fNIRS signals during an n-back paradigm, demonstrating differences in NVC under varying cognitive loads.
4. Methods Comparison on Simulated Data
4.1. Methods comparison
From our survey, we found that source-decomposition techniques were underrepresented in the literature. We, therefore, sought to evaluate them directly, as we hypothesize they are particularly well suited to continuous naturalistic brain imaging, due to their promising capabilities of disentangling the complex interplay between neural and systemic physiological activity in an unsupervised fashion, that is, without the need for labeled stimuli or events. To this end, a semi-synthetic dataset was used to compare the performance of previously reviewed methods (Section 3.6) on source reconstruction and spatial pattern estimation (Fig. 10). The dedicated goals here were to explore the differences between methods and their trends on the same benchmark, highlighting the critical roles of regularization techniques and temporal embedding. However, it is important to note that an exhaustive comparative evaluation falls beyond the scope of this review and is, therefore, reserved for future studies.
Fig. 10.
Comparison of methods’ performance for different SNR values, measured by the correlation between true and estimated quantities on a test (unseen) dataset. Each bullet point represents the bias-corrected average of such correlations across 50 simulations, and horizontal bars correspond to the average error. The first row shows correlations between reconstructed and ground-truth sources for EEG (left) and fNIRS (right), while the second row shows correlations between estimated and true spatial patterns for EEG (left) and fNIRS (right).
4.2. Simulated data
To systematically evaluate the performance of classical machine-learning fusion methods for EEG–fNIRS integration, it is essential to benchmark them, ideally on data with a known ground truth. However, this presents a significant challenge: publicly available EEG–fNIRS datasets are limited in number (Buccino et al., 2016; Ortega et al., 2020; Shin, Von Lühmann, et al., 2018; Shin et al., 2017), and to the best of our knowledge, there are no existing datasets providing synthetic ground-truth information in both modalities while also maintaining a realistic (physiological) noise background. This lack of accessible, well-characterized data hinders the development and validation of source-decomposition methods. To address this gap, we generated a realistic multimodal HD-fNIRS–EEG dataset with synthetic ground truth that simulates a typical motor task, such as finger tapping (Fig. 9). The resulting dataset comprises biologically plausible EEG–fNIRS neural activity, ensuring shared spatiotemporal co-modulation between sources, localized in the primary motor cortex, via simultaneous EEG alpha-band suppression and oxygenated hemoglobin increase in fNIRS. The simulated neural activity sits on top of a real concurrent HD-fNIRS–EEG recording from a single subject at rest, providing correlation between underlying processes already at noise level. By incorporating a high-density fNIRS configuration, our dataset offers an enriched platform for developing and testing fusion methods, benefiting from improved spatial resolution and enabling applications such as 3D image reconstruction through DOT. It should be noted, however, that this is not intended as the most advanced or comprehensive simulation framework; rather, it serves as a benchmarking resource to illustrate and evaluate multimodal fusion approaches within the scope of this review, with limitations and future directions discussed in Section 5. We nevertheless still describe it as “realistic” to distinguish it from fully synthetic toy datasets lacking physiological grounding, which often yield overly optimistic performance estimates.
Fig. 9.
Schematics of the synthetic data pipeline and its integration into the source-decomposition methods comparison. The montage on the left shows the location of EEG channels (green), and fNIRS sources (red) and detectors (blue). The exaggerated bandpower decay in and HbO increase in result from choosing a very high SNR, just for visualization purposes.
4.3. Realistic HD-fNIRS–EEG data with synthetic ground truth
4.3.1. Paradigm, montage, and background data
The EEG–fNIRS data simulate concurrent channel-space recordings on a single subject performing a motor task, such as finger tapping. The fNIRS dataset comprises optical-density (OD) recordings at 2 wavelengths (760 and 850 nm) across 100 measurement channels, created by pairings of 14 light sources and 32 detectors arranged in a bilateral high-density hexagonal grid over the motor cortex, with source–detector separations of 19 and 33 mm (Y. Gao, Rogers, et al., 2023). EEG data represent electrical activity collected from the 32-channel actiCAP arrangement (Brain Products GmbH, Gilching, Germany). The simulated experiment consists of one single block composed of 12 trials, each of which begins with a 10-second stimulus period followed by a recovery period whose duration is randomly chosen between 8 and 16 seconds. To produce realistic background noise underlying the simulated experiment, we collected 5 minutes of real concurrent HD-fNIRS–EEG resting-state data from a single subject, using the aforementioned montage on a custom-designed NinjaCap (von Lühmann et al., 2024). fNIRS data were recorded by two NIRSport2 devices (NIRx GmbH, Berlin, Germany) at 12.6 Hz sampling rate. EEG data were collected using a 32-channel LiveAmp amplifier (Brain Products GmbH, Gilching, Germany) with active Ag/AgCl electrodes mounted in the NinjaCap following the actiCAP arrangement, using common average reference. The sites normally occupied by FT9, FT10, T7, and T8 were repurposed to record horizontal and vertical EOG channels. Impedances were brought below 10 k before data acquisition, and data were digitized at 1000 Hz. Given the importance of spatial accuracy around motor areas, two in-house co-localized fNIRS–EEG probes (Rogers et al., 2025) were placed on the C3 and C4 landmarks. As a preprocessing step on the background data, prior to synthetic data augmentation, we removed as much alpha motor activity from EEG data as possible so true sources did not interfere with our synthetically generated ones. To this end, we used the spatiospectral decomposition (SSD) algorithm (Nikulin et al., 2011) to identify alpha-band neural oscillations and removed the components whose highest activity lay in the left-motor region.
4.3.2. Source spatiotemporal profile
We assumed a single common source between modalities. The source spatial profiles were created using EEG and fNIRS linear forward models that shared the same head model of the anatomy of Colin27 (Holmes et al., 1998), segmentation masks (gray/white matter, air, CSF, bone, and skin surfaces calculated using the segmentation pipeline of Harmening & Miklody (2022) based on SPM12 (Wellcome Centre for Human Neuroimaging, UCL, 2014)), FEM volume mesh (using CGAL (The CGAL Project, 2024)), and source location. The latter was chosen to be the point in the brain surface closest to the C3 landmark on the scalp. The sensitivity matrix was calculated via photon simulation using NIRFASTer (Dehghani et al., 2009) within Cedalion (IBS-Lab & Cedalion-Developers, 2024), and the EEG leadfield was computed with the SimBio FEM solver (Vorwerk et al., 2018) as integrated in the FieldTrip toolbox (Oostenveld et al., 2011). These forward matrices were then used to create fNIRS and EEG mixing matrices by combining them with modality-specific spatial information. For EEG, we used a single dipole structure and for fNIRS we generated a Gaussian-distributed activity mask, both centered around the common source location as explained above.
Simulated source time courses followed the stimulus/recovery structure of the experiment described above, where activity periods resembles biologically plausible motor activity: increasing fNIRS HbO activity while decreasing power in the alpha band (ERD). The EEG source was created using band-limited (8–12 Hz) noise modulated by slow task-related temporal envelopes, while the fNIRS HbO source was modeled using canonical hemodynamic response functions time locked to the beginning of the ERD (EEG source), including the typical hemodynamic time delay relative to activity onset. Such a synthetic hemodynamic response (together with the Gaussian-distributed spatial pattern) was generated from Cedalion’s functionalities, corresponding to the image-space extension of the method developed in von Lühmann, Li, Gilmore, et al. (2020). In each trial, the HRF amplitude was chosen randomly and proportionally to the corresponding EEG bandpower decrease. In this way, EEG source power reduction correlates with an increase in the hemodynamic response. After projecting the sources to channel space via the corresponding mixing matrices, the synthetic activity was normalized, and added to the background data, together with a scaling parameter that allows regulating the SNR. The latter were chosen to span from -25 dB to 10 dB, covering and extending ranges typically observed for EEG and fNIRS, and consistent with a prior multimodal simulation framework (Dähne et al., 2013).
4.4. Evaluation of source-decomposition methods
4.4.1. Methods
We used our synthetic data to compare some of the source-decomposition methods encountered in the survey (Section 3.6): the mSPoC algorithm and temporally embedded versions of CCA, sCCA, Ridge CCA, ElasticNet CCA, and ssCCA. The regularization parameters were chosen through a coarse-grained guided search, by exploring the performance of a few runs, rather than via an exhaustive grid or random search. More precisely, we selected regularization strengths that lie midway between very low and very high penalties, yielding reasonable performance across all methods without favoring one approach over another. For sCCA we chose the L1-penalty parameters to be , for Ridge CCA, we chose as the L2-penalty parameters, and for ElasticNet CCA we used their combination.7 ssCCA used the same regularization parameters, while the EEG and fNIRS Laplacian matrices were built using a binary nearest neighbors approach in which the adjacency matrices contain unit values where features (channels) were closer than the (pre-defined) distances of 25 mm for fNIRS and 60 mm for EEG. Due to the time delay of the hemodynamic response, we found temporal embedding to be mandatory in all CCA-based methods to achieve satisfactory performance levels. For all of them, we chose the same set of time lags: ranging from 1 to 4 s in 1-s steps. To demonstrate the need for temporal embedding, we also tested standard CCA. mSPoC’s original formulation (Dähne, 2015) does not include any regularization mechanism (14), so we included an L2 penalty term to ( ) together with PCA-based dimensional reduction, keeping only of the explained variance, to obtain a competitive performance.
4.4.2. Simulation and preprocessing
We ran 50 simulations with 12 trials each. Each simulation produced different temporal profiles for the EEG and fNIRS true sources due to the randomness in stimulus amplitudes and recovery times, but the source spatial profiles and the background data were always the same. For each simulation, we combined synthetic and background data using 20 different SNR values, from -25 to 10 dB. For each run and SNR, standard preprocessing pipelines were applied using Cedalion for fNIRS and the MNE toolbox (Gramfort et al., 2014; Larson et al., 2024) for EEG. fNIRS preprocessing involved channel pruning using the scalp coupling index and peak spectral power quality metrics (Pollonini et al., 2016), bandpass filtering from 0.01 to 0.6 Hz, linear detrending, and 2 Hz dowsampling. For EEG, we used linear interpolation to replace bad channels, a linear EOG regression algorithm for ocular artifact removal, and bandpass filtering in the alpha band. Then, EEG data were duplicated, and one of the copies was used to estimate its bandpower time course by splitting the data into equal-length segments and calculating the per-segment variance, resulting in an effective 2 Hz downsampling. The three datasets, namely fNIRS, EEG, and EEG bandpower time series, were segmented into individual trials, composed of the stimulus period (10 s) plus pre- and post-stimulus resting periods of constant length (8 s) each. We then performed a random 80%/20% train/test set split, resulting in 10 and 2 trials, respectively. Trials in the train set were concatenated along the time direction to generate a single time-course input dataset for each modality, while test trials were kept separated.
4.4.3. Training and testing
Pairs of fNIRS and EEG sets were used as inputs of the mSPoC algorithm, while pairs of fNIRS and EEG bandpower sets entered the CCA-based methods. Each method learned a single pair of filters from the train sets. These filters were then applied to each test trial to obtain reconstructed sources via (5) and estimate spatial patterns via (6). We computed the correlation between these reconstructed sources and filters and their corresponding ground-truth counterparts , and averaged across test trials. Figure 10 shows the bias-corrected average of such correlations for each SNR across 50 simulations, computed by first applying Fisher’s z-transformation to the correlations, averaging in z-space, and then inverse-transforming back to the original correlation space (Corey et al., 1998).
4.4.4. Results
The poor performance of standard CCA on the fNIRS side () demonstrates the need for temporal embedding in EEG–fNIRS fusion when using time-domain inputs. tCCA shows a slight improvement in this regard, at expenses of a trade off in the EEG side, which speaks in favor of incorporating some regularization mechanism. This is confirmed by the overall best performance achieved by ElasticNet tCCA on three of the reconstruction tasks. For EEG spatial pattern estimation, however, mSPoC achieves the best overall correlation, which is not surprising considering it was introduced exactly for the purpose of dealing with bandpower EEG data in a way consistent with the linear forward model (4). sstCCA comes second in estimation due to the use of local spatial information, although it performs slightly worse that ElasticNet in the rest of the tasks.
It is worth highlighting that such an assessment of the reviewed source-decomposition methods is far from exhaustive: it represents only a small proof-of-concept comparison and does not sweep across all possible algorithms, parameter settings, or regularization strategies. In particular, we did not study the temporal patterns learned by mSPoC, encoded in , that theoretically should resemble the ground-truth HRF from the fNIRS simulated data (Dähne, 2015). This is a key ingredient, specific to mSPoC, that gives the method an advantage over the others. We leave a more comprehensive evaluation and analysis of these and other topics for future work. To foster transparency and make it easy for others to build on this study, we provide full implementations of each method in Cedalion8 together with a Jupyter notebook that reproduces these results.9 We encourage researchers to extend this work by conducting similar systematic comparisons on additional benchmarks.
5. Current Limitations and Future Directions
5.1. Current challenges
Despite the growing interest and progress in multimodal EEG–fNIRS/DOT fusion for continuous brain imaging, several key limitations still hinder the field’s full potential. First and foremost is the unequal treatment of the two modalities in most studies. EEG data typically benefit from extensive artifact removal techniques and auxiliary recordings (e.g., EOG, EMG), while fNIRS preprocessing often remains limited to motion correction and basic filtering. This imbalance results in suboptimal fusion, where noise and systemic physiology in fNIRS are insufficiently addressed, ultimately degrading the quality of the integrated signal (Scholkmann et al., 2022). From a methodological perspective, non-directional and late-stage fusion strategies dominate the literature, while unsupervised, data-driven symmetric methods, particularly source-decomposition approaches, remain underutilized. While this may not represent a problem in conventional neuroscientific scenarios, it poses both a challenge and opportunity if addressed for continuous and naturalistic brain imaging. Most of the algorithms used in many of the reviewed articles may currently be limited in their ability to uncover rich cross-modal relationships. In this context, source-decomposition methods promise alternatives for greater flexibility, interpretability, and better real-time, single-trial analysis. However, another major challenge lies in the limited availability of publicly accessible fNIRS–EEG datasets and the complete absence of high-density or synthetic extensions. This lack of resources significantly hinders objective benchmarking of fusion methods and undermines reproducibility across studies. While our realistic synthetic dataset represents an initial step toward addressing this issue, there remains a critical need for more diverse and ecologically valid datasets that incorporate controlled artifacts, auxiliary signals, and/or known ground-truth activation patterns.
5.2. Limitations
While our proof-of-concept comparison offers valuable insights, some limitations must be acknowledged. First, the analysis did not include an exhaustive hyperparameter search or systematic sensitivity analysis for each algorithm, so alternative parameter settings may yield different relative performances. We believe that such an investigation could favor the most complex methods, namely (regularized) mSPoC, ssCCA, and kCCA, since they are the ones that depend on a larger parameter space. Second, guided by the review’s main focus on naturalistic scenarios, we only compared data-driven, symmetric fusion methods and did not benchmark them against model-based, asymmetric, or directionally informed techniques, limiting the generalizability of our findings to that particular class of algorithms (see Nia et al., 2025 for a first step in this direction). Finally, although our synthetic dataset was designed to mimic key aspects of continuous EEG–fNIRS recordings, it does not capture the full complexity of real-world signals (Pfurtscheller & Lopes da Silva, 1999). In particular, our synthetic dataset does not model the transient ipsilateral motor-cortex desynchronization that often precedes or accompanies contralateral mu-band ERD in simple movement tasks (Muellbacher et al., 2000). We have also omitted the characteristic post-movement beta rebound (beta-band event-related synchronization) observed in sensorimotor cortex following movement termination (Jurkiewicz et al., 2006). Such omissions may lead to deviations from real data, which we expect to primarily increase variability in the results when comparing fusion methods, rather than systematically biasing performance in favor of one method over another. Moreover, the resting-state background activity was derived from a single subject. While this provides a controlled test bed for benchmarking methods in this manuscript, it does not reflect the inter-individual variability in physiological noise typically encountered in empirical datasets. Incorporating these phenomena would further improve the physiological realism of our simulations and provide a more rigorous testbed for fusion algorithms.
5.3. Future directions
Looking forward, we identify several important directions for future work. First, greater integration of physiological priors and auxiliary measurements, such as heart rate, respiration, and SS channels, will be crucial for improving artifact rejection and isolating neural signals (Scholkmann et al., 2022; Yücel et al., 2021). A second promising direction lies in the development of more sophisticated classical machine-learning fusion methods that integrate spatial and/or context-aware priors. In particular, we identify mSPoC-like approaches, which already incorporate more accurate physiological models, as strong candidates for such enhancements. These methods could be significantly improved by embedding regularization techniques, including not only traditional L1/L2 penalties, but also structured spatial priors derived from anatomical layouts or sensitivity profiles, such as those used in fNIRS/DOT-informed fusion (Cao et al., 2021). Furthermore, these methods could also be enhanced by data-driven constraints, derived from auxiliary recordings, a technique that has been successfully implemented already in the context of single-modality (Lu & Rajapakse, 2000; Yang et al., 2023) and multimodal (J. Liu & Calhoun, 2007) decomposition, although never in fNIRS–EEG fusion.
As a third future direction, it would be highly beneficial to apply these advanced source-decomposition frameworks and the next generation of algorithms built upon them to scientific questions that our survey did not address. The infra-slow EEG–fNIRS coupling is one of these examples (Pfurtscheller et al., 2012; Watson, 2018), in which one could look for shared sources underlying Hz EEG DC shifts and fNIRS hemodynamic oscillations, to test hypotheses about slow neurovascular autoregulation, and resting-state network dynamics (Thompson et al., 2014). Another compelling direction would be the generalization of these source-decomposition methods for the fusion of other modalities, such as fMRI-EEG, exploiting the similarities between (fast) fMRI and fNIRS/DOT in volume space. Moreover, such methods could be further extended for tri-modal fusion, such as fMRI–fNIRS–EEG (Anwar et al., 2016; Muthalib et al., 2013). While not directly applicable to naturalistic scenarios, (fast) fMRI could complement fNIRS rapid hemodynamic acquisition and enhance fusion by its unique full-volume spatial resolution. For such extension, one could build on top of the tools already developed in V. Calhoun, Adali, and Liu (2006), a successful case of tri-modal integration covered in the review, which demonstrated the feasibility of feature-based fusion across functional MRI, structural MRI, and EEG.
Lastly, the field would benefit from community-driven open-source toolkits for multimodal data-driven analysis and datasets. This is one of the main motivators for our development of the Cedalion toolbox (IBS-Lab & Cedalion-Developers, 2024): With Cedalion, we aim to enable state-of-the-art multimodal fNIRS methods development and benchmarking, and to foster reproducibility and standardized evaluation protocols. Regarding community-driven open-source datasets, a public challenge on fNIRS–EEG fusion would be a highly promising step toward this direction by promoting open data sharing and accelerating methodological innovation, much like the successful impact of competitions in the EEG-based BCI field (Blankertz et al., 2006; Tangermann et al., 2012). It is our intention to engage in this direction in the near future by releasing a challenge on single-trial brain activity decoding on a systemic physiology-augmented fNIRS–EEG dataset.
We believe that pursuing all of these avenues combined will advance the field of fNIRS–EEG fusion for continuous brain imaging toward more robust, interpretable, and reproducible brain–body investigations.
Acknowledgments
We thank Bilal Siddique for developing the custom hardware used in our EEG–fNIRS recordings and for his invaluable assistance during data collection. We also thank Nils Harmening for his scripts and expert guidance on EEG head modeling in the synthetic data pipeline.
Appendix A
The following tables contain surveyed articles that used nondirectional fusion strategies (Fig. 2). Stats refers to statistical quantities such as mean, maximum, slope, variance, skewness, kurtosis, or median, applied to raw time-course data. PSR: Phase-Space Reconstruction, TPF: Time-Phase-Frequency features, FBCSP: Filter-Bank Common Spatial Pattern (Ang et al., 2008).
Appendix Table A1.
Surveyed studies that used a concatenation fusion strategy.
| Reference | EEG feature | fNIRS feature | Study objective |
|---|---|---|---|
| Almajidy et al. (2015) | PSD | HbO slope | BCI classification |
| Buccino et al. (2016) | CSP + Bandpower | HbO, HbR CSP + stats | |
| Banville et al. (2017) | Bandpower | HbO, HbR mean | |
| Ge et al. (2017) | PSR + CSP | HbO Hurst exponent | |
| Mandal et al. (2020) | Stats | HbO, HbR stats | |
| Al-Quraishi et al. (2021) | ERD | HbO, HbR stats | |
| Alhudhaif (2021) | Stats | HbO, HbR stats | |
| Z. Wang et al. (2023) | CSP | HbO, HbR CSP | |
| Ergün et al. (2024) | Bandpower | HbO, HbR bandpower | |
| Y. Sun et al. (2020) | PSD | HbO, HbR stats | |
| Fu et al. (2017) | TPF | HbO, HbR raw time-course | |
| X. Lin et al. (2018) | ERP | HbO raw time-course | |
| Firooz and Setarehdan (2019) | Bandpower + PSD | HbO stats | |
| Cao et al. (2022) | PSD and FC | HbO, HbR raw time-course | |
| Gamboa et al. (2022) | Stats | HbO, HbR stats | |
| Zakeri et al. (2023) | Bandpower | HbO, HbR raw time-course | |
| Gemignani and Gervain (2024) | Raw time-course | HbO, HbR raw time-course | |
| Guo et al. (2024) | FBCSP | HbO, HbR CSP + Relief | |
| Y. Meng et al. (2025) | FBCSP | HbO, HbR stats | |
| Abtahi et al. (2020) | PSD | HbO, HbR raw time-course | Clinical assessment |
| Guevara et al. (2020) | Raw time-course | HbO, HbR raw time-course | |
| Güven et al. (2020) | Complexity + ERP | HbO raw time-course | |
| Liang et al. (2022) | ERD/ERS | HbO raw time-course | |
| Maddury (2024) | TPF | HbO, HbR TPF | |
| Y.-C. Jiang et al. (2022) | ERD/ERS | HbO raw time-course | Method presentation |
Appendix Table A2.
Surveyed studies that used a decision-level fusion strategy.
| Reference | EEG feature | fNIRS feature | Study objective |
|---|---|---|---|
| Pfurtscheller (2010) | ERS/ERD | HbO raw time-course | BCI classification |
| Fazli et al. (2012) | FBCSP | HbO, HbR mean | |
| Tomita et al. (2014) | SSVEP | HbO, HbR slope | |
| Putze et al. (2014) | Bandpower + ERP | HbO, HbR mean | |
| Blokland et al. (2014) | PSD | HbO, HbR raw time-course | |
| Koo et al. (2015) | Bandpower | HbO slope | |
| Lee et al. (2015) | FBCSP | HbO, HbR mean | |
| Ahn et al. (2016) | PSD | HbO, HbR raw time-course | |
| Y. Liu et al. (2017) | PSD | HbO, HbR mean | |
| Nguyen et al. (2017) | PSD | HbO, HbR raw time-course | |
| Gupta et al. (2017) | ERS/ERD | HbO, HbR stats | |
| Al-Shargie et al. (2017a) | Bandpower | HbO raw time-course | |
| Shin, Kim, et al. (2018) | FBCSP | HbO, HbR mean + slope | |
| Shin, Kwon, and Im (2018) | FBCSP | HbO, HbR stats | |
| Shin, Müller, and Hwang (2018) | FBCSP | HbO, HbR mean + slope | |
| Rezazadeh et al. (2019) | DWT | THbO, HbR stats | |
| X. Jiang et al. (2019) | Multiple feature paths | Multiple feature paths | |
| Kwon et al. (2020) | FBCSP | HbO, HbR mean | |
| Han et al. (2020) | FBCSP | HbO, HbR mean + slope | |
| Qiu et al. (2022) | TPF | HbO, HbR TPF |
Footnotes
By low-process we mean operations whose outputs are multivariate time-series, the same data type as raw sensor data. Any derived quantity that is not of this kind here is defined as a feature.
Statistics are inflated due to unrelated keyword matches, leaving only 35 deduplicated articles when “DOT” is excluded.
Although adoption has increased in recent years with improved hardware and software support, only 10 of the 124 surveyed studies included SS channels, so they remain rarely recorded overall.
Unless stated otherwise, we use an abstract feature-sample notation whose interpretation depends on the level of analysis, such as single-trial or group-level. For instance, features can refer to channels or voxels, while samples can be time-points, trials, or subjects.
See Cedalion’s implementation for a detailed description of each parameter.
Data and Code Availability
All compared source-decomposition methods can be found in the “sigdecomp” package10 of our Python-based fNIRS/DOT analysis toolbox Cedalion (IBS-Lab & Cedalion-Developers, 2024). We also provide an example notebook containing the methods comparison on our synthetic data to enable the reproducibility of our results.11
Author Contributions
T.C.: Conceptualization; Literature survey; Data collection and curation; Methodology; Software and Synthetic data generation; Formal analysis; Writing—original draft, review, and editing. B.B.: Supervision—final stage; Synthetic data refinement; Writing—review and editing. A.v.L.: Conceptualization; Supervision—all stages; Project administration; Funding acquisition; Writing—original draft, review, and editing.
Funding
The IBS Lab gratefully acknowledges support from BMBF BIFOLD25B, the European Research Council (ERC) under the European Union’s Horizon Europe research and innovation programme (ERC Starting Grant, Grant Agreement No. 101163363) and from the Open Access Publication Fund of TU Berlin.
Declaration of Competing Interest
The authors have no conflicts of interest to declare.
Supplementary Materials
Example notebook containing the methods comparison on our synthetic data: https://github.com/ibs-lab/EEG-fNIRS-Fusion/tree/main/multimodal_fnirs_eeg_sensor_fusion_review.
References
- Abdelnour, A. F., & Huppert, T. (2009). Real-time imaging of human brain function by near-infrared spectroscopy using an adaptive general linear model. NeuroImage, 46(1), 133–143. 10.1016/j.neuroimage.2009.01.033 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abdelnour, F., Genovese, C., & Huppert, T. (2010). Hierarchical Bayesian regularization of reconstructions for diffuse optical tomography using multiple priors. Biomedical Optics Express, 1(4), 1084. 10.1364/BOE.1.001084 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abtahi, M., Bahram Borgheai, S., Jafari, R., Constant, N., Diouf, R., Shahriari, Y., & Mankodiya, K. (2020). Merging fNIRS-EEG brain monitoring and body motion capture to distinguish Parkinsons disease. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 28(6), 1246–1253. 10.1109/TNSRE.2020.2987888 [DOI] [PubMed] [Google Scholar]
- Aghajani, H., Garbey, M., & Omurtag, A. (2017). Measuring mental workload with EEG+fNIRS. Frontiers in Human Neuroscience, 11, 359. 10.3389/fnhum.2017.00359 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ahn, S., & Jun, S. C. (2017). Multi-modal integration of EEG-fNIRS for brain-computer interfaces—Current limitations and future directions. Frontiers in Human Neuroscience, 11, 503. 10.3389/fnhum.2017.00503 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ahn, S., Nguyen, T., Jang, H., Kim, J. G., & Jun, S. C. (2016). Exploring neuro-physiological correlates of drivers’ mental fatigue caused by sleep deprivation using simultaneous EEG, ECG, and fNIRS data. Frontiers in Human Neuroscience, 10, 219. 10.3389/fnhum.2016.00219 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aihara, T., Takeda, Y., Takeda, K., Yasuda, W., Sato, T., Otaka, Y., Hanakawa, T., Honda, M., Liu, M., Kawato, M., Sato, M.-A., & Osu, R. (2012). Cortical current source estimation from electroencephalography in combination with near-infrared spectroscopy as a hierarchical prior. Neuroimage, 59(4), 4006–4021. 10.1016/j.neuroimage.2011.09.087 [DOI] [PubMed] [Google Scholar]
- Alhudhaif, A. (2021). An effective classification framework for brain-computer interface system design based on combining of fNIRS and EEG signals. PeerJ Computer Science, 7, e537. 10.7717/peerj-cs.537 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ali, M. U., Kim, K. S., Kallu, K. D., Zafar, A., & Lee, S. W. (2023). OptEF-BCI: An optimization-based hybrid EEG and fNIRS-brain computer interface. Bioengineering (Basel), 10(5), 608. 10.3390/bioengineering10050608 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ali, M. U., Zafar, A., Kallu, K. D., Masood, H., Mannan, M. M. N., Ibrahim, M. M., Kim, S., & Khan, M. A. (2024). Correlation-filter-based channel and feature selection framework for hybrid EEG-fNIRS BCI applications. IEEE Journal of Biomedical and Health Informatics, 28(6), 3361–3370. 10.1109/JBHI.2023.3294586 [DOI] [PubMed] [Google Scholar]
- Almajidy, R. K., Boudria, Y., Hofmann, U. G., Besio, W., & Mankodiya, K. (2015). Multimodal 2D brain computer interface. In 2015 37th Annual International Conference of the IEEE Engineering in Medicine and Biology Society (EMBC) (pp. 1067–1070). IEEE. 10.1109/EMBC.2015.7318549 [DOI] [PubMed] [Google Scholar]
- AlQahtani, N. J., Al-Naib, I., & Althobaiti, M. (2024). Recent progress on smart lower prosthetic limbs: A comprehensive review on using EEG and fNIRS devices in rehabilitation. Frontiers in Bioengineering and Biotechnology, 12, 1454262. 10.3389/fbioe.2024.1454262 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Al-Quraishi, M. S., Elamvazuthi, I., Tang, T. B., Al-Qurishi, M., Adil, S. H., & Ebrahim, M. (2021). Bimodal data fusion of simultaneous measurements of EEG and fNIRS during lower limb movements. Brain Sciences, 11(6), 713. 10.3390/brainsci11060713 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Al-Shargie, F., Kiguchi, M., Badruddin, N., Dass, S. C., Hani, A. F. M., & Tang, T. B. (2016). Mental stress assessment using simultaneous measurement of EEG and fNIRS. Biomedical Optics Express, 7(10), 3882–3898. 10.1364/BOE.7.003882 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Al-Shargie, F., Tang, T. B., & Kiguchi, M. (2017a). Stress assessment based on decision fusion of EEG and fNIRS signals. IEEE Access, 5, 19889–19896. 10.1109/ACCESS.2017.2754325 [DOI] [Google Scholar]
- Al-Shargie, F., Tang, T. B., & Kiguchi, M. (2017b). Assessment of mental stress effects on prefrontal cortical activities using canonical correlation analysis: An fNIRS-EEG study. Biomedical Optics Express, 8(5), 2583–2598. 10.1364/BOE.8.002583 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Anaya, D., Batra, G., Bracewell, P., Catoen, R., Chakraborty, D., Chevillet, M., Damodara, P., Dominguez, A., Emms, L., Jiang, Z., Kim, E., Klumb, K., Lau, F., Le, R., Li, J., Mateo, B., Matloff, L., Mehta, A., Mugler, E. M., … Yin, A. (2023). Scalable, modular continuous wave functional near-infrared spectroscopy system (Spotlight). Journal of Biomedical Optics, 28(6), 065003. 10.1117/1.JBO.28.6.065003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Andrew, G., Arora, R., Bilmes, J., & Livescu, K. (2013, 17–19 Jun). Deep canonical correlation analysis. In Dasgupta S. & McAllester D. (Eds.), Proceedings of the 30th international conference on machine learning (Vol. 28, pp. 1247–1255). PMLR. https://proceedings.mlr.press/v28/andrew13.html [Google Scholar]
- Ang, K. K., Chin, Z. Y., Zhang, H., & Guan, C. (2008). Filter bank common spatial pattern (FBCSP) in brain-computer interface. In 2008 IEEE International Joint Conference on Neural Networks (IEEE World Congress on Computational Intelligence) (pp. 2390–2397). IEEE. 10.1109/IJCNN.2008.4634130 [DOI] [Google Scholar]
- Anwar, A. R., Muthalib, M., Perrey, S., Galka, A., Granert, O., Wolff, S., Heute, U., Deuschl, G., Raethjen, J., & Muthuraman, M. (2016). Effective connectivity of cortical sensorimotor networks during finger movement tasks: A simultaneous fNIRS, fMRI, EEG study. Brain Topography, 29(5), 645–660. 10.1007/s10548-016-0507-1 [DOI] [PubMed] [Google Scholar]
- Arif, A., Jawad Khan, M., Javed, K., Sajid, H., Rubab, S., Naseer, N., & Irfan Khan, T. (2022). Hemodynamic response detection using integrated EEG-fNIRS-VPA for BCI. Computers, Materials & Continua, 70(1), 535–555. 10.32604/cmc.2022.018318 [DOI] [Google Scholar]
- Arridge, S. R., & Schotland, J. C. (2009). Optical tomography: Forward and inverse problems. Inverse Problems, 25(12), 123010. 10.1088/0266-5611/25/12/123010 [DOI] [Google Scholar]
- Ayaz, H., Shewokis, P. A., Bunce, S., Izzetoglu, K., Willems, B., & Onaral, B. (2012). Optical brain monitoring for operator training and mental workload assessment. NeuroImage, 59(1), 36–47. 10.1016/j.neuroimage.2011.06.023 [DOI] [PubMed] [Google Scholar]
- Balconi, M., & Vanutelli, M. E. (2016). Hemodynamic (fNIRS) and EEG (N200) correlates of emotional inter-species interactions modulated by visual and auditory stimulation. Scientific Reports, 6, 23083. 10.1038/srep23083 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Banville, H., Gupta, R., & Falk, T. H. (2017). Mental task evaluation for hybrid NIRS-EEG brain-computer interfaces. Computational Intelligence and Neuroscience, 2017, 3524208. 10.1155/2017/3524208 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barker, J. W., Aarabi, A., & Huppert, T. J. (2013). Autoregressive model based algorithm for correcting motion and serially correlated errors in fNIRS. Biomedical Optics Express, 4(8), 1366. 10.1364/BOE.4.001366 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Berger, A., Horst, F., Müller, S., Steinberg, F., & Doppelmayr, M. (2019). Current state and future prospects of EEG and fNIRS in robot-assisted gait rehabilitation: A brief review. Frontiers in Human Neuroscience, 13, 172. 10.3389/fnhum.2019.00172 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bie, T., & De Moor, B. (2003). On the regularization of canonical correlation analysis. In 4th International Symposium on Independent Component Analysis (ICA) and Blind Source Separation (ICA2003) (pp. 785–790), Nara, Japan. https://www.rd.ntt/cs/team_project/icl/signal/ica2003/ [Google Scholar]
- Bießmann, F., Meinecke, F. C., Gretton, A., Rauch, A., Rainer, G., Logothetis, N. K., & Müller, K.-R. (2010). Temporal kernel CCA and its application in multimodal neuronal data analysis. Machine Learning, 79(1–2), 5–27. 10.1007/s10994-009-5153-3 [DOI] [Google Scholar]
- Blanco, R., Koba, C., & Crimi, A. (2024). Investigating the interaction between EEG and fNIRS: A multimodal network analysis of brain connectivity. Journal of Computational Science, 82, 102416. 10.1016/j.jocs.2024.102416 [DOI] [Google Scholar]
- Blankertz, B., Muller, K.-R., Krusienski, D., Schalk, G., Wolpaw, J., Schlogl, A., Pfurtscheller, G., Millan, J., Schroder, M., & Birbaumer, N. (2006). The BCI competition III: Validating alternative approaches to actual BCI problems. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 14(2), 153–159. 10.1109/TNSRE.2006.875642 [DOI] [PubMed] [Google Scholar]
- Blokland, Y., Spyrou, L., Thijssen, D., Eijsvogels, T., Colier, W., Floor-Westerdijk, M., Vlek, R., Bruhn, J., & Farquhar, J. (2014). Combined EEG-fNIRS decoding of motor attempt and imagery for brain switch control: An offline study in patients with tetraplegia. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 22(2), 222–229. 10.1109/TNSRE.2013.2292995 [DOI] [PubMed] [Google Scholar]
- Bluestone, A. Y., Abdoulaev, G., Schmitz, C. H., Barbour, R. L., & Hielscher, A. H. (2001). Three-dimensional optical tomography of hemodynamics in the human head. Optics Express, 9(6), 272. 10.1364/OE.9.000272 [DOI] [PubMed] [Google Scholar]
- Boas, D., Brooks, D., Miller, E., DiMarzio, C., Kilmer, M., Gaudette, R., & Zhang, Q. (2001). Imaging the body with diffuse optical tomography. IEEE Signal Processing Magazine, 18(6), 57–75. 10.1109/79.962278 [DOI] [Google Scholar]
- Borgheai, S. B., Deligani, R. J., McLinden, J., Zisk, A., Hosni, S. I., Abtahi, M., Mankodiya, K., & Shahriari, Y. (2019). Multimodal exploration of non-motor neural functions in ALS patients using simultaneous EEG-fNIRS recording. Journal of Neural Engineering, 16(6), 066036. 10.1088/1741-2552/ab456c [DOI] [PMC free article] [PubMed] [Google Scholar]
- Borgheai, S. B., Zisk, A. H., McLinden, J., Mcintyre, J., Sadjadi, R., & Shahriari, Y. (2024). Multimodal pre-screening can predict BCI performance variability: A novel subject-specific experimental scheme. Computers in Biology and Medicine, 168(107658), 107658. 10.1016/j.compbiomed.2023.107658 [DOI] [PubMed] [Google Scholar]
- Bourguignon, N. J., Bue, S. L., Guerrero-Mosquera, C., & Borragán, G. (2022). Bimodal EEG-fNIRS in neuroergonomics. Current evidence and prospects for future research. Frontiers in Neuroergonomics, 3, 934234. 10.3389/fnrgo.2022.934234 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Buccino, A. P., Keles, H. O., & Omurtag, A. (2016). Hybrid EEG-fNIRS asynchronous Brain-computer interface for multiple motor tasks. PLoS One, 11(1), e0146610. 10.1371/journal.pone.0146610 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Calhoun, V., Adali, T., Giuliani, N., Pekar, J., Kiehl, K., & Pearlson, G. (2006). Method for multimodal analysis of independent source differences in schizophrenia: Combining gray matter structural and auditory oddball functional data. Human Brain Mapping, 27(1), 47–62. 10.1002/hbm.20166 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Calhoun, V., Adali, T., & Liu, J. (2006). A feature-based approach to combine functional MRI, structural MRI and EEG brain imaging data. In 2006 International Conference of the IEEE Engineering in Medicine and Biology Society (pp. 3672–3675). IEEE. 10.1109/IEMBS.2006.259810 [DOI] [PubMed] [Google Scholar]
- Calhoun, V., Adali, T., Pearlson, G., & Kiehl, K. (2006). Neuronal chronometry of target detection: Fusion of hemodynamic and event-related potential data. NeuroImage, 30(2), 544–553. 10.1016/j.neuroimage.2005.08.060 [DOI] [PubMed] [Google Scholar]
- Calhoun, V. D., Adalı, T., Kiehl, K. A., Astur, R., Pekar, J. J., & Pearlson, G. D. (2006). A method for multitask fMRI data fusion applied to schizophrenia. Human Brain Mapping, 27(7), 598–610. 10.1002/hbm.20204 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cao, J., Bulger, E., Shinn-Cunningham, B., Grover, P., & Kainerstorfer, J. M. (2023). Diffuse optical tomography spatial prior for EEG source localization in human visual cortex. NeuroImage, 277, 120210. 10.1016/j.neuroimage.2023.120210 [DOI] [PubMed] [Google Scholar]
- Cao, J., Garro, E. M., & Zhao, Y. (2022). EEG/fNIRS based workload classification using functional brain connectivity and machine learning. Sensors (Basel), 22(19), 7623. 10.3390/s22197623 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cao, J., Huppert, T. J., Grover, P., & Kainerstorfer, J. M. (2021). Enhanced spatiotemporal resolution imaging of neuronal activity using joint electroencephalography and diffuse optical tomography. Neurophotonics, 8(1), 015002. 10.1117/1.NPh.8.1.015002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Casson, A., Yates, D., Smith, S., Duncan, J., & Rodriguez-Villegas, E. (2010). Wearable electroencephalography. IEEE Engineering in Medicine and Biology Magazine, 29(3), 44–56. 10.1109/MEMB.2010.936545 [DOI] [PubMed] [Google Scholar]
- Chalia, M., Lee, C. W., Dempsey, L. A., Edwards, A. D., Singh, H., Michell, A. W., Everdell, N. L., Hill, R. W., Hebden, J. C., Austin, T., & Cooper, R. J. (2016). Hemodynamic response to burst-suppressed and discontinuous electroencephalography activity in infants with hypoxic ischemic encephalopathy. Neurophotonics, 3(3), 1. 10.1117/1.NPh.3.3.031408 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chang, C.-Y., Hsu, S.-H., Pion-Tonachini, L., & Jung, T.-P. (2020). Evaluation of artifact subspace reconstruction for automatic artifact components removal in multi-channel EEG recordings. IEEE Transactions on Biomedical Engineering, 67(4), 1114–1121. 10.1109/TBME.2019.2930186 [DOI] [PubMed] [Google Scholar]
- Charly, D. T., & Onabid, M. (2024). EF-CorrCA: A multi-modal EEG-fNIRS subject independent model to assess speech quality on brain activity using correlated component analysis. Cognitive Computation and Systems, 6(1–3), 36–48. 10.1049/ccs2.12111 [DOI] [Google Scholar]
- Chen, J., Bushman, F. D., Lewis, J. D., Wu, G. D., & Li, H. (2013). Structure-constrained sparse canonical correlation analysis with an application to microbiome data analysis. Biostatistics, 14(2), 244–258. 10.1093/biostatistics/kxs038 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen, J., Xia, Y., Zhou, X., Vidal Rosas, E., Thomas, A., Loureiro, R., Cooper, R. J., Carlson, T., & Zhao, H. (2023). fNIRS-EEG BCIs for motor rehabilitation: A review. Bioengineering, 10(12), 1393. 10.3390/bioengineering10121393 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen, L.-C., Sandmann, P., Thorne, J. D., Herrmann, C. S., & Debener, S. (2015). Association of concurrent fNIRS and EEG signatures in response to auditory and visual stimuli. Brain Topography, 28(5), 710–725. 10.1007/s10548-015-0424-8 [DOI] [PubMed] [Google Scholar]
- Chen, Y., Tang, J., Chen, Y., Farrand, J., Craft, M. A., Carlson, B. W., & Yuan, H. (2020). Amplitude of fNIRS resting-state global signal is related to EEG vigilance measures: A simultaneous fNIRS and EEG study. Frontiers in Neuroscience, 14, 560878. 10.3389/fnins.2020.560878 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chiarelli, A. M., Maclin, E. L., Fabiani, M., & Gratton, G. (2015). A kurtosis-based wavelet algorithm for motion artifact correction of fNIRS data. NeuroImage, 112, 128–137. 10.1016/j.neuroimage.2015.02.057 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chiarelli, A. M., Perpetuini, D., Croce, P., Filippini, C., Cardone, D., Rotunno, L., Anzoletti, N., Zito, M., Zappasodi, F., & Merla, A. (2021). Evidence of neurovascular Un-coupling in mild Alzheimer’s disease through multimodal EEG-fNIRS and multivariate analysis of resting-state data. Biomedicines, 9(4), 337. 10.3390/biomedicines9040337 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chu, H., Cao, Y., Jiang, J., Yang, J., Huang, M., Li, Q., Jiang, C., & Jiao, X. (2022). Optimized electroencephalogram and functional near-infrared spectroscopy-based mental workload detection method for practical applications. BioMedical Engineering OnLine, 21(1), 9. 10.1186/s12938-022-00980-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chung, F. R. (1997). Spectral graph theory (Vol. 92). American Mathematical Soc. https://www.ams.org/books/cbms/092/ [Google Scholar]
- Cicalese, P. A., Li, R., Ahmadi, M. B., Wang, C., Francis, J. T., Selvaraj, S., Schulz, P. E., & Zhang, Y. (2020). An EEG-fNIRS hybridization technique in the four-class classification of Alzheimer’s disease. Journal of Neuroscience Methods, 336, 108618. 10.1016/j.jneumeth.2020.108618 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Corey, D. M., Dunlap, W. P., & Burke, M. J. (1998). Averaging correlations: Expected values and bias in combined Pearson rs and Fisher’s z transformations. The Journal of General Psychology, 125(3), 245–261. 10.1080/00221309809595548 [DOI] [Google Scholar]
- Correa, N. M., Adali, T., Li, Y.-O., & Calhoun, V. D. (2010). Canonical correlation analysis for data fusion and group inferences. IEEE Signal Processing Magazine, 27(4), 39–50. 10.1109/MSP.2010.936725 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dähne, S. (2015). Decomposition Methods for the Fusion of Multimodal Neuroimaging Data. [Doctoral thesis, Institute of Software Engineering and Theoretical Computer Science]. 10.14279/depositonce-4683 [DOI]
- Dähne, S., Biessmann, F., Meinecke, F. C., Mehnert, J., Fazli, S., & Muller, K.-R. (2013). Integration of multivariate data streams with bandpower signals. IEEE Transactions on Multimedia, 15(5), 1001–1013. 10.1109/TMM.2013.2250267 [DOI] [Google Scholar]
- Dashtestani, H., Miguel, H. O., Condy, E. E., Zeytinoglu, S., Millerhagen, J. B., Debnath, R., Smith, E., Adali, T., Fox, N. A., & Gandjbakhche, A. H. (2022). Structured sparse multiset canonical correlation analysis of simultaneous fNIRS and EEG provides new insights into the human action-observation network. Scientific Reports, 12(1), 6878. 10.1038/s41598-022-10942-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dehghani, H., Eames, M. E., Yalavarthy, P. K., Davis, S. C., Srinivasan, S., Carpenter, C. M., Pogue, B. W., & Paulsen, K. D. (2009). Near infrared optical tomography using NIRFAST: Algorithm for numerical model and image reconstruction. Communications in Numerical Methods in Engineering, 25(6), 711–732. 10.1002/cnm.1162 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deligani, R. J., Borgheai, S. B., McLinden, J., & Shahriari, Y. (2021). Multimodal fusion of EEG-fNIRS: A mutual information-based hybrid classification framework. Biomedical Optics Express, 12(3), 1635–1650. 10.1364/BOE.413666 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Diamond, S. G., Huppert, T. J., Kolehmainen, V., Franceschini, M. A., Kaipio, J. P., Arridge, S. R., & Boas, D. A. (2006). Dynamic physiological modeling for functional diffuse optical tomography. NeuroImage, 30(1), 88–101. 10.1016/j.neuroimage.2005.09.016 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dissanayake, T., Müller, K. R., & Luhmann, A. V. (2025). Deep learning from diffuse optical oximetry time-series: An fnirs-focused review of recent advancements and future directions. IEEE Reviews in Biomedical Engineering, 19. 10.1109/RBME.2025.3617858 [DOI] [PubMed] [Google Scholar]
- Eastmond, C., Subedi, A., De, S., & Intes, X. (2022). Deep learning in fNIRS: A review. Neurophotonics, 9(04), 041411. 10.1117/1.NPh.9.4.041411 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eggebrecht, A. T., Ferradal, S. L., Robichaux-Viehoever, A., Hassanpour, M. S., Dehghani, H., Snyder, A. Z., Hershey, T., & Culver, J. P. (2014). Mapping distributed brain function and networks with diffuse optical tomography. Nature Photonics, 8(6), 448–454. 10.1038/nphoton.2014.107 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ergün, E., & Aydemir, O. (2023). A hybrid BCI using singular value decomposition values of the fast walsh–hadamard transform coefficients. IEEE Transactions on Cognitive and Developmental Systems, 15(2), 454–463. 10.1109/TCDS.2020.3028785 [DOI] [Google Scholar]
- Ergün, E., Aydemir, Ö., & Korkmaz, O. E. (2024). Investigating the informative brain region in multiclass electroencephalography and near infrared spectroscopy based BCI system using band power based features. Computer Methods in Biomechanics and Biomedical Engineering, 28(10), 1655–1670. 10.1080/10255842.2024.2333924 [DOI] [PubMed] [Google Scholar]
- Fazli, S., Mehnert, J., Steinbrink, J., Curio, G., Villringer, A., Müller, K.-R., & Blankertz, B. (2012). Enhanced performance by a hybrid NIRS-EEG brain computer interface. NeuroImage, 59(1), 519–529. 10.1016/j.neuroimage.2011.07.084 [DOI] [PubMed] [Google Scholar]
- Ferradal, S. L., Liao, S. M., Eggebrecht, A. T., Shimony, J. S., Inder, T. E., Culver, J. P., & Smyser, C. D. (2016). Functional imaging of the developing brain at the bedside using diffuse optical tomography. Cerebral Cortex, 26(4), 1558–1568. 10.1093/cercor/bhu320 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Firooz, S., & Setarehdan, S. K. (2019). IQ estimation by means of EEG-fNIRS recordings during a logical-mathematical intelligence test. Computers in Biology and Medicine, 110, 218–226. 10.1016/j.compbiomed.2019.05.017 [DOI] [PubMed] [Google Scholar]
- Fishburn, F. A., Ludlum, R. S., Vaidya, C. J., & Medvedev, A. V. (2019). Temporal Derivative Distribution Repair (TDDR): A motion correction method for fNIRS. NeuroImage, 184, 171–179. 10.1016/j.neuroimage.2018.09.025 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fitzgibbon, S. P., Lewis, T. W., Powers, D. M. W., Whitham, E. W., Willoughby, J. O., & Pope, K. J. (2013). Surface laplacian of central scalp electrical signals is insensitive to muscle contamination. IEEE Transactions on Biomedical Engineering, 60(1), 4–9. 10.1109/TBME.2012.2195662 [DOI] [PubMed] [Google Scholar]
- Franceschini, M. A., Joseph, D. K., Huppert, T. J., Diamond, S. G., & Boas, D. A. (2006). Diffuse optical imaging of the whole head. Journal of Biomedical Optics, 11(5), 054007. 10.1117/1.2363365 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Friston, K., Mechelli, A., Turner, R., & Price, C. (2000). Nonlinear responses in fMRI: The balloon model, volterra kernels, and other hemodynamics. NeuroImage, 12(4), 466–477. 10.1006/nimg.2000.0630 [DOI] [PubMed] [Google Scholar]
- Fu, Y., Xiong, X., Jiang, C., Xu, B., Li, Y., & Li, H. (2017). Imagined hand clenching force and speed modulate brain activity and are classified by NIRS combined with EEG. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 25(9), 1641–1652. 10.1109/TNSRE.2016.2627809 [DOI] [PubMed] [Google Scholar]
- Fukumizu, K., & Gretton, A. (2007). Statistical consistency of kernel canonical correlation analysis. Journal of Machine Learning Research, 8, 361–383. 10.1109/mlsp.2007.4414286 [DOI] [Google Scholar]
- Funane, T., Atsumori, H., Katura, T., Obata, A. N., Sato, H., Tanikawa, Y., Okada, E., & Kiguchi, M. (2014). Quantitative evaluation of deep and shallow tissue layers’ contribution to fNIRS signal using multi-distance optodes and independent component analysis. NeuroImage, 85, 150–165. 10.1016/j.neuroimage.2013.02.026 [DOI] [PubMed] [Google Scholar]
- Gagnon, L., Perdue, K., Greve, D. N., Goldenholz, D., Kaskhedikar, G., & Boas, D. A. (2011). Improved recovery of the hemodynamic response in diffuse optical imaging using short optode separations and state-space modeling. NeuroImage, 56(3), 1362–1371. 10.1016/j.neuroimage.2011.03.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gamboa, P., Varandas, R., Rodrigues, J., Cepeda, C., Quaresma, C., & Gamboa, H. (2022). Attention classification based on biosignals during standard cognitive tasks for occupational domains. Computers, 11, 49. 10.3390/computers11040049 [DOI] [Google Scholar]
- Gao, C., Huang, H., Zhan, J., Li, W., Li, Y., Li, J., Zhou, J., Wang, Y., Jiang, Z., Chen, W., Zhu, Y., Zhuo, Y., & Wu, K. (2024). Adaptive changes in neurovascular properties with binocular accommodation functions in myopic participants by 3D visual training: An EEG and fNIRS study. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 32, 2749–2758. 10.1109/TNSRE.2024.3434492 [DOI] [PubMed] [Google Scholar]
- Gao, T., Chen, C., Liang, G., Ran, Y., Huang, Q., Liao, Z., He, B., Liu, T., Tang, X., Chen, H., & Fan, Y. (2025). Feature fusion analysis approach based on synchronous EEG-fNIRS signals: Application in etomidate use disorder individuals. Biomedical Optics Express, 16(2), 382. 10.1364/BOE.542078 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao, Y., Rogers, D., Von Lühmann, A., Ortega-Martinez, A., Boas, D. A., & Yücel, M. A. (2023). Short-separation regression incorporated diffuse optical tomography image reconstruction modeling for high-density functional near-infrared spectroscopy. Neurophotonics, 10(2), 025007. 10.1117/1.NPh.10.2.025007 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao, Y., Jia, B., Houston, M., & Zhang, Y. (2023). Hybrid eeg-fnirs brain computer interface based on common spatial pattern by using eeg-informed general linear model. IEEE Transactions on Instrumentation and Measurement, 72, 1–10. 10.1109/TIM.2023.327650937323850 [DOI] [Google Scholar]
- Gao, Y., Liu, H., Fang, F., & Zhang, Y. (2023). Classification of working memory loads via assessing broken detailed balance of EEG-FNIRS neurovascular coupling measures. IEEE Transactions on Biomedical Engineering, 70(3), 877–887. 10.1109/TBME.2022.3204718 [DOI] [PubMed] [Google Scholar]
- Ge, S., Wang, P., Liu, H., Lin, P., Gao, J., Wang, R., Iramina, K., Zhang, Q., & Zheng, W. (2019). Neural activity and decoding of action observation using combined EEG and fNIRS measurement. Frontiers in Human Neuroscience, 13, 357. 10.3389/fnhum.2019.00357 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ge, S., Yang, Q., Wang, R., Lin, P., Gao, J., Leng, Y., Yang, Y., & Wang, H. (2017). A brain-computer interface based on a few-channel EEG-fNIRS bimodal system. IEEE Access, 5, 208–218. 10.1109/ACCESS.2016.2637409 [DOI] [Google Scholar]
- Gemignani, J., & Gervain, J. (2024). A within-subject multimodal NIRS-EEG classifier for infant data. Sensors (Basel), 24(13), 4161. 10.3390/s24134161 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gentile, E., Brunetti, A., Ricci, K., Delussi, M., Bevilacqua, V., & de Tommaso, M. (2020). Mutual interaction between motor cortex activation and pain in fibromyalgia: EEG-fNIRS study. PLoS One, 15(1), e0228158. 10.1371/journal.pone.0228158 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ghasimi, A., & Shamekhi, S. (2024). Assessment of cognitive workload using simultaneous EEG and fNIRS: A comparison of feature combinations. Computers and Electrical Engineering, 119, 109619. 10.1016/j.compeleceng.2024.109619 [DOI] [Google Scholar]
- Giacometti, P., & Diamond, S. G. (2014). Correspondence of electroencephalography and near-infrared spectroscopy sensitivities to the cerebral cortex using a high-density layout. Neurophotonics, 1(2), 025001. 10.1117/1.NPh.1.2.025001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Glover, G. H. (1999). Deconvolution of impulse response in event-related BOLD fMRI. NeuroImage, 9(4), 416–429. 10.1006/nimg.1998.0419 [DOI] [PubMed] [Google Scholar]
- Goodwin, J. R., Gaudet, C. R., & Berger, A. J. (2014). Short-channel functional near-infrared spectroscopy regressions improve when source-detector separation is reduced. Neurophotonics, 1(1), 015002. 10.1117/1.NPh.1.1.015002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Govindan, R. B., Massaro, A., Chang, T., Vezina, G., & du Plessis, A. (2016). A novel technique for quantitative bedside monitoring of neurovascular coupling. Journal of Neuroscience Methods, 259, 135–142. 10.1016/j.jneumeth.2015.11.025 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gramfort, A., Luessi, M., Larson, E., Engemann, D. A., Strohmeier, D., Brodbeck, C., Parkkonen, L., & Hämäläinen, M. S. (2014). Mne software for processing MEG and EEG data. NeuroImage, 86, 446–460. 10.1016/j.neuroimage.2013.10.027 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guevara, E., Flores-Castro, J.-A., Peng, K., Nguyen, D. K., Lesage, F., Pouliot, P., & Rosas-Romero, R. (2020). Prediction of epileptic seizures using fnirs and machine learning. Journal of Intelligent & Fuzzy Systems, 38(2), 2055–2068. 10.3233/JIFS-190738 [DOI] [Google Scholar]
- Guo, M., Feng, L., Chen, X., Li, M., & Xu, G. (2024). A novel strategy for differentiating motor imagination brain-computer interface tasks by fusing EEG and functional near-infrared spectroscopy signals. Biomedical Signal Processing and Control, 95, 106448. 10.1016/j.bspc.2024.106448 [DOI] [Google Scholar]
- Gupta, R., Banville, H. J., & Falk, T. H. (2017). Multimodal physiological quality-of-experience assessment of text-to-speech systems. IEEE Journal of Selected Topics in Signal Processing, 11(1), 22–36. 10.1109/JSTSP.2016.2638538 [DOI] [Google Scholar]
- Güven, A., Altınkaynak, M., Dolu, N., İzzetoğlu, M., Pektaş, F., Özmen, S., Demirci, E., & Batbat, T. (2020). Combining functional near-infrared spectroscopy and EEG measurements for the diagnosis of attention-deficit hyperactivity disorder. Neural Computing and Applications, 32(12), 8367–8380. 10.1007/s00521-019-04294-7 [DOI] [Google Scholar]
- Habermehl, C., Holtze, S., Steinbrink, J., Koch, S. P., Obrig, H., Mehnert, J., & Schmitz, C. H. (2012). Somatosensory activation of two fingers can be discriminated with ultrahigh-density diffuse optical tomography. NeuroImage, 59(4), 3201–3211. 10.1016/j.neuroimage.2011.11.062 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Han, C.-H., Muller, K.-R., & Hwang, H.-J. (2020). Enhanced performance of a brain switch by simultaneous use of EEG and NIRS data for asynchronous brain-computer interface. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 28(10), 2102–2112. 10.1109/TNSRE.2020.3017167 [DOI] [PubMed] [Google Scholar]
- Handwerker, D. A., Ollinger, J. M., & D’Esposito, M. (2004). Variation of BOLD hemodynamic responses across subjects and brain regions and their effects on statistical analyses. NeuroImage, 21(4), 1639–1651. 10.1016/j.neuroimage.2003.11.029 [DOI] [PubMed] [Google Scholar]
- Hardoon, D. R., Szedmak, S., & Shawe-Taylor, J. (2004). Canonical correlation analysis: An overview with application to learning methods. Neural Computation, 16(12), 2639–2664. 10.1162/0899766042321814 [DOI] [PubMed] [Google Scholar]
- Harmening, N., & Miklody, D. (2022. November). MRIsegmentation (Version 1.1). 10.5281/zenodo.7357674 [DOI]
- Hasan, M. A. H., Khan, M. U., & Mishra, D. (2020). A computationally efficient method for hybrid EEG-fNIRS BCI based on the Pearson correlation. BioMed Research International, 2020, 1838140. 10.1155/2020/1838140 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hastie, T., Tibshirani, R., & Wainwright, M. (2015). Statistical learning with sparsity: The lasso and generalizations. Chapman & Hall/CRC. 10.1201/b18401 [DOI] [Google Scholar]
- Haufe, S., Meinecke, F., Görgen, K., Dähne, S., Haynes, J.-D., Blankertz, B., & Bießmann, F. (2014). On the interpretation of weight vectors of linear models in multivariate neuroimaging. NeuroImage, 87, 96–110. 10.1016/j.neuroimage.2013.10.067 [DOI] [PubMed] [Google Scholar]
- Hauk, O. (2004). Keep it simple: A case for using classical minimum norm estimation in the analysis of EEG and MEG data. NeuroImage, 21(4), 1612–1621. 10.1016/j.neuroimage.2003.12.018 [DOI] [PubMed] [Google Scholar]
- He, X., & Niyogi, P. (2003). Locality preserving projections. In Proceedings of the 16th International Conference on Neural Information Processing Systems (pp. 153–160). Vancouver, Canada. https://dl.acm.org/doi/10.5555/2981345.2981365 [Google Scholar]
- Holland, P. W., & Welsch, R. E. (1977). Robust regression using iteratively reweighted least-squares. Communications in Statistics—Theory and Methods, 6(9), 813–827. 10.1080/03610927708827533 [DOI] [Google Scholar]
- Holmes, C. J., Hoge, R., Collins, L., Woods, R., Toga, A. W., & Evans, A. C. (1998). Enhancement of MR images using registration for signal averaging. Journal of Computer Assisted Tomography, 22(2), 324–333. 10.1097/00004728-199803000-00032 [DOI] [PubMed] [Google Scholar]
- Hong, K.-S., Khan, M. J., & Hong, M. J. (2018). Feature extraction and classification methods for hybrid fNIRS-EEG brain-computer interfaces. Frontiers in Human Neuroscience, 12, 246. 10.3389/fnhum.2018.00246 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hosni, S. M. I., Borgheai, S. B., McLinden, J., Zhu, S., Huang, X., Ostadabbas, S., & Shahriari, Y. (2022). A graph-based nonlinear dynamic characterization of motor imagery toward an enhanced hybrid BCI. Neuroinformatics, 20(4), 1169–1189. 10.1007/s12021-022-09595-2 [DOI] [PubMed] [Google Scholar]
- Hossain, M. S., Reaz, M. B. I., Chowdhury, M. E. H., Ali, S. H. M., Bakar, A. A. A., Kiranyaz, S., Khandakar, A., Alhatou, M., & Habib, R. (2022). Motion artifacts correction from EEG and fNIRS signals using novel multiresolution analysis. IEEE Access, 10, 29760–29777. 10.1109/ACCESS.2022.3159155 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hotelling, H. (1936). Relations between two sets of variates. Biometrika, 28(3–4), 321–377. 10.1093/biomet/28.3-4.321 [DOI] [Google Scholar]
- Hou, M., Zhang, X., Chen, G., Huang, L., & Sun, Y. (2024). Emotion recognition based on a EEG–fNIRS hybrid brain network in the source space. Brain Sciences, 14(12), 1166. 10.3390/brainsci14121166 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huppert, T., Barker, J., Schmidt, B., Walls, S., & Ghuman, A. (2017). Comparison of group-level, source localized activity for simultaneous functional near-infrared spectroscopy-magnetoencephalography and simultaneous fNIRS-fMRI during parametric median nerve stimulation. Neurophotonics, 4(1), 015001. 10.1117/1.NPh.4.1.015001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huppert, T. J. (2016). Commentary on the statistical properties of noise and its implication on general linear models in functional near-infrared spectroscopy. Neurophotonics, 3(1), 010401. 10.1117/1.NPh.3.1.010401 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huppert, T. J., Diamond, S. G., Franceschini, M. A., & Boas, D. A. (2009). Homer: A review of time-series analysis methods for near-infrared spectroscopy of the brain. Applied Optics, 48(10), D280–D298. 10.1364/AO.48.00D280 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huppert, T., Hoge, R., Diamond, S., Franceschini, M., & Boas, D. (2006). A temporal comparison of BOLD, ASL, and NIRS hemodynamic responses to motor stimuli in adult humans. NeuroImage, 29(2), 368–382. 10.1016/j.neuroimage.2005.08.065 [DOI] [PMC free article] [PubMed] [Google Scholar]
- IBS-Lab, & Cedalion-Developers. (2024, September). Cedalion Python Toolbox (Version 1.0.0). https://github.com/ibs-lab/cedalion
- Ieong, H. F.-H., Gao, F., & Yuan, Z. (2019). Machine learning: Assessing neurovascular signals in the prefrontal cortex with non-invasive bimodal electro-optical neuroimaging in opiate addiction. Scientific Reports, 9(1), 18262. 10.1038/s41598-019-54316-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Islam, M. K., Rastegarnia, A., & Yang, Z. (2016). Methods for artifact detection and removal from scalp EEG: A review. Neurophysiologie Clinique, 46(4–5), 287–305. 10.1016/j.neucli.2016.07.002 [DOI] [PubMed] [Google Scholar]
- Jahani, S., Setarehdan, S. K., Boas, D. A., & Yücel, M. A. (2018). Motion artifact detection and correction in functional near-infrared spectroscopy: A new hybrid method based on spline interpolation method and Savitzky–Golay filtering. Neurophotonics, 5(1), 1. 10.1117/1.NPh.5.1.015003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang, X., Gu, X., Xu, K., Ren, H., & Chen, W. (2019). Independent decision path fusion for bimodal asynchronous brain–computer interface to discriminate multiclass mental states. IEEE Access, 7, 165303–165317. 10.1109/ACCESS.2019.2953535 [DOI] [Google Scholar]
- Jiang, Y.-C., Ma, R., Qi, S., Ge, S., Sun, Z., Li, Y., Song, J., & Zhang, M. (2022). Characterization of bimanual cyclical tasks from single-trial EEG-fNIRS measurements. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 30, 146–156. 10.1109/TNSRE.2022.3144216 [DOI] [PubMed] [Google Scholar]
- Jurkiewicz, M. T., Gaetz, W. C., Bostan, A. C., & Cheyne, D. (2006). Post-movement beta rebound is generated in motor cortex: Evidence from neuromagnetic recordings. NeuroImage, 32, 1281–1289. 10.1016/j.neuroimage.2006.06.005 [DOI] [PubMed] [Google Scholar]
- Kaga, Y., Ueda, R., Tanaka, M., Kita, Y., Suzuki, K., Okumura, Y., Egashira, Y., Shirakawa, Y., Mitsuhashi, S., Kitamura, Y., Nakagawa, E., Yamashita, Y., & Inagaki, M. (2020). Executive dysfunction in medication-naive children with ADHD: A multi-modal fNIRS and EEG study. Brain & Development, 42(8), 555–563. 10.1016/j.braindev.2020.05.007 [DOI] [PubMed] [Google Scholar]
- Keles, H. O., Barbour, R. L., & Omurtag, A. (2016). Hemodynamic correlates of spontaneous neural activity measured by human whole-head resting state EEG plus fNIRS. NeuroImage, 138, 76–87. 10.1016/j.neuroimage.2016.05.058 [DOI] [PubMed] [Google Scholar]
- Kettenring, J. R. (1971). Canonical analysis of several sets of variables. Biometrika, 58(3), 433–451. 10.1093/biomet/58.3.433 [DOI] [Google Scholar]
- Khan, H., Naseer, N., Yazidi, A., Eide, P. K., Hassan, H. W., & Mirtaheri, P. (2021). Analysis of human gait using hybrid EEG-fNIRS-based BCI system: A review. Frontiers in Human Neuroscience, 14, 613254. 10.3389/fnhum.2020.613254 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Khan, M. J., Ghafoor, U., & Hong, K.-S. (2018). Early detection of hemodynamic responses using EEG: A hybrid EEG-fNIRS study. Frontiers in Human Neuroscience, 12, 479. 10.3389/fnhum.2018.00479 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Khan, M. U., & Hasan, M. A. H. (2020). Hybrid EEG-fNIRS BCI fusion using multi-resolution singular value decomposition (MSVD). Frontiers in Human Neuroscience, 14, 599802. 10.3389/fnhum.2020.599802 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kirilina, E., Jelzow, A., Heine, A., Niessing, M., Wabnitz, H., Brühl, R., Ittermann, B., Jacobs, A. M., & Tachtsidis, I. (2012). The physiological origin of task-evoked systemic artefacts in functional near infrared spectroscopy. NeuroImage, 61(1), 70–81. 10.1016/j.neuroimage.2012.02.074 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kirilina, E., Yu, N., Jelzow, A., Wabnitz, H., Jacobs, A. M., & Tachtsidis, I. (2013). Identifying and quantifying main components of physiological noise in functional near infrared spectroscopy on the prefrontal cortex. Frontiers in Human Neuroscience, 7, 864. 10.3389/fnhum.2013.00864 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kiverstein, J., & Miller, M. (2015). The embodied brain: Towards a radical embodied cognitive neuroscience. Frontiers in Human Neuroscience, 9, 237. 10.3389/fnhum.2015.00237 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koo, B., Lee, H.-G., Nam, Y., Kang, H., Koh, C. S., Shin, H.-C., & Choi, S. (2015). A hybrid NIRS-EEG system for self-paced brain computer interface with online motor imagery. Journal of Neuroscience Methods, 244, 26–32. 10.1016/j.jneumeth.2014.04.016 [DOI] [PubMed] [Google Scholar]
- Kwon, J., Shin, J., & Im, C.-H. (2020). Toward a compact hybrid brain-computer interface (BCI): Performance evaluation of multi-class hybrid EEG-fNIRS BCIs with limited number of channels. PLoS One, 15(3), e0230491. 10.1371/journal.pone.0230491 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Larson, E., Gramfort, A., Engemann, D. A., Leppakangas, J., Brodbeck, C., Jas, M., Brooks, T. L., Sassenhagen, J., McCloy, D., Luessi, M., King, J.-R., Höchenberger, R., Goj, R., Brunner, C., Favelier, G., van Vliet, M., Wronkiewicz, M., Rockhill, A., Holdgraf, C., … luzpaz. (2024, December). Mne-python (Version v1.9.0). Zenodo. 10.5281/zenodo.14519545 [DOI]
- Lee, M.-H., Fazli, S., Mehnert, J., & Lee, S.-W. (2015). Subject-dependent classification for robust idle state detection using multi-modal neuroimaging and data-fusion techniques in BCI. Pattern Recognition, 48(8), 2725–2737. 10.1016/j.patcog.2015.03.010 [DOI] [Google Scholar]
- Li, R., Li, S., Roh, J., Wang, C., & Zhang, Y. (2020). Multimodal neuroimaging using concurrent EEG/fNIRS for poststroke recovery assessment: An exploratory study. Neurorehabil. Neural Repair, 34(12), 1099–1110. 10.1177/1545968320969937 [DOI] [PubMed] [Google Scholar]
- Li, R., Nguyen, T., Potter, T., & Zhang, Y. (2019). Dynamic cortical connectivity alterations associated with Alzheimer’s disease: An EEG and fNIRS integration study. NeuroImage: Clinical, 21, 101622. 10.1016/j.nicl.2018.101622 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li, R., Potter, T., Huang, W., & Zhang, Y. (2017). Enhancing performance of a hybrid EEG-fNIRS system using channel selection and early temporal features. Frontiers in Human Neuroscience, 11, 462. 10.3389/fnhum.2017.00462 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li, R., Yang, D., Fang, F., Hong, K.-S., Reiss, A. L., & Zhang, Y. (2022). Concurrent fNIRS and EEG for brain function investigation: A systematic, methodology-focused review. Sensors, 22(15), 5865. 10.3390/s22155865 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li, R., Zhao, C., Wang, C., Wang, J., & Zhang, Y. (2020). Enhancing fNIRS analysis using EEG rhythmic signatures: An EEG-informed fNIRS analysis study. IEEE Transactions on Biomedical Engineering, 67(10), 2789–2797. 10.1109/TBME.2020.2971679 [DOI] [PubMed] [Google Scholar]
- Li, X., Fang, F., Li, R., & Zhang, Y. (2022). Functional brain controllability alterations in stroke. Frontiers in Bioengineering and Biotechnology, 10, 925970. 10.3389/fbioe.2022.925970 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liang, J., Song, Y., Belkacem, A. N., Li, F., Liu, S., Chen, X., Wang, X., Wang, Y., & Wan, C. (2022). Prediction of balance function for stroke based on EEG and fNIRS features during ankle dorsiflexion. Frontiers in Neuroscience, 16, 968928. 10.3389/fnins.2022.968928 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lin, C.-T., King, J.-T., Chuang, C.-H., Ding, W., Chuang, W.-Y., Liao, L.-D., & Wang, Y.-K. (2020). Exploring the brain responses to driving fatigue through simultaneous EEG and fNIRS measurements. International Journal of Neural Systems, 30(1), 19500187. 10.1142/S0129065719500187 [DOI] [PubMed] [Google Scholar]
- Lin, J., Lu, J., Shu, Z., Han, J., & Yu, N. (2024). Subject-specific modeling of EEG-fNIRS neurovascular coupling by task-related tensor decomposition. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 32, 452–461. 10.1109/TNSRE.2024.3355121 [DOI] [PubMed] [Google Scholar]
- Lin, J., Lu, J., Shu, Z., Yu, N., & Han, J. (2023). An EEG-fNIRS neurovascular coupling analysis method to investigate cognitive-motor interference. Computers in Biology and Medicine, 160, 106968. 10.1016/j.compbiomed.2023.106968 [DOI] [PubMed] [Google Scholar]
- Lin, X., Sai, L., & Yuan, Z. (2018). Detecting concealed information with fused electroencephalography and functional near-infrared spectroscopy. Neuroscience, 386, 284–294. 10.1016/j.neuroscience.2018.06.049 [DOI] [PubMed] [Google Scholar]
- Liu, J., & Calhoun, V. (2007). Parallel independent component analysis for multimodal analysis: Application to fMRI and EEG data. In 2007 4th IEEE International Symposium on Biomedical Imaging: From Nano to Macro (pp. 1028–1031). IEEE. 10.1109/ISBI.2007.357030 [DOI] [Google Scholar]
- Liu, Y., Ayaz, H., & Shewokis, P. A. (2017). Multisubject “learning” for mental workload classification using concurrent EEG, fNIRS, and physiological measures. Frontiers in Human Neuroscience, 11, 389. 10.3389/fnhum.2017.00389 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Logothetis, N. K. (2008). What we can do and what we cannot do with fMRI. Nature, 453(7197), 869–878. 10.1038/nature06976 [DOI] [PubMed] [Google Scholar]
- Lu, W., & Rajapakse, J. C. (2000). Constrained Independent Component Analysis. In NIPS’00: Proceedings of the 14th International Conference on Neural Information Processing Systems (pp. 549–555). https://dl.acm.org/doi/10.5555/3008751.3008831 [Google Scholar]
- Mackey, L. (2008). Deflation methods for sparse pca. In Koller D., Schuurmans D., Bengio Y., & Bottou L. (Eds.), Advances in neural information processing systems (Vol. 21). Curran Associates, Inc. https://proceedings.neurips.cc/paper_files/paper/2008/file/85d8ce590ad8981ca2c8286f79f59954-Paper.pdf [Google Scholar]
- Maddury, S. (2024). The performance of domain-based feature extraction on EEG, ECG, and fNIRS for Huntington’s disease diagnosis via shallow machine learning. Frontiers in Signal Processing, 4, 1321861. 10.3389/frsip.2024.1321861 [DOI] [Google Scholar]
- Mäkelä, N., Stenroos, M., Sarvas, J., & Ilmoniemi, R. J. (2018). Truncated rap-music (trap-music) for MEG and EEG source localization. NeuroImage, 167, 73–83. 10.1016/j.neuroimage.2017.11.013 [DOI] [PubMed] [Google Scholar]
- Mallat, S. (1999). A wavelet tour of signal processing. Academic Press. 10.1016/B978-0-12-374370-1.X0001-8 [DOI] [Google Scholar]
- Mandal, S., Singh, B., & Thakur, K. (2020). Classification of working memory loads using hybrid EEG and fNIRS in machine learning paradigm. Electronics Letters, 56(25), 1386–1389. 10.1049/el.2020.2710 [DOI] [Google Scholar]
- McLinden, J., Rahimi, N., Kumar, C., Krusienski, D. J., Shao, M., Spencer, K. M., & Shahriari, Y. (2024). Investigation of electro-vascular phase-amplitude coupling during an auditory task. Computers in Biology and Medicine, 169, 107902. 10.1016/j.compbiomed.2023.107902 [DOI] [PubMed] [Google Scholar]
- Meng, M., Dai, L., She, Q., Ma, Y., & Kong, W. (2021). Crossing time windows optimization based on mutual information for hybrid BCI. Mathematical Biosciences and Engineering, 18(6), 7919–7935. 10.3934/mbe.2021392 [DOI] [PubMed] [Google Scholar]
- Meng, Y., Su, Y., Li, D., Nan, J., & Xia, Y. (2025). Decoding of multi-modal signals for motor imagery based on window positioning. Signal, Image and Video Processing, 19(3), 242. 10.1007/s11760-025-03841-1 [DOI] [Google Scholar]
- Mesquita, R. C., Franceschini, M. A., & Boas, D. A. (2010). Resting state functional connectivity of the whole head with near-infrared spectroscopy. Biomedical Optics Express, 1(1), 324. 10.1364/BOE.1.000324 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mihara, M., & Miyai, I. (2016). Review of functional near-infrared spectroscopy in neurorehabilitation. Neurophotonics, 3(3), 031414. 10.1117/1.NPh.3.3.031414 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Molavi, B., & Dumont, G. A. (2012). Wavelet-based motion artifact removal for functional near-infrared spectroscopy. Physiological Measurement, 33(2), 259–270. 10.1088/0967-3334/33/2/259 [DOI] [PubMed] [Google Scholar]
- Morioka, H., Kanemura, A., Morimoto, S., Yoshioka, T., Oba, S., Kawanabe, M., & Ishii, S. (2014). Decoding spatial attention by using cortical currents estimated from electroencephalography with near-infrared spectroscopy prior information. NeuroImage, 90, 128–139. 10.1016/j.neuroimage.2013.12.035 [DOI] [PubMed] [Google Scholar]
- Mosher, J. C., & Leahy, R. M. (1999). Source localization using recursively applied and projected (RAP) MUSIC. IEEE Transactions on Signal Processing, 47(2), 332–340. 10.1109/78.740118 [DOI] [PubMed] [Google Scholar]
- Moslehi, A. H., & Davies, T. C. (2021). EEG electrode selection for a two-class motor imagery task in a BCI using fNIRS prior data. In 2021 43rd Annual International Conference of the IEEE Engineering in Medicine & Biology Society (EMBC) (pp. 6627–6630). IEEE. 10.1109/EMBC46164.2021.9630786 [DOI] [PubMed] [Google Scholar]
- Muellbacher, W., Facchini, S., Boroojerdi, B., & Hallett, M. (2000). Changes in motor cortex excitability during ipsilateral hand muscle activation in humans. Clinical Neurophysiology, 111(2), 344–349. 10.1016/S1388-2457(99)00243-6 [DOI] [PubMed] [Google Scholar]
- Muthalib, M., Anwar, A. R., Perrey, S., Dat, M., Galka, A., Wolff, S., Heute, U., Deuschl, G., Raethjen, J., & Muthuraman, M. (2013). Multimodal integration of fNIRS, fMRI and EEG neuroimaging. Clinical Neurophysiology, 124(10), 2060–2062. 10.1016/j.clinph.2013.03.018 [DOI] [PubMed] [Google Scholar]
- Nair, V. V., Kish, B. R., Yang, H.-C., Yu, Z., Guo, H., Tong, Y., & Liang, Z. (2021). Monitoring anesthesia using simultaneous functional Near Infrared Spectroscopy and Electroencephalography. Clinical Neurophysiology, 132(7), 1636–1646. 10.1016/j.clinph.2021.03.025 [DOI] [PubMed] [Google Scholar]
- Nguyen, T., Ahn, S., Jang, H., Jun, S. C., & Kim, J. G. (2017). Utilization of a combined EEG/NIRS system to predict driver drowsiness. Scientific Reports, 7, 43933. 10.1038/srep43933 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nguyen, T., Potter, T., Nguyen, T., Karmonik, C., Grossman, R., & Zhang, Y. (2016). EEG source imaging guided by spatiotemporal specific fMRI: Toward an understanding of dynamic cognitive processes. Neural Plasticity, 2016, 1–10. 10.1155/2016/4182483 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nia, A. F., Tang, V., Maso Talou, G. D., & Billinghurst, M. (2025). Decoding emotions through personalized multi-modal fNIRS-EEG Systems: Exploring deterministic fusion techniques. Biomedical Signal Processing and Control, 105, 107632. 10.1016/j.bspc.2025.107632 [DOI] [Google Scholar]
- Nikulin, V. V., Nolte, G., & Curio, G. (2011). A novel method for reliable and fast extraction of neuronal EEG/MEG oscillations on the basis of spatio-spectral decomposition. NeuroImage, 55(4), 1528–1535. 10.1016/j.neuroimage.2011.01.057 [DOI] [PubMed] [Google Scholar]
- Niso, G., Romero, E., Moreau, J. T., Araujo, A., & Krol, L. R. (2023). Wireless EEG: A survey of systems and studies. NeuroImage, 269, 119774. 10.1016/j.neuroimage.2022.119774 [DOI] [PubMed] [Google Scholar]
- Nourhashemi, M., Mahmoudzadeh, M., Goudjil, S., Kongolo, G., & Wallois, F. (2020). Neurovascular coupling in the developing neonatal brain at rest. Human Brain Mapping, 41(2), 503–519. 10.1002/hbm.24818 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nunez, P. L., & Srinivasan, R. (2006, January). Electric fields of the brain: The neurophysics of EEG. Oxford University Press. 10.1093/acprof:oso/9780195050387.001.0001 [DOI] [Google Scholar]
- Omurtag, A., Aghajani, H., & Keles, H. O. (2017). Decoding human mental states by whole-head EEG+fNIRS during category fluency task performance. Journal of Neural Engineering, 14(6), 066003. 10.1088/1741-2552/aa814b [DOI] [PubMed] [Google Scholar]
- Oostenveld, R., Fries, P., Maris, E., & Schoffelen, J.-M. (2011). Fieldtrip: Open source software for advanced analysis of MEG, EEG, and invasive electrophysiological data. Computational Intelligence and Neuroscience, 2011, 156869. 10.1155/2011/156869 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ortega, P., Zhao, T., & Faisal, A. A. (2020). HYGRIP: Full-stack characterization of neurobehavioral signals (fNIRS, EEG, EMG, Force, and Breathing) during a bimanual grip force control task. Frontiers in Neuroscience, 14, 919. 10.3389/fnins.2020.00919 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Othman, M. H., Bhattacharya, M., Møller, K., Kjeldsen, S., Grand, J., Kjaergaard, J., Dutta, A., & Kondziella, D. (2021). Resting-state NIRS-EEG in unresponsive patients with acute brain injury: A proof-of-concept study. Neurocritical Care, 34(1), 31–44. 10.1007/s12028-020-00971-x [DOI] [PubMed] [Google Scholar]
- Pang, J. C., Aquino, K. M., Oldehinkel, M., Robinson, P. A., Fulcher, B. D., Breakspear, M., & Fornito, A. (2023). Geometric constraints on human brain function. Nature, 618(7965), 566–574. 10.1038/s41586-023-06098-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Park, J. L., Dudchenko, P. A., & Donaldson, D. I. (2018). Navigation in real-world environments: New opportunities afforded by advances in mobile brain imaging. Frontiers in Human Neuroscience, 12, 361. 10.3389/fnhum.2018.00361 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parkhomenko, E., Tritchler, D., & Beyene, J. (2009). Sparse canonical correlation analysis with application to genomic data integration. Statistical Applications in Genetics and Molecular Biology, 8(1), 1–34. 10.2202/1544-6115.1406 [DOI] [PubMed] [Google Scholar]
- Parra, L. C., Spence, C. D., Gerson, A. D., & Sajda, P. (2005). Recipes for the linear analysis of EEG. NeuroImage, 28(2), 326–341. 10.1016/j.neuroimage.2005.05.032 [DOI] [PubMed] [Google Scholar]
- Peng, K., Nguyen, D. K., Tayah, T., Vannasing, P., Tremblay, J., Sawan, M., Lassonde, M., Lesage, F., & Pouliot, P. (2014). fNIRS-EEG study of focal interictal epileptiform discharges. Epilepsy Research, 108(3), 491–505. 10.1016/j.eplepsyres.2013.12.011 [DOI] [PubMed] [Google Scholar]
- Peng, K., Nguyen, D. K., Vannasing, P., Tremblay, J., Lesage, F., & Pouliot, P. (2016). Using patient-specific hemodynamic response function in epileptic spike analysis of human epilepsy: A study based on EEG-fNIRS. NeuroImage, 126, 239–255. 10.1016/j.neuroimage.2015.11.045 [DOI] [PubMed] [Google Scholar]
- Perpetuini, D., Chiarelli, A. M., Filippini, C., Cardone, D., Croce, P., Rotunno, L., Anzoletti, N., Zito, M., Zappasodi, F., & Merla, A. (2020). Working memory decline in Alzheimer’s disease is detected by complexity analysis of multimodal EEG-fNIRS. Entropy (Basel), 22(12), 1380. 10.3390/e22121380 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pfurtscheller, G. (2010). The hybrid BCI. Frontiers in Neuroscience. 10.3389/fnpro.2010.00003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pfurtscheller, G., Daly, I., Bauernfeind, G., & Müller-Putz, G. R. (2012). Coupling between intrinsic prefrontal HbO and central EEG beta power oscillations in the resting brain. PLoS One, 7(8), e43640. 10.1371/journal.pone.0043640 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pfurtscheller, G., & Lopes da Silva, F. H. (1999). Event-related eeg/meg synchronization and desynchronization: Basic principles. Clinical Neurophysiology, 110(11), 1842–1857. 10.1016/S1388-2457(99)00141-8 [DOI] [PubMed] [Google Scholar]
- Pinti, P., Aichelburg, C., Gilbert, S., Hamilton, A., Hirsch, J., Burgess, P., & Tachtsidis, I. (2018). A review on the use of wearable functional near-infrared spectroscopy in naturalistic environments: Review of fNIRS measurements in naturalistic environments. Japanese Psychological Research, 60(4), 347–373. 10.1111/jpr.12206 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pinti, P., Tachtsidis, I., Hamilton, A., Hirsch, J., Aichelburg, C., Gilbert, S., & Burgess, P. W. (2020). The present and future use of functional near-infrared spectroscopy (fNIRS) for cognitive neuroscience. Annals of the New York Academy of Sciences, 1464(1), 5–29. 10.1111/nyas.13948 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pollonini, L., Bortfeld, H., & Oghalai, J. S. (2016). Phoebe: A method for real time mapping of optodes-scalp coupling in functional near-infrared spectroscopy. Biomedical Optics Express, 7(12), 5104–5119. 10.1364/BOE.7.005104 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pouliot, P., Tran, T. P. Y., Birca, V., Vannasing, P., Tremblay, J., Lassonde, M., & Nguyen, D. K. (2014). Hemodynamic changes during posterior epilepsies: An EEG-fNIRS study. Epilepsy Research, 108(5), 883–890. 10.1016/j.eplepsyres.2014.03.007 [DOI] [PubMed] [Google Scholar]
- Putze, F., Hesslinger, S., Tse, C.-Y., Huang, Y., Herff, C., Guan, C., & Schultz, T. (2014). Hybrid fNIRS-EEG based classification of auditory and visual perception processes. Frontiers in Neuroscience, 8, 373. 10.3389/fnins.2014.00373 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qin, Y., Wu, J., Bulger, E., Cao, J., Dehghani, H., Shinn-Cunningham, B., & Kainerstorfer, J. M. (2024). Optimizing spatial accuracy in electroencephalography reconstruction through diffuse optical tomography priors in the auditory cortex. Biomedical Optics Express, 15(8), 4859–4876. 10.1364/BOE.531576 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qiu, L., Zhong, Y., He, Z., & Pan, J. (2022). Improved classification performance of EEG-fNIRS multimodal brain-computer interface based on multi-domain features and multi-level progressive learning. Frontiers in Human Neuroscience, 16, 973959. 10.3389/fnhum.2022.973959 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qu, J., Du, Y., Jing, J., Wang, J., Bu, L., & Wang, Y. (2025). Short-term longitudinal study on brain network informatics of stroke patients under acupuncture and motor imagery intervention. IEEE Journal of Biomedical and Health Informatics, 29(5), 3356–3365. 10.1109/JBHI.2025.3527074 [DOI] [PubMed] [Google Scholar]
- Ran, Y., Fan, Y., Wu, S., Chen, C., Li, Y., Gao, T., Zhang, H., Han, C., & Tang, X. (2025). TdCCA with dual-modal signal fusion: Degenerated occipital and frontal connectivity of adult moyamoya disease for early identification. Translational Stroke Research, 16(4), 1340–1357. 10.1007/s12975-024-01313-1 [DOI] [PubMed] [Google Scholar]
- Rezaee, Z., Ranjan, S., Solanki, D., Bhattacharya, M., Srivastava, M. V. P., Lahiri, U., & Dutta, A. (2021). Feasibility of combining functional near-infrared spectroscopy with electroencephalography to identify chronic stroke responders to cerebellar transcranial direct current stimulation-a computational modeling and portable neuroimaging methodological study. Cerebellum, 20(6), 853–871. 10.1007/s12311-021-01249-4 [DOI] [PubMed] [Google Scholar]
- Rezazadeh, A., Yousefi, R., Wong, A., Rudzicz, F., & Chau, T. (2019). Development of a ternary hybrid fNIRS-EEG brain–computer interface based on imagined speech. Brain-Computer Interfaces, 6, 1–13. 10.1080/2326263X.2019.1698928 [DOI] [Google Scholar]
- Rogers, D., O’Brien, W. J., Gao, Y., Zimmermann, B., Grover, S., Zhang, Y., Gaona, A. K., Duwadi, S., Anderson, J. E., Carlton, L., Rahimi, P., Farzam, P. Y., Von Lühmann, A., Reinhart, R. M. G., Boas, D. A., & Yücel, M. A. (2025). Co-localized optode-electrode design for multimodal functional near infrared spectroscopy and electroencephalography. Neurophotonics, 12(2), 025006. 10.1117/1.NPh.12.2.025006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Safaie, J., Grebe, R., Moghaddam, H. A., & Wallois, F. (2013). Toward a fully integrated wireless wearable EEG-NIRS bimodal acquisition system. Journal of Neural Engineering, 10(5), 056001. 10.1088/1741-2560/10/5/056001 [DOI] [PubMed] [Google Scholar]
- Saricaoglu, M., Yücel, M. A., Budak, M., Omurtag, A., & Hanoglu, L. (2025). Different cortex activation between young and middle-aged people during different type problem-solving: An EEG&fNIRS study. NeuroImage, 308, 121062. 10.1016/j.neuroimage.2025.121062 [DOI] [PubMed] [Google Scholar]
- Sato, M.-a., Yoshioka, T., Kajihara, S., Toyama, K., Goda, N., Doya, K., & Kawato, M. (2004). Hierarchical Bayesian estimation for MEG inverse problem. NeuroImage, 23(3), 806–826. 10.1016/j.neuroimage.2004.06.037 [DOI] [PubMed] [Google Scholar]
- Schmuckler, M. A. (2001). What is ecological validity? A dimensional analysis. Infancy, 2(4), 419–436. 10.1207/S15327078IN0204_02 [DOI] [PubMed] [Google Scholar]
- Schneider, S., Wagels, L., Haeussinger, F. B., Fallgatter, A. J., Ehlis, A.-C., & Rapp, A. M. (2015). Haemodynamic and electrophysiological markers of pragmatic language comprehension in schizophrenia. The World Journal of Biological Psychiatry, 16(6), 398–410. 10.3109/15622975.2015.1019359 [DOI] [PubMed] [Google Scholar]
- Scholkmann, F., Spichtig, S., Muehlemann, T., & Wolf, M. (2010). How to detect and reduce movement artifacts in near-infrared imaging using moving standard deviation and spline interpolation. Physiological Measurement, 31(5), 649–662. 10.1088/0967-3334/31/5/004 [DOI] [PubMed] [Google Scholar]
- Scholkmann, F., Tachtsidis, I., Wolf, M., & Wolf, U. (2022). Systemic physiology augmented functional near-infrared spectroscopy: A powerful approach to study the embodied human brain. Neurophotonics, 9(3), 030801. 10.1117/1.NPh.9.3.030801 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shin, J., Kim, D.-W., Müller, K.-R., & Hwang, H.-J. (2018). Improvement of information transfer rates using a hybrid EEG-NIRS brain-computer interface with a short trial length: Offline and pseudo-online analyses. Sensors (Basel), 18(6), 1827. 10.3390/s18061827 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shin, J., Kwon, J., & Im, C.-H. (2018). A ternary hybrid EEG-NIRS brain-computer interface for the classification of brain activation patterns during mental arithmetic, motor imagery, and idle state. Frontiers in Neuroinformatics, 12, 5. 10.3389/fninf.2018.00005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shin, J., Müller, K.-R., & Hwang, H.-J. (2018). Eyes-closed hybrid brain-computer interface employing frontal brain activation. PLoS One, 13(5), e0196359. 10.1371/journal.pone.0196359 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shin, J., Von Luhmann, A., Blankertz, B., Kim, D.-W., Jeong, J., Hwang, H.-J., & Muller, K.-R. (2017). Open access dataset for EEG+NIRS single-trial classification. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 25(10), 1735–1745. 10.1109/TNSRE.2016.2628057 [DOI] [PubMed] [Google Scholar]
- Shin, J., Von Lühmann, A., Kim, D.-W., Mehnert, J., Hwang, H.-J., & Müller, K.-R. (2018). Simultaneous acquisition of EEG and NIRS during cognitive tasks for an open access dataset. Scientific Data, 5(1), 180003. 10.1038/sdata.2018.3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Si, X., Huang, H., Yu, J., & Ming, D. (2024). EEG microstates and fNIRS metrics reveal the spatiotemporal joint neural processing features of human emotions. IEEE Transactions on Affective Computing, 15(4), 2128–2138. 10.1109/TAFFC.2024.3399729 [DOI] [Google Scholar]
- Sirpal, P., Kassab, A., Pouliot, P., Nguyen, D. K., & Lesage, F. (2019). fNIRS improves seizure detection in multimodal EEG-fNIRS recordings. Journal of Biomedical Optics, 24(5), 1–9. 10.1117/1.JBO.24.5.051408 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sonkusare, S., Breakspear, M., & Guo, C. (2019). Naturalistic stimuli in neuroscience: Critically acclaimed. Trends in Cognitive Sciences, 23(8), 699–714. 10.1016/j.tics.2019.05.004 [DOI] [PubMed] [Google Scholar]
- Sood, M., Besson, P., Muthalib, M., Jindal, U., Perrey, S., Dutta, A., & Hayashibe, M. (2016). NIRS-EEG joint imaging during transcranial direct current stimulation: Online parameter estimation with an autoregressive model. Journal of Neuroscience Methods, 274, 71–80. 10.1016/j.jneumeth.2016.09.008 [DOI] [PubMed] [Google Scholar]
- Su, W.-C., Dashtestani, H., Miguel, H. O., Condy, E., Buckley, A., Park, S., Perreault, J. B., Nguyen, T., Zeytinoglu, S., Millerhagen, J., Fox, N., & Gandjbakhche, A. (2023). Simultaneous multimodal fNIRS-EEG recordings reveal new insights in neural activity during motor execution, observation, and imagery. Scientific Reports, 13(1), 5151. 10.1038/s41598-023-31609-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun, L., Ji, S., Yu, S., & Ye, J. (2009). On the equivalence between canonical correlation analysis and orthonormalized partial least squares. In Proceedings of the 21st International Joint Conference on Artificial Intelligence (pp. 1230–1235). ACM. 10.1145/1390156.1390285 [DOI] [Google Scholar]
- Sun, Y., Ayaz, H., & Akansu, A. N. (2020). Multimodal affective state assessment using fNIRS + EEG and spontaneous facial expression. Brain Sciences, 10(2), 85. 10.3390/brainsci10020085 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Suzuki, K., Okumura, Y., Kita, Y., Oi, Y., Shinoda, H., & Inagaki, M. (2018). The relationship between the superior frontal cortex and alpha oscillation in a flanker task: Simultaneous recording of electroencephalogram (EEG) and near infrared spectroscopy (NIRS). Neuroscience Research, 131, 30–35. 10.1016/j.neures.2017.08.011 [DOI] [PubMed] [Google Scholar]
- Tachtsidis, I., & Scholkmann, F. (2016). False positives and false negatives in functional near-infrared spectroscopy: Issues, challenges, and the way forward. Neurophotonics, 3(3), 031405. 10.1117/1.NPh.3.3.031405 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Talukdar, M. T., Frost, H. R., & Diamond, S. G. (2015). Modeling neurovascular coupling from clustered parameter sets for multimodal EEG-NIRS. Computational and Mathematical Methods in Medicine, 2015, 830849. 10.1155/2015/830849 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Talukdar, T., Moore, J. H., & Diamond, S. G. (2013). Continuous correction of differential path length factor in near-infrared spectroscopy. Journal of Biomedical Optics, 18(5), 056001. 10.1117/1.JBO.18.5.056001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tang, J., LeBel, A., Jain, S., & Huth, A. G. (2023). Semantic reconstruction of continuous language from non-invasive brain recordings. Nature Neuroscience, 26(5), 858–866. 10.1038/s41593-023-01304-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tangermann, M., Müller, K.-R., Aertsen, A., Birbaumer, N., Braun, C., Brunner, C., Leeb, R., Mehring, C., Miller, K. J., Müller-Putz, G. R., Nolte, G., Pfurtscheller, G., Preissl, H., Schalk, G., Schlögl, A., Vidaurre, C., Waldert, S., & Blankertz, B. (2012). Review of the BCI Competition IV. Frontiers in Neuroscience, 6, 55. 10.3389/fnins.2012.00055 [DOI] [PMC free article] [PubMed] [Google Scholar]
- The CGAL Project. (2024). CGAL user and reference manual (6.0.1). CGAL Editorial Board. https://www.cgal.org/
- Thompson, G. J., Pan, W.-J., Magnuson, M. E., Jaeger, D., & Keilholz, S. D. (2014). Quasi-periodic patterns (QPP): Large-scale dynamics in resting state fmri that correlate with local infraslow electrical activity. NeuroImage, 84, 1018–1031. 10.1016/j.neuroimage.2013.09.029 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tomita, Y., Vialatte, F.-B., Dreyfus, G., Mitsukura, Y., Bakardjian, H., & Cichocki, A. (2014). Bimodal BCI using simultaneously NIRS and EEG. IEEE Transactions on Biomedical Engineering, 61(4), 1274–1284. 10.1109/TBME.2014.2300492 [DOI] [PubMed] [Google Scholar]
- Uchitel, J., Vidal-Rosas, E. E., Cooper, R. J., & Zhao, H. (2021). Wearable, integrated EEG–fNIRS technologies: A review. Sensors, 21(18), 6106. 10.3390/s21186106 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Urbanowicz, R. J., Meeker, M., La Cava, W., Olson, R. S., & Moore, J. H. (2018). Relief-based feature selection: Introduction and review. Journal of Biomedical Informatics, 85, 189–203. 10.1016/j.jbi.2018.07.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Urigüen, J. A., & Garcia-Zapirain, B. (2015). EEG artifact removal—State-of-the-art and guidelines. Journal of Neural Engineering, 12(3), 031001. 10.1088/1741-2560/12/3/031001 [DOI] [PubMed] [Google Scholar]
- Vakitbilir, N., Islam, A., Gomez, A., Stein, K. Y., Froese, L., Bergmann, T., Sainbhi, A. S., McClarty, D., Raj, R., & Zeiler, F. A. (2024). Multivariate modelling and prediction of high-frequency sensor-based cerebral physiologic signals: Narrative review of machine learning methodologies. Sensors, 24(24), 8148. 10.3390/s24248148 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vidal-Rosas, E. E., Von Lühmann, A., Pinti, P., & Cooper, R. J. (2023). Wearable, high-density fNIRS and diffuse optical tomography technologies: A perspective. Neurophotonics, 10(2), 023513. 10.1117/1.NPh.10.2.023513 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vinck, M., Oostenveld, R., Van Wingerden, M., Battaglia, F., & Pennartz, C. M. (2011). An improved index of phase-synchronization for electrophysiological data in the presence of volume-conduction, noise and sample-size bias. NeuroImage, 55(4), 1548–1565. 10.1016/j.neuroimage.2011.01.055 [DOI] [PubMed] [Google Scholar]
- von Lühmann, A. (2018). Multimodal Instrumentation and Methods for Neurotechnology Out of the Lab [Doctoral dissertation, Technische Universität Berlin] [10.14279/depositonce-7445]. https://depositonce.tu-berlin.de/items/da8ba104-5dfd-4875-a0a2-871a5890e288
- von Lühmann, A., Kura, S., O’Brien, W. J., Zimmermann, B. B., Duwadi, S., Rogers, D., Anderson, J. E., Farzam, P., Snow, C., Chen, A., Yücel, M. A., Perkins, N., & Boas, D. A. (2024). Ninjacap: A fully customizable and 3D printable headgear for fNIRS and EEG brain imaging. bioRxiv. 10.1101/2024.05.14.594159 [DOI] [PMC free article] [PubMed] [Google Scholar]
- von Lühmann, A., Li, X., Gilmore, N., Boas, D. A., & Yücel, M. A. (2020). Open access multimodal fnirs resting state dataset with and without synthetic hemodynamic responses. Frontiers in Neuroscience, 14, 579353. 10.3389/fnins.2020.579353 [DOI] [PMC free article] [PubMed] [Google Scholar]
- von Lühmann, A., Li, X., Müller, K.-R., Boas, D. A., & Yücel, M. A. (2020). Improved physiological noise regression in fnirs: A multimodal extension of the general linear model using temporally embedded canonical correlation analysis. NeuroImage, 208, 116472. 10.1016/j.neuroimage.2019.116472 [DOI] [PMC free article] [PubMed] [Google Scholar]
- von Lühmann, A., Wabnitz, H., Sander, T., & Müller, K.-R. (2017). M3ba: A mobile, modular, multimodal biosignal acquisition architecture for miniaturized EEG-NIRS-based hybrid BCI and monitoring. IEEE Transactions on Biomedical Engineering, 64(6), 1199–1210. 10.1109/TBME.2016.2594127 [DOI] [PubMed] [Google Scholar]
- von Lühmann, A., Zheng, Y., Ortega-Martinez, A., Kiran, S., Somers, D. C., Cronin-Golomb, A., Awad, L. N., Ellis, T. D., Boas, D. A., & Yücel, M. A. (2021). Toward Neuroscience of the Everyday World (NEW) using functional near-infrared spectroscopy. Current Opinion in Biomedical Engineering, 18, 100272. 10.1016/j.cobme.2021.100272 [DOI] [PMC free article] [PubMed] [Google Scholar]
- von Lühmann, A., Boukouvalas, Z., Müller, K.-R., & Adalı, T. (2019). A new blind source separation framework for signal analysis and artifact rejection in functional near-infrared spectroscopy. NeuroImage, 200, 72–88. 10.1016/j.neuroimage.2019.06.021 [DOI] [PubMed] [Google Scholar]
- Vorwerk, J., Oostenveld, R., Piastra, M. C., Magyari, L., & Wolters, C. H. (2018). The fieldtrip–simbio pipeline for EEG forward solutions. Biomedical Engineering Online, 17, 37. 10.1186/s12938-018-0463-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- Walia, P., Fu, Y., Norfleet, J., Schwaitzberg, S. D., Intes, X., De, S., Cavuoto, L., & Dutta, A. (2022). Error-related brain state analysis using electroencephalography in conjunction with functional near-infrared spectroscopy during a complex surgical motor task. Brain Informatics, 9(1), 29. 10.1186/s40708-022-00179-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wallois, F., Patil, A., Héberlé, C., & Grebe, R. (2010). EEG-NIRS in epilepsy in children and neonates. Neurophysiologie Clinique/Clinical Neurophysiology, 40(5–6), 281–292. 10.1016/j.neucli.2010.08.004 [DOI] [PubMed] [Google Scholar]
- Wan, L., Li, Y., Zhu, G., Yang, D., Li, F., Wang, W., Chen, J., Yang, G., & Li, R. (2024). Multimodal investigation of dynamic brain network alterations in autism spectrum disorder: Linking connectivity dynamics to symptoms and developmental trajectories. NeuroImage, 302, 120895. 10.1016/j.neuroimage.2024.120895 [DOI] [PubMed] [Google Scholar]
- Wang, Y., Gao, S., & Gao, X. (2005). Common spatial pattern method for channel selelction in motor imagery based brain-computer interface. In 2005 IEEE Engineering in Medicine and Biology 27th Annual Conference (pp. 5392–5395). IEEE. 10.1109/IEMBS.2005.1615701 [DOI] [PubMed] [Google Scholar]
- Wang, Y., Yang, Z., Ji, H., Li, J., Liu, L., & Zhuang, J. (2022). Cross-modal transfer learning from EEG to functional near-infrared spectroscopy for classification task in brain-computer interface system. Frontiers in Psychology, 13, 833007. 10.3389/fpsyg.2022.833007 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang, Z., Yang, L., Zhou, Y., Chen, L., Gu, B., Liu, S., Xu, M., He, F., & Ming, D. (2023). Incorporating EEG and fNIRS patterns to evaluate cortical excitability and MI-BCI performance during motor training. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 31, 2872–2882. 10.1109/TNSRE.2023.3281855 [DOI] [PubMed] [Google Scholar]
- Watson, B. O. (2018). Cognitive and physiologic impacts of the infraslow oscillation. Frontiers in Systems Neuroscience, 12, 44. 10.3389/fnsys.2018.00044 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wellcome Centre for Human Neuroimaging, UCL. (2014). Spm12 manual. Statistical Parametric Mapping (SPM) Toolbox. https://www.fil.ion.ucl.ac.uk/spm/doc/spm12_manual.pdf
- Wheelock, M. D., Culver, J. P., & Eggebrecht, A. T. (2019). High-density diffuse optical tomography for imaging human brain function. Review of Scientific Instruments, 90(5), 051101. 10.1063/1.5086809 [DOI] [PMC free article] [PubMed] [Google Scholar]
- White, B. R., & Culver, J. P. (2010). Quantitative evaluation of high-density diffuse optical tomography: In vivo resolution and mapping performance. Journal of Biomedical Optics, 15(2), 1. 10.1117/1.3368999 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Witten, D. M., Tibshirani, R., & Hastie, T. (2009). A penalized matrix decomposition, with applications to sparse principal components and canonical correlation analysis. Biostatistics, 10(3), 515–534. 10.1093/biostatistics/kxp008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Witten, D. M., & Tibshirani, R. J. (2009). Extensions of sparse canonical correlation analysis with applications to genomic data. Statistical Applications in Genetics and Molecular Biology, 8(1), Article 28. 10.2202/1544-6115.1470 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wold, S., Ruhe, A., Wold, H., & Dunn, W. J., III. (1984). The collinearity problem in linear regression. The partial least squares (PLS) approach to generalized inverses. SIAM Journal on Scientific and Statistical Computing, 5(3), 735–743. 10.1137/0905052 [DOI] [Google Scholar]
- Worsley, K. J., Poline, J. B., Friston, K. J., & Evans, A. C. (1997). Characterizing the response of PET and fMRI data using multivariate linear models. NeuroImage, 6(4), 305–319. 10.1006/nimg.1997.0294 [DOI] [PubMed] [Google Scholar]
- Xu, T., Zhou, Z., Yang, Y., Li, Y., Li, J., Bezerianos, A., & Wang, H. (2023). Motor imagery decoding enhancement based on hybrid eeg-fnirs signals. IEEE Access, 11, 65277–65288. 10.1109/ACCESS.2023.3289709 [DOI] [Google Scholar]
- Yang, H., Ghayem, F., Gabrielson, B., Akhonda, M. A. B. S., Calhoun, V. D., & Adali, T. (2023). Constrained independent component analysis based on entropy bound minimization for subgroup identification from multi-subject fMRI data. In ICASSP 2023—2023 IEEE International Conference on Acoustics, Speech and Signal Processing (ICASSP) (pp. 1–5). IEEE. 10.1109/ICASSP49357.2023.10095816 [DOI] [Google Scholar]
- Yeung, M. K., & Chu, V. W. (2022). Viewing neurovascular coupling through the lens of combined EEG–fNIRS: A systematic review of current methods. Psychophysiology, 59(6), e14054. 10.1111/psyp.14054 [DOI] [PubMed] [Google Scholar]
- Yi, L., Xie, G., Li, Z., Li, X., Zhang, Y., Wu, K., Shao, G., Lv, B., Jing, H., Zhang, C., Liang, W., Sun, J., Hao, Z., & Liang, J. (2023). Automatic depression diagnosis through hybrid EEG and near-infrared spectroscopy features using support vector machine. Frontiers in Neuroscience, 17, 1205931. 10.3389/fnins.2023.1205931 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yin, X., Xu, B., Jiang, C., Fu, Y., Wang, Z., Li, H., & Shi, G. (2015). A hybrid BCI based on EEG and fNIRS signals improves the performance of decoding motor imagery of both force and speed of hand clenching. Journal of Neural Engineering, 12(3), 036004. 10.1088/1741-2560/12/3/036004 [DOI] [PubMed] [Google Scholar]
- Yücel, M. A., Lühmann, A. V., Scholkmann, F., Gervain, J., Dan, I., Ayaz, H., Boas, D., Cooper, R. J., Culver, J., Elwell, C. E., Eggebrecht, A., Franceschini, M. A., Grova, C., Homae, F., Lesage, F., Obrig, H., Tachtsidis, I., Tak, S., Tong, Y., … Wolf, M. (2021). Best practices for fNIRS publications. Neurophotonics, 8(1), 012101. 10.1117/1.NPh.8.1.012101 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yücel, M. A., Luke, R., Mesquita, R. C., von Lühmann, A., Mehler, D. M. A., Lührs, M., Gemignani, J., Abdalmalak, A., Albrecht, F., & Almeida, I. (2024, September). The fNIRS reproducibility study hub (FRESH): Exploring variability and enhancing transparency in FNIRS neuroimaging research. ArXiv. 10.31222/osf.io/pc6x8 [DOI]
- Yücel, M. A., Selb, J., Aasted, C. M., Petkov, M. P., Becerra, L., Borsook, D., & Boas, D. A. (2015). Short separation regression improves statistical significance and better localizes the hemodynamic response obtained by near-infrared spectroscopy for tasks with differing autonomic responses. Neurophotonics, 2(3), 035005. 10.1117/1.NPh.2.3.035005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zakeri, Z., Arif, A., Omurtag, A., Breedon, P., & Khalid, A. (2023). Multimodal assessment of cognitive workload using neural, subjective and behavioural measures in smart factory settings. Sensors (Basel), 23(21), 8926. 10.3390/s23218926 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zama, T., Takahashi, Y., & Shimada, S. (2019). Simultaneous EEG-NIRS measurement of the inferior parietal lobule during a reaching task with delayed visual feedback. Frontiers in Human Neuroscience, 13, 301. 10.3389/fnhum.2019.00301 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zeff, B. W., White, B. R., Dehghani, H., Schlaggar, B. L., & Culver, J. P. (2007). Retinotopic mapping of adult human visual cortex with high-density diffuse optical tomography. Proceedings of the National Academy of Sciences of the United States of America, 104(29), 12169–12174. 10.1073/pnas.0611266104 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zennifa, F., Ageno, S., Hatano, S., & Iramina, K. (2018). Hybrid system for engagement recognition during cognitive tasks using a CFS + KNN algorithm. Sensors (Basel), 18(11), 3691. 10.3390/s18113691 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang, X., Noah, J. A., & Hirsch, J. (2016). Separation of the global and local components in functional near-infrared spectroscopy signals using principal component spatial filtering. Neurophotonics, 3(1), 015004. 10.1117/1.NPh.3.1.015004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang, Y., & Zhu, C. (2019). Assessing brain networks by resting-state dynamic functional connectivity: An fNIRS-EEG study. Frontiers in Neuroscience, 13, 1430. 10.3389/fnins.2019.01430 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao, H., & Cooper, R. J. (2017). Review of recent progress toward a fiberless, whole-scalp diffuse optical tomography system. Neurophotonics, 5(1), 1. 10.1117/1.NPh.5.1.011012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhu, R., She, Q., Li, R., Tan, T., & Zhang, Y. (2024). Neurovascular coupling analysis based on multivariate variational Gaussian process convergent cross-mapping. IEEE Transactions on Neural Systems and Rehabilitation Engineering, 32, 1873–1883. 10.1109/TNSRE.2024.3398662 [DOI] [PubMed] [Google Scholar]
- Zhuang, X., Yang, Z., & Cordes, D. (2020). A technical review of canonical correlation analysis for neuroscience applications. Human Brain Mapping, 41(13), 3807–3833. 10.1002/hbm.25090 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zich, C., Debener, S., Thoene, A.-K., Chen, L.-C., & Kranczioch, C. (2017). Simultaneous EEG-fNIRS reveals how age and feedback affect motor imagery signatures. Neurobiology of Aging, 49, 183–197. 10.1016/j.neurobiolaging.2016.10.011 [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
All compared source-decomposition methods can be found in the “sigdecomp” package10 of our Python-based fNIRS/DOT analysis toolbox Cedalion (IBS-Lab & Cedalion-Developers, 2024). We also provide an example notebook containing the methods comparison on our synthetic data to enable the reproducibility of our results.11









