Skip to main content

Specific Issues Related to EEG-fMRI at B 0 > 3 T

  • Chapter
  • First Online:
EEG - fMRI

Abstract

Functional MRI (fMRI) can be used to map regional changes in cerebral blood flow and the level of haemoglobin oxygenation (BOLD) associated with neuronal activity (Belliveau et al. 1991; Kwong et al. 1992; Ogawa et al. 1992). In 2003 the US Food and Drug Administration raised the value of the static field of “no significant risk” for MRI to 8 Tesla (T), potentially opening up this technology to large numbers of laboratories in the USA. Regulatory agencies in Europe and Asia reached similar conclusions, and as a result the number of high-field systems worldwide is growing rapidly. The increased static field B 0 allows the potential for improved signal-to-noise ratio (SNR), offering the possibility of increasing spatial resolution and reducing scan times (Wiesinger et al. 2006; Harel et al. 2006). In this chapter, we outline the safety issues raised and the challenges involved in performing EEG at high-field MRI or at static magnetic fields greater than 3 T.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 89.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 119.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  • Angelone LM, and G. Bonmassar. Use of resistances and resistive leads: implications on computed electric field and SAR values. in ISMRM Twelfth Scientific Meeting. 2004

    Google Scholar 

  • Angelone LM, et al. (2004) Metallic electrodes and leads in simultaneous EEG-MRI: specific absorption rate (SAR) simulation studies. Bioelectromagnetics 25(4):285–295

    Article  PubMed  Google Scholar 

  • Angelone LM, et al, (2005) New high resolution head model for accurate electromagnetic field computation. Proc 13th ISMRM Sci Meet, Miami, FL, USA, 7–13 May 2005, p 881

    Google Scholar 

  • Angelone LM, et al. (2006) On the effect of resistive EEG electrodes and leads during 7 Tesla MRI: simulation and temperature measurement studies. Magn Reson Imaging 24(6):801–812

    Article  PubMed  Google Scholar 

  • Armenean C, et al. (2004) RF-induced temperature elevation along metallic wires in clinical magnetic resonance imaging: influence of diameter and length. Magn Reson Med 52(5):1200–1206

    Article  PubMed  Google Scholar 

  • Baker KB, et al. (2004) Evaluation of specific absorption rate as a dosimeter of MRI-related implant heating. J Magn Reson Imaging 20(2):315–320

    Article  PubMed  Google Scholar 

  • Belliveau JW, et al. (1991) Functional mapping of the human visual cortex by magnetic resonance imaging. Science 254:716–719

    Article  PubMed  CAS  Google Scholar 

  • Berenger JP (1994) A perfectly matched layer for the absorption of electromagnetic waves. Comput Phys 114:185–200

    Article  Google Scholar 

  • Black PM, et al. (1999) Craniotomy for tumor treatment in an intraoperative magnetic resonance imaging unit. Neurosurgery 45(3):423–431; discussion 431–433

    Article  PubMed  CAS  Google Scholar 

  • Bottomley PA, Andrew ER (1978) RF magnetic field penetration, phase shift and power dissipation in biological tissue: Implications for NMR imaging. Phys Med Biol 23:630–643

    Article  PubMed  CAS  Google Scholar 

  • Bottomley P, Redington R, Edelstein W (1985) Estimating radiofrequency power deposition in body NMR imaging. Magn Reson Med 2:336–349

    Article  PubMed  CAS  Google Scholar 

  • Boundy KL, et al. (1996) Localization of temporal lobe epileptic foci with iodine-123 iododexetimide cholinergic neuroreceptor single-photon emission computed tomography. Neurology 47(4):1015–1020

    Article  PubMed  CAS  Google Scholar 

  • Carmichael DW, et al. Functional MRI with active, fully implanted, deep brain stimulation systems: safety and experimental confounds. Neuroimage, 2007. 37(2): p. 508–517

    Article  PubMed  Google Scholar 

  • Carmichael DW, et al. Safety of localizing epilepsy monitoring intracranial electroencephalograph electrodes using MRI: radiofrequency-induced heating. J Magn Reson Imaging, 2008. 28(5): p. 1233–1244

    Article  PubMed  Google Scholar 

  • Cangellaris AC, Wright SM (1991) Analysis of the numerical error caused by the stair-stepped approximation of a conducting boundary in FDTD simulations of electromagnetic phenomena. IEEE Trans Antennas Propagat 39:1518–1525

    Article  Google Scholar 

  • Chou CK, Guy AW (1979) Carbon-loaded teflon electrodes for chronic EEG recordings in microwave research. J Microw Power 14(4):399–404

    PubMed  CAS  Google Scholar 

  • Chou CK, McDougall JA, Chan KW (1997) RF heating of implanted spinal fusion stimulator during magnetic resonance imaging. IEEE Trans Biomed Eng 44(5):367–373

    Article  PubMed  CAS  Google Scholar 

  • Collins CM, Smith MB (2003) Spatial resolution of numerical models of man and calculated specific absorption rate using the FDTD method: a study at 64 MHz in a magnetic resonance imaging coil. J Magn Reson Imaging 18(3):383–388

    Article  PubMed  Google Scholar 

  • Cone-Wesson B, et al. (2002) The auditory steady-state response: comparisons with the auditory brainstem response. J Am Acad Audiol 13(4):173–187; quiz 225–226

    PubMed  Google Scholar 

  • Debener S, et al. (2008) Properties of the ballistocardiogram artifact as revealed by EEG recordings at 1.5, 3 and 7 Tesla. Int J Psychophysiol 67:189–199

    Article  PubMed  Google Scholar 

  • Dimbylow PJ, Gandhi OP (1991) Finite-difference time-domain calculations of SAR in a realistic heterogeneous model of the head for plane-wave exposure from 600 MHz to 3 GHz. Phys Med Biol 36(8):1075–1089

    Article  PubMed  CAS  Google Scholar 

  • FDA (2003) Criteria for significant risk investigations of magnetic resonance diagnostic devices. Center for Devices and Radiological Health, US FDA, Silver Spring

    Google Scholar 

  • Filipek PA, et al. (1994) The young adult human brain: an MRI-based morphometric analysis. Cereb Cortex 4(4): 344–360

    Article  PubMed  CAS  Google Scholar 

  • Gabriel C, Gabriel S, Corthout E (1996a) The dielectric properties of biological tissues: I. Literature survey. Phys Med Biol 41:2231–2249

    Article  PubMed  CAS  Google Scholar 

  • Gabriel C, Gabriel S, Corthout E (1996b) The dielectric properties of biological tissues: II. Measurements in the frequency range 10 Hz to 20 GHz. Phys Med Biol 41:2251–2269

    Article  PubMed  CAS  Google Scholar 

  • Galambos R, Makeig S, Talmachoff P (1981) A 40-Hz auditory potential recorded from the human scalp. Proc Natl Acad Sci USA 78:2643–2647

    Article  PubMed  CAS  Google Scholar 

  • Gandhi OP, Chen XB (1999) Specific absorption rates and induced current densities for an anatomy-based model of the human for exposure to time-varying magnetic fields of MRI. Magn Reson Med 41(4):816–823

    Article  PubMed  CAS  Google Scholar 

  • Gangarosa RE, et al. (1987) Operational safety issues in MRI. Magn Reson Imaging 5(4):287–292

    Article  PubMed  CAS  Google Scholar 

  • Glover GH, et al. (1985) Comparison of linear and circular polarization for magnetic resonance imaging. J Magn Reson 64:255–270

    CAS  Google Scholar 

  • Harel N, et al. (2006) Frontiers of brain mapping using MRI. J Magn Reson Imaging 23(6):945–957

    Article  PubMed  Google Scholar 

  • Haykin S (1996) Adaptive filter theory, 3rd edn. Prentice Hall, Upper Saddle River

    Google Scholar 

  • Haykin S (2002) Adaptive filter theory, 4th edn. Prentice Hall, Upper Saddle River

    Google Scholar 

  • Ho HS (2001) Safety of metallic implants in magnetic resonance imaging. J Magn Reson Imaging 14(4):472–477

    Article  PubMed  CAS  Google Scholar 

  • Hoult DI, Foreman D, Kolansky G, Kripiakevich D (2008) Overcoming high-field RF problems with non-magnetic cartesian feedback transceivers. MAGMA 21(1–2):15–29

    Article  PubMed  CAS  Google Scholar 

  • Huang-Hellinger FR, et al. (1995) Simultaneous functional magnetic resonance imaging and electrophysiological recording. Hum Brain Mapp 3:13–23

    Article  Google Scholar 

  • IEC (2002) International standard, medical equipment, part 2–33: particular requirements for the safety of the magnetic resonance equipment for medical diagnosis, 2nd revision. International Electrotechnical Commission 601-2-33: Geneva, pp 29–31

    Google Scholar 

  • Ives JR, et al. (1993) Monitoring the patient’s EEG during echo-planar MRI. Electroenceph Clin Neurophysiol 87:417–420

    Article  PubMed  CAS  Google Scholar 

  • Jin J-M (1999) Electromagnetic analysis and design in magnetic resonance imaging (Biomedical Engineering Series). CRC Press, Boca Raton

    Google Scholar 

  • Kainz W, et al. (2005) Dosimetric comparison of the specific anthropomorphic mannequin (SAM) to 14 anatomical head models using a novel definition for the mobile phone positioning. Phys Med Biol 50(14):3423–3445

    Article  PubMed  Google Scholar 

  • Kangarlu A, Shellock FG, Chakeres DW (2003) 8.0-Tesla human MR system: temperature changes associated with radiofrequency-induced heating of a head phantom. J Magn Reson Imaging 17(2):220–226

    Article  PubMed  Google Scholar 

  • Keltner JR, et al. (1991) Electromagnetic fields of surface coil in vivo NMR at high frequencies. Magn Reson Med 22:467–480

    Article  PubMed  CAS  Google Scholar 

  • Konings MK, Bartels LW, Smits HFM, Bakker CJG (2000) Heating around intravascular guidewires by resonating RF waves. JMRI 12:79–85

    Article  PubMed  CAS  Google Scholar 

  • Kwong KK, et al. (1992) Functional MR imaging of primary visual and motor cortex. In: 10th Annu Meet Soc Magnetic Resonance Imaging, New York, 25–29 April 1992

    Google Scholar 

  • Lazeyras F, et al. (2001) Functional MRI with simultaneous EEG recording: Feasibility and application to motor and visual activation. J Magn Reson Imaging 13(6):943–948

    Article  PubMed  CAS  Google Scholar 

  • Lemieux L, et al. (1997) Recording of EEG during fMRI experiments: patient safety. Magn Reson Med 38(6):943–952

    Article  PubMed  CAS  Google Scholar 

  • Lu, H, et al. Routine clinical brain MRI sequences for use at 3.0 Tesla. J Magn Reson Imaging, 2005. 22(1): p. 13–22

    Article  PubMed  Google Scholar 

  • Lu HZ, et al. (2005) Routine clinical brain MRI sequences for use at 3.0 Tesla. J Magn Reson Imaging 22:13–22

    Article  PubMed  Google Scholar 

  • Mansfield P, Morris PG (1982) NMR imaging in biomedicine (Advances in Magnetic Resonance, Suppl 2). Academic, New York

    Google Scholar 

  • Mirsattari SM, et al. (2004) MRI compatible EEG electrode system for routine use in the epilepsy monitoring unit and intensive care unit. Clin Neurophysiol 115(9):2175–2180

    Article  PubMed  Google Scholar 

  • Mullinger K (2008) Exploring the feasibility of simultaneous EEG/fMRI at 7T. Magn Reson Imaging 26(7):607–616

    Article  Google Scholar 

  • Mullinger KJ, et al. (2008) Effects of simultaneous EEG recording on MRI data quality at 1.5, 3 and 7 Tesla. Int J Psychophysiol 67:178–188

    Article  PubMed  Google Scholar 

  • Nakamura W, et al. (2006) Removal of ballistocardiogram artifacts from simultaneously recorded EEG and fMRI data using independent component analysis. IEEE Trans Biomed Eng 53(7):1294–1308

    Article  PubMed  Google Scholar 

  • Nitz WR, et al. (2005) Specific absorption rate as a poor indicator of magnetic resonance-related implant heating. Invest Radiol 40(12):773–776

    Article  PubMed  Google Scholar 

  • Ogawa, S., et al. (1992) Intrinsic signal changes accompanying sensory stimulation: Functional brain mapping with magnetic resonance imaging. Proc Natl Acad Sci USA 89:5951–5955

    Article  PubMed  CAS  Google Scholar 

  • Percival DB, Walden AT (1993) Spectral analysis for physical applications: multitaper and conventional univariate techniques. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Purdon PL, et al. (2005) Concurrent EEG/fMRI of temporal fluctuations in the 40-Hz auditory steady state response. Proc Intl Soc Magn Reson Med 13:1421

    Google Scholar 

  • Purdon PL, et al. (2008) High-Field One: an open-source hardware and software platform for acquisition and real-time processing of electrophysiology during MRI. J Neurosci Methods 175(2):165–186

    Article  PubMed  Google Scholar 

  • Purdon PL, et al. Concurrent Recording of 40-Hz Auditory Steady State Response and Functional MR. in 10th Annual Meeting of the Organization for Human Brain Mapping. 2004. Budapest, Hungary: Neuroimage

    Google Scholar 

  • Rector DM, George JS (2001) Continuous image and electrophysiological recording with real-time processing and control. Methods 25(2):151–163

    Article  PubMed  CAS  Google Scholar 

  • Reilly JP (1992) Electrical stimulation and electropathology. Cambridge University Press, New York

    Google Scholar 

  • Rezai AR, Finelli D, Nyenhuis JA, Hrdlicka G, Tkach J, Sharan A, Rugieri P, Stypulkowski PH, Shellock FG (2002) Neurostimulation systems for deep brain stimulation: in vitro evaluation of magnetic resonance imaging-related heating at 1.5 Tesla. J Magn Reson Imaging 15(3):241–250

    Article  PubMed  Google Scholar 

  • Row R (1955) Theoretical and experimental study of electromagnetic scattering by two identical conducting cylinders. J Appl Phys 26(6):666–675

    Article  Google Scholar 

  • Schomer LD, et al. (2000) EEG linked functional magnetic resonance imaging in epilepsy and cognitive neurophysiology. J Clin Neurophysiol 17(1):43–58

    Article  PubMed  CAS  Google Scholar 

  • Sharan A, et al. (2003) MR safety in patients with implanted deep brain stimulation systems (DBS). Acta Neurochir Suppl 87:141–145

    PubMed  CAS  Google Scholar 

  • Shellock FG (2007) Comments on MR heating tests of critical implants. J Magn Reson Imaging 26(5):1182–1185

    Article  PubMed  Google Scholar 

  • Sled JG, Pike GB (1998) Standing-wave and RF penetration artifacts caused by elliptic geometry: an electrodynamic analysis of MRI. IEEE Trans Med Imaging 17(4):653–662

    Article  PubMed  CAS  Google Scholar 

  • Stach BA (2002) The auditory steady-state response: a primer. Hearing J 55(9):10–14

    Google Scholar 

  • Stevens, T., J. Ives, and R. Bartha. Energy Coupling between RF Electric Fields and Conductive Wires: Image Artifacts and Heating. in International Society for Magnetic Resonance in Medicine. 2007. Berlin, Germany

    Google Scholar 

  • Stevens, T., J. Ives, and R. Bartha. Avoiding Resonant Lengths of Wire with RF Chokes at 4 Tesla. in International Society for Magnetic Resonance in Medicine. 2007. Berlin, Germany

    Google Scholar 

  • Stevens TK, et al. (2007) MR compatibility of EEG scalp electrodes at 4 Tesla. J Magn Reson Imaging 25:872–877

    Article  PubMed  Google Scholar 

  • Vasios CE, et al. (2005) An ink cap for recording EEG during 7 Tesla MRI. In: 35th Annu Meet Soc Neuroscience, Washington, DC, USA, 12–16 Nov 2005

    Google Scholar 

  • Vasios CE, et al. (2006) EEG/(f)MRI measurements at 7 Tesla using a new EEG cap (“InkCap”). Neuroimage 33(4):1082–1092

    Article  PubMed  Google Scholar 

  • Wendt M, et al. Dynamic tracking in interventional MRI using wavelet-encoded gradient-echo sequences. IEEE Trans Med Imaging, 1998. 17(5): p. 803–809

    Article  PubMed  CAS  Google Scholar 

  • Wiesinger F, et al. (2006) Potential and feasibility of parallel MRI at high field. NMR Biomed 19(3):368–378

    Article  PubMed  Google Scholar 

  • Wiggins GC, et al. (2005) Eight-channel phased array coil and detunable TEM volume coil for 7 T brain imaging. Magn Reson Med 54(1):235–240

    Article  PubMed  CAS  Google Scholar 

  • Young J, Wait J (1989) Shielding properties of an ensemble of thin, infinitely long, parallel wires over a lossy half space. IEEE Trans Electromagn Compat 31(3):238–244

    Article  Google Scholar 

Download references

Acknowledgements

Thanks to Prof Richard Bowtell, of the Sir Peter Mansfield Magnetic Resonance Centre, University of Nottingham, and Dr David Carmichael of the UCL Institute of Neurology, for their input.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Giorgio Bonmassar .

Editor information

Editors and Affiliations

Appendices

Appendix 1: The Multidimensional Kalman Adaptive Filtering Method

We provide an outline of a 3D extension of the Kalman filter for pulse-related noise cancellation in multiple channels simultaneously based on the adaptive Kalman filter algorithm (Haykin 1996). The adaptive algorithm makes use of any correlation between each of the motion sensor signals m i (t) and the observed signal y(t c) to estimate the finite impulse response (FIR) kernel (k c) and to remove the noise signals n i (t, c). Since the true underlying (neuronal) EEG signal s(t,c) is uncorrelated with the motion signals m i (t), it is unaffected by the adaptive algorithm and, on average, the result of the noise cancellation process ŝ(t, c) will be the true underlying EEG.

Each channel “c” of the EEG signal y(t, c) is modeled as the sum of a “true” underlying EEG signal s(t, c) and a sum of 3D signals n x (t, c), n y (t, c) and n z (t, c) containing motion and ballistocardiogram components along the three cardinal directions:

$y(t,c) = s(t,c) + n_x (t,c) + n_y (t,c) + n_z (t,c)\quad c = 1, \ldots,32.$
((1))

Next, we enumerate the spatial components of both the noise n i (t,c), i = 1, 2, 3 (i.e. n x (t,c), n y (t,c) and n z (t,c)) and the motion sensor signals m i (t). The relationship between the noise signals n i (t,c) and the motion sensor signals m i (t) is modeled linearly using a time-varying, FIR kernel (k) with the equation:

$n_i (t,c) = \sum\limits_{k = 0}^{N_i - 1} {w_i^t (k,c)m_i (t - k)\quad i = 1,2,3,} $
((2))

where N i is the order of the FIR kernel (k, c), chosen to be the same for all channels. An adaptive filtering algorithm is used to produce an estimate of the FIR kernel (k, c), which in turn is used to estimate the noise signal n i (t, c). The sum of estimated noise components signals is then subtracted from the recorded signal y(t, c) to reveal the underlying EEG signal s(t, c),

$\displaylines{ \hat n_i (t,c) = \sum\limits_{k = 0}^{N - 1} {\hat w_i^t (k)m_i (t - k)}, \cr \hat s(t,c) = y(t,c) - \sum\limits_{i = 1}^3 {\hat n_i (t,c)}. \cr} $
((3))

In previous studies (Schomer et al. 2000), we showed that the ballistocardiogram is spatially smooth and that the first eigenvalue of the principal component analysis (PCA) alone accounts for over 90% of the variance. Building on those results, we consider adaptive filtering on the first Q components of the PCA of y(t,c) (i.e. ŷq (t): q = 1,…,Q), and select Q according to the desired resulting minimum fractional variance. Figure 6 shows the filtering scheme, which includes: (a) the dimensionality reduction for the 32-channel EEG signals by means of PCA, and (b) singular value decomposition (SVD) transformation along the principal direction of the 3D motion. Figure 6 also illustrates the algorithm with the simplified value of Q = 2, which is a remarkable reduction in computation time: 16 times faster than when Q = 2.

Fig. 6
figure 6

Adaptive-filtering multichannel scheme used for artifact removal. The input signal (decimated bipolar EEG), after band pass and decimation, is a signal of 32 channels that is first transformed into two principal component analysis (PCA) components. Each component is then separately filtered using the proposed Kalman filter. According to this filtering scheme, three motion sensor signals are first singular value decomposition (SVD) transformed and vectorized before Kalman filtering. After filtering, each “clean” PCA component is used to reconstruct the 32-channel EEG signal. There is a module of Z-10 for causality, given that the motion sensors have an intrinsic delay due to mechanical inertia. Z-L is used to compensate for the Z-10 and the delay in the FIR filter

The synchronization in Fig. 6 occurs because every channel contains a certain mixture of the two filtered PCA components. m 1(t) is the vector composed of the N samples from the first motion sensor prior to and equal to “t”, m 2(t) are the data from the second motion sensor, and m 3(t) are data from the third motion sensor. The matrix m(t) = [m 1(t) m 2(t) m 3(t)] is then separated into three components which are approximately mutually orthogonal, and the final result is obtained by diagonalizing the new motion-sensor covariance matrix \({\bf{\bar m}}(t)\):

$\begin{array}{l} {\rm w(}t + 1,q{\rm ) = }a{\rm w(}t,q{\rm ) + v}_{\rm 3} (t) \\ y_q (t) = {\rm \tilde m}^T (t) \cdot {\rm w}(t,q) + s_q (t) \\ {\rm w}(t,q) = \left[ {{\rm w}_{\rm 1}^t (0,q){\rm w}_{\rm 2}^t (0,q)\quad {\rm w}_{\rm 3}^t (0,q)\quad {\rm w}_{\rm 1}^t (1,q)\quad {\rm w}_{\rm 2}^t (1,q)\quad {\rm w}_{\rm 3}^t (1,q)\quad \ldots \quad {\rm w}_{\rm 1}^t (N,q)\quad {\rm w}_{\rm 2}^t (N,q)\quad {\rm w}_{\rm 3}^t (N,q)} \right]^T \\ {\rm \tilde m(}t{\rm ) = }\left[ {\tilde m_1 (t)\quad \tilde m_2 (t)\quad \tilde m_3 (t)\quad \tilde m_1 (t - 1)\quad \tilde m_2 (t - 1)\quad \tilde m_3 (t - 1)\quad \ldots \quad \tilde m_3 (t - N + 1)} \right]^{\rm T} . \\ \end{array}$
((4))

The proposed transformation (i.e. rotation) is found by taking the SVD of the original motion-sensor matrix, m(t) = U·S·VT, and transforming m(t) into a new matrix: \({\bf{\tilde m}}(t) = {\bf{m}}(t) \cdot {\bf{V}}^{\text{T}} \). \({\bf{\tilde m}}(t)\) is created by concatenating every row of \({\bf{\bar m}}(t)\) in a vector of 3N elements. Furthermore, we have expressed the filter tap weights (k,q) and the three components of the motion sensor m i (t) in vector form, with T denoting matrix transposition, and where v 3(t) can be a singular white noise matrix with the covariance matrix Λi,j = qiÅ3δiÅ3,jÅ3I, where Å3 denotes addition modulus 3, δi,j is the Kronecker delta, a is a scalar state-transition parameter, and s(t) is the underlying “true” EEG signal, modeled as a white noise with variance λ. We can then recursively estimate the filter taps using the Kalman filter with updated equations (Haykin 2002).

Appendix 2: The Open Hardware and Software Project. The High-Field One System for Real-Time EEG–fMRI

While many commercial and academic systems have been developed to perform EEG recordings during fMRI, rapidly evolving technical, clinical, and scientific requirements have created an opportunity for hardware and software systems that can be customized for specific electrophysiology-fMRI applications. Hardware platforms may require customization to enable a variety of recording types (e.g. electroencephalogram, local field potentials, or multiunit activity) while meeting the stringent and costly requirements of MRI safety and compatibility. Real-time signal processing tools are an enabling technology for electrophysiology-fMRI studies, particularly for application areas such as sleep, epilepsy, neurofeedback, and drug studies, yet real-time signal processing tools are difficult to develop. See the chapter “EEG Quality: The Image Acquisition Artifact” for a discussion of existing hardware solutions. In this section we outline this system, which we call ­High-Field One [i.e. HF-1 (Purdon et al.2008)]. Since high-field MRI is a very challenging environment, it imposes the following requirements: (a) a high dynamic range to avoid saturation during scanning, (b) low RF emissions, and (c) low eddy currents.

RF noise immunity was obtained in HF-1 despite the EEG-MRI requirements of: (1) an absence of ferromagnetic components, (2) resolution of input signals down to 100 nV (exceeding the requirements described in the chapter “The Added Value of EEG–fMRI in Imaging Neuroscience”), and (3) RF noise immunity in the link with the computer system outside the MRI. HF-1 has a bandwidth from DC to 4 kHz (no high-pass filter is needed thanks to the large dynamic range), and a 24-bit sampling rate of 20,833 S/s. It also has a complete set of digital circuits for analogue-to-digital conversion, storage, and USB connection with an external PC. HF-1 is connected to a laptop via a USB optical interface. Furthermore, an optical digital line is used to synchronize EEG acquisition with the PCMCIA digital acquisition card (National Instruments, Austin, TX, USA), enabling us to record the trigger signals from a parallel port of a stimulus PC in order to do real-time averaging of epochs.

Although the system (Fig. 7) has been successfully used at 1.5 and 3 T fields, it was specifically designed for very high fields and was therefore tested on humans and animals at 7 T with the following specifications for the gradient coils: maximum gradient strength 40 mT/m (whole-body) and 60 mT/m (head); slew rate of 200 T/m/s (whole body) and 333 T/m/s (head).

Fig. 7
figure 7

High-Field One (HF-1) hardware overview. The hardware is organized into three main stages: (1) an amplification system, (2) an analogue-to-digital converter (ADC), and (3) a programmable microcontroller unit (MCU) for controlling ADC functions. The EEG data are relayed outside the MRI via a USB-based optical link. A laptop system is used for data display, recording, integration with external signals such as event triggers and physiological monitoring, and real-time signal processing. The HF-1 hardware is compatible with MRI systems, has been tested at field strengths of up to 7 T, and can be used for both EEG and more general electrophysiological recordings such as local field potential or multiunit activity

Main Design Features

The first stage of the system increases the EEG SNR by amplifying the signals before relaying them to the rest of the system. The purpose of this first stage is twofold: (1) to further attenuate (Aw) the fraction of the RF pulse picked up by the Ink Cap, and (2) to amplify the EEG signal in order to improve its SNR, which is given by:

${\text{SNR}} = \frac{{P^{\text{s}} }}{{P^{\text{n}} + \lambda \rho _1 }},$
((5))

where Ps is the power of the observed signal (i.e. EEG), λ = (Ri + R)/Ri, Ri is the input resistance of the second-stage amplifier, ,\({P^n} = |V_{_{rms} }^{n} |2/(4RB)\) and ρ1 is the intrinsic noise power of the first stage amplifier in the band of interest, B. Hence, the SNR is linearly dependent on the first-stage gain.

The third and last stage (Fig. 7) consists of a microcontroller unit (MCU) that is used to supervise all of the functions of the instruments and the clock management circuit, which generates three clocks: (a) an MCU clock (i.e. 48 MHz), (b) an analogue-to-digital conversion (ADC) clock (i.e. 8 MHz), and (c) a USB clock (i.e. 6 MHz). The MCU synchronises, communicates through a handshake protocol, and controls the input of each ADC. The MCU also controls the internal USB section. The data from the ADC is streamed in a FIFO controller using an internal bus controlled by a line driver. The USB serial port section consists of a USB controller connected to the internal data bus. The USB section is connected to an external custom-shielded USB optical extension. The optical sampling clock section informs the acquisition computer of the sampling frequency and is connected to the MCU; it can be located up to 120 m away and is connected to a DAQ card in the acquisition computer for acquisition of the ERP trigger. A PC runs the Labview code and allows monitoring of the electrophysiological traces in real time. It also performs band-pass filtering, saves this large set of data in real time (Rector and George 2001), and computes steady-state responses and spectrograms, also in real time. Complete synchronisation of the MRI and EEG systems was achieved by generating a 48 MHz clock from the 10 MHz master clock of the Siemens Trio (Siemens AG, Erlangen, Germany) scanner using a direct digital synthesizer.

The HF-1 software and hardware designs are freely available at http://nmr.mgh.harvard.edu/abilab/, subject to a free licensing agreement. The HF-1 hardware system can be constructed based on the complete specifications provided. These specifications include circuit diagrams in Orcad (Cadence, San Jose, CA, USA) format, print circuit board layouts in both Orcad and Gerber (Gerber Systems Corporation) formats, and bills for materials specifying all required components and part numbers. Print circuit board layouts were designed to permit both automated pick-and-place assembly and soldering, as well as hand assembly and soldering. With these specifications, it is possible to achieve turnkey production of the complete hardware system, or components of it, as desired by the end-user.

Rights and permissions

Reprints and permissions

Copyright information

© 2009 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Bonmassar, G., Mullinger, K.J. (2009). Specific Issues Related to EEG-fMRI at B 0 > 3 T. In: Mulert, C., Lemieux, L. (eds) EEG - fMRI. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-540-87919-0_11

Download citation

  • DOI: https://doi.org/10.1007/978-3-540-87919-0_11

  • Published:

  • Publisher Name: Springer, Berlin, Heidelberg

  • Print ISBN: 978-3-540-87918-3

  • Online ISBN: 978-3-540-87919-0

  • eBook Packages: MedicineMedicine (R0)

Publish with us

Policies and ethics