Functional ultrasound imaging

Functional ultrasound imaging (fUS) is a medical ultrasound imaging technique of detecting or measuring changes in neural activities or metabolism, for example, the loci of brain activity, typically through measuring blood flow or hemodynamic changes. The method can be seen as an extension of Doppler imaging.

Main applications and features of functional ultrasound (fUS) imaging

Background

edit
 
Main brain functional imaging technique resolutions

Brain activation can either be directly measured by imaging electrical activity of neurons using voltage sensitive dyes, calcium imaging, electroencephalography, or magnetoencephalography, or indirectly by detecting hemodynamic changes in blood flow in the neurovascular systems through functional magnetic resonance imaging (fMRI), positron emission tomography (PET), Functional near-infrared spectroscopy (fNIRS), or Doppler ultrasonography )...[1]

Optical based methods generally provide the highest spatial and temporal resolutions; however, due to scattering, they are intrinsically limited to the investigation of the cortex. Thus, they are often used on animal models after partially removing or thinning the skull to allow for the light to penetrate into tissue. fMRI and PET, which measure the blood-oxygen level dependent (BOLD) signal, were the only techniques capable of imaging brain activation in depth. BOLD signal increases when neuronal activation exceeds oxygen consumption, where blood flow increases significantly. In fact, in-depth imaging of cerebral hemodynamic responses by fMRI, being noninvasive, paved the way for major discoveries in neurosciences in the early stage, and is applicable on humans. However, fMRI also suffers limitations. First, the cost and size of MR machines can be prohibitive. Also, spatially resolved fMRI is achieved at the expense of a substantial drop in temporal resolution and/or SNR. As a result, the imaging of transient events such as epilepsy is particularly challenging. Finally, fMRI is not appropriate for all clinical applications. For example, fMRI is rarely performed on infants because of specific issues concerning infant sedation.[2]

Like fMRI, Doppler based functional ultrasound approach are based on the neurovascular coupling and are thus limited by the spatiotemporal features of neurovascular coupling as they measure cerebral blood volume (CBV) changes. CBV is a pertinent parameter for functional imaging that is already used by other modalities such as intrinsic optical imaging or CBV-weighted fMRI. The spatiotemporal extent of CBV response was extensively studied. The spatial resolution of sensory-evoked CBV response can go down to cortical column (~100 μm). Temporally, the CBV impulse response function was measured to typically start at ~0.3 s and peak at ~1 s in response to ultrashort stimuli (300μs), which is much slower than the underlying electrical activity.[3]

Conventional doppler based approaches

edit

Hemodynamic changes in the brain are often used as a surrogate indicator of neuronal activity to map the loci of brain activity. Major part of the hemodynamic response occurs in small vessels; however, conventional Doppler ultrasound is not sensitive enough to detect the blood flow in such small vessels.[2]

Functional transcranial doppler (fTCD)

edit

Ultrasound Doppler imaging can be used to obtain basic functional measurements of brain activity using blood flow. In functional transcranial Doppler sonography, a low frequency (1-3 MHz) transducer is used through the temporal bone window with a conventional pulse Doppler mode to estimate blood flow at a single focal location. The temporal profile of blood velocity is usually acquired in main large arteries such as the middle cerebral artery (MCA). The peak velocity is compared between rest and task conditions or between right and left sides when studying lateralization.[4]

Power doppler

edit

Power Doppler is a Doppler sequence that measures the ultrasonic energy backscattered from red blood cells in each pixel of the image. It provides no information on blood velocity but is proportional to blood volume within the pixel. However, conventional power Doppler imaging lacks sensitivity to detect small arterioles/venules and thus is unable to provide local neurofunctional information through neurovascular coupling.[2]

Ultrasensitive doppler

edit

Functional ultrasound imaging was pioneered at ESPCI by Mickael Tanter [fr]'s team[5] following work on ultrafast imaging[6] and ultrafast Doppler.[7]

Ultrasensitive doppler principle

edit

Ultrasensitive Doppler relies on ultrafast imaging scanners[6] able to acquire images at thousands of frames per second, thus boosting the power Doppler SNR without any contrast agents. Instead of the line by line acquisition of conventional ultrasound devices, ultra-fast ultrasound takes advantage of successive tilted plane wave transmissions that afterward coherently compounded to form images at high frame rates. Coherent Compound Beamforming consists of the recombination of backscattered echoes from different illuminations achieved on the acoustic pressure field with various angles (as opposed to the acoustic intensity for the incoherent case). All images are added coherently to obtain a final compounded image. This very addition is produced without taking the envelope of the beamformed signals or any other nonlinear procedure to ensure a coherent addition. As a result, coherent adding of several echo waves leads to cancellation of out-of-phase waveforms, narrowing the point spread function (PSF), and thus increasing spatial resolution. A theoretical model demonstrates that the gain in sensitivity of the ultrasensitive Doppler method is due to the combination of the high signal-to-noise ratio (SNR) of the gray scale images, due to the synthetic compounding of backscattered echoes and the extensive signal samples averaging due to the high temporal resolution of ultrafast frame rates.[2] The sensitivity was recently further improved using multiple plane wave transmissions[8] and advanced spatiotemporal clutter filters for better discrimination between low blood flow and tissue motion. Ultrasound researchers have been using ultrafast imaging research platforms with parallel acquisition of channels and custom sequences programming to investigate ultrasensitive Doppler/fUS modalities. A custom real-time high-performance GPU beamforming code with a high data transfer rate (several GBytes per second) must then be implemented to perform imaging at high frame rate. Acquisitions can also typically easily provide gigabytes of data depending on acquisition duration.

Ultrasensitive Doppler has a typical 50-200 μm spatial resolution depending on the ultrasound frequency used.[2] It features a temporal resolution in the tens of milliseconds, can image the full depth of the brain and can provide 3D angiography.[9]

functional Ultrasound imaging

edit

This signal boost enables the sensitivity required to map subtle blood variations in small arterioles (down to 1mm/s) related to neuronal activity. By applying an external stimulus such as a sensory, auditory or visual stimulation, it is then possible to construct a map of brain activation from the ultrasensitive Doppler movie.

fUS measures indirectly cerebral blood volume which provides an effect size close to 20% and as such is quite more sensitive than fMRI whose BOLD response is typically only a couple of percents. Correlation maps or statistical parametric maps can be constructed to highlight the activated areas. fUS has been shown to have a spatial resolution on the order 100 micrometers at 15 MHz in ferrets[10] and is sensitive enough to perform single trial detection in awake primates.[11] Other fMRI-like modalities such as functional connectivity can also be implemented.

Commercial scanners with specialized hardware and software[12] are enabling fUS to rapidly expand behind ultrasound research labs to the neuroscience community.

4D functional ultrasound imaging

edit

Some researchers conducted 4D functional ultrasound imaging of whole-brain activity in rodents. Currently, two different technological solutions are proposed for the acquisition of 3D and 4D fUS data, each with its own advantages and drawbacks.[13] The first is a tomographic approach based on motorized translation of linear probes. This approach proved to be a successful method for several applications such as 3D retinotopic mapping in the rodent brain[14][15] and 3D tonotopic mapping of the auditory system in ferrets.[10] The second approach relies on high frequency 2D matrix array transducer technology coupled with a high channel count electronic system for fast 3D imaging. To counterbalance the intrinsically poor sensitivity of matrix elements, they devised a 3D multiplane-wave scheme with 3D spatiotemporal encoding of transmit signals using Hadamard coefficients. For each transmission, the backscattered signals containing mixed echoes from the different plane waves are decoded using the summation of echoes from successive receptions with appropriate Hadamard coefficients. This summation enables the synthetic building of echoes from a virtual individual plane wave transmission with a higher amplitude. Finally, they perform coherent compounding beamforming of decoded echoes to produce 3D ultrasonic images and apply a spatiotemporal clutter filter separating blood flow from tissue motion to compute a power Doppler volume, which is proportional to the cerebral blood volume.[16]

Applications

edit

Preclinical

edit
 
Preclinical applications of fUS imaging

fUS can benefit in monitoring cerebral function in the whole brain which is important to understanding how the brain works on a large scale under normal or pathological conditions. The ability to image cerebral blood volume at high spatiotemporal resolution and with high sensitivity using fUS could be of great interest for applications in which fMRI reaches its limits, such as imaging of epileptic-induced changes in blood volume.[5] fUS can be applied for chronic studies in animal models through a thinned-skull[17] or smaller cranial window or directly through the skull in mice.

Brain activity mapping

edit

Tonotopics or retinotopics maps[18] can be constructed by mapping the response of frequency-varying sounds[10] or moving visual targets.[14][18][15]

Functional connectivity / resting state

edit

When no stimulus is applied, fUS can be used to study functional connectivity during resting state. The method has been demonstrated in rats[19] and awake mice[20] and can be used for pharmacological studies when testing drugs.[21] Seed-based maps, independent component analysis of resting states modes or functional connectivity matrix between atlas-based regions of interests can be constructed with high resolution.

Awake fUS imaging

edit

Using dedicated ultralight probes, it is possible to perform freely-moving experiments in rats or mice.[22][23] The size of the probes and electromagnetic-compatibility of fUS means it can also be used easily on head-fixed setups for mice[15] or in electrophysiology chambers in primate.[11]

Clinical

edit
 
Clinical neuroimaging using ultrasound

Neonates

edit

Thanks to its portability, fUS has also been used in clinics in awake neonates.[24] Functional ultrasound imaging can be applied to neonatal brain imaging in a non-invasive manner through the fontanel window. Ultrasound is usually performed in this case, which means that the current procedures does not have to be changed. High quality angiographic images could help diagnose vascular diseases such as perinatal ischemia or ventricular hemorrhage.

Adults / intraoperative

edit

For adults, this method can be used during neurosurgery to guide the surgeon through the vasculature and to monitor the patient's brain function prior to tumor resection[25][26]

See also

edit

References

edit
  1. ^ Petersen CC (October 2007). "The functional organization of the barrel cortex". Neuron. 56 (2): 339–55. doi:10.1016/j.neuron.2007.09.017. PMID 17964250.
  2. ^ a b c d e Mace E, Montaldo G, Osmanski BF, Cohen I, Fink M, Tanter M (March 2013). "Functional ultrasound imaging of the brain: theory and basic principles". IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control. 60 (3): 492–506. doi:10.1109/tuffc.2013.2592. PMID 23475916. S2CID 27482186.
  3. ^ Deffieux T, Demene C, Pernot M, Tanter M (June 2018). "Functional ultrasound neuroimaging: a review of the preclinical and clinical state of the art". Current Opinion in Neurobiology. 50: 128–135. doi:10.1016/j.conb.2018.02.001. PMID 29477979.
  4. ^ Knecht S, Deppe M, Ebner A, Henningsen H, Huber T, Jokeit H, et al. (January 1998). "Noninvasive determination of language lateralization by functional transcranial Doppler sonography: a comparison with the Wada test". Stroke. 29 (1): 82–6. doi:10.1161/01.str.29.1.82. PMID 9445333.
  5. ^ a b Macé E, Montaldo G, Cohen I, Baulac M, Fink M, Tanter M (July 2011). "Functional ultrasound imaging of the brain". Nature Methods. 8 (8): 662–664. doi:10.1038/nmeth.1641. PMID 21725300.
  6. ^ a b Tanter M, Fink M (January 2014). "Ultrafast imaging in biomedical ultrasound". IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control. 61 (1): 102–19. doi:10.1109/TUFFC.2014.6689779. PMID 24402899.
  7. ^ Bercoff J, Montaldo G, Loupas T, Savery D, Mézière F, Fink M, et al. (January 2011). "Ultrafast compound Doppler imaging: providing full blood flow characterization". IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control. 58 (1): 134–47. doi:10.1109/TUFFC.2011.1780. PMID 21244981.
  8. ^ Tiran E, Deffieux T, Correia M, Maresca D, Osmanski BF, Sieu LA, et al. (November 2015). "Multiplane wave imaging increases signal-to-noise ratio in ultrafast ultrasound imaging". Physics in Medicine and Biology. 60 (21): 8549–66. Bibcode:2015PMB....60.8549T. doi:10.1088/0031-9155/60/21/8549. PMID 26487501.
  9. ^ Demené C, Tiran E, Sieu LA, Bergel A, Gennisson JL, Pernot M, et al. (February 2016). "4D microvascular imaging based on ultrafast Doppler tomography". NeuroImage. 127: 472–483. doi:10.1016/j.neuroimage.2015.11.014. PMID 26555279.
  10. ^ a b c Bimbard C, Demene C, Girard C, Radtke-Schuller S, Shamma S, Tanter M, et al. (June 2018). "Multi-scale mapping along the auditory hierarchy using high-resolution functional UltraSound in the awake ferret". eLife. 7. doi:10.7554/eLife.35028. PMC 6039176. PMID 29952750.
  11. ^ a b Dizeux A, Gesnik M, Ahnine H, Blaize K, Arcizet F, Picaud S, et al. (March 2019). "Functional ultrasound imaging of the brain reveals propagation of task-related brain activity in behaving primates". Nature Communications. 10 (1): 1400. Bibcode:2019NatCo..10.1400D. doi:10.1038/s41467-019-09349-w. PMC 6438968. PMID 30923310.
  12. ^ "Iconeus Functional Ultrasound (fUS) preclinical imaging systems - Iconeus". iconeus.com. 2021-08-13. Retrieved 2024-06-15.
  13. ^ "The path to 4D fUS" (PDF). Iconeus. Retrieved 25 May 2020.
  14. ^ a b Gesnik M, Blaize K, Deffieux T, Gennisson JL, Sahel JA, Fink M, et al. (April 2017). "3D functional ultrasound imaging of the cerebral visual system in rodents". NeuroImage. 149: 267–274. doi:10.1016/j.neuroimage.2017.01.071. PMC 5387157. PMID 28167348.
  15. ^ a b c Macé É, Montaldo G, Trenholm S, Cowan C, Brignall A, Urban A, et al. (December 2018). "Whole-Brain Functional Ultrasound Imaging Reveals Brain Modules for Visuomotor Integration". Neuron. 100 (5): 1241–1251.e7. doi:10.1016/j.neuron.2018.11.031. PMC 6292977. PMID 30521779.
  16. ^ Rabut C, Correia M, Finel V, Pezet S, Pernot M, Deffieux T, et al. (October 2019). "4D functional ultrasound imaging of whole-brain activity in rodents". Nature Methods. 16 (10): 994–997. doi:10.1038/s41592-019-0572-y. PMC 6774790. PMID 31548704.
  17. ^ Drew PJ, Shih AY, Driscoll JD, Knutsen PM, Blinder P, Davalos D, et al. (December 2010). "Chronic optical access through a polished and reinforced thinned skull". Nature Methods. 7 (12): 981–4. doi:10.1038/nmeth.1530. PMC 3204312. PMID 20966916.
  18. ^ a b Blaize K, Arcizet F, Gesnik M, Ahnine H, Ferrari U, Deffieux T, et al. (June 2020). "Functional ultrasound imaging of deep visual cortex in awake nonhuman primates". Proceedings of the National Academy of Sciences of the United States of America. 117 (25): 14453–14463. Bibcode:2020PNAS..11714453B. doi:10.1073/pnas.1916787117. PMC 7321983. PMID 32513717.
  19. ^ Osmanski BF, Pezet S, Ricobaraza A, Lenkei Z, Tanter M (October 2014). "Functional ultrasound imaging of intrinsic connectivity in the living rat brain with high spatiotemporal resolution". Nature Communications. 5: 5023. Bibcode:2014NatCo...5.5023O. doi:10.1038/ncomms6023. PMC 4205893. PMID 25277668.
  20. ^ Ferrier J, Tiran E, Deffieux T, Tanter M, Lenkei Z (June 2020). "Functional imaging evidence for task-induced deactivation and disconnection of a major default mode network hub in the mouse brain". Proceedings of the National Academy of Sciences of the United States of America. 117 (26): 15270–15280. Bibcode:2020PNAS..11715270F. doi:10.1073/pnas.1920475117. PMC 7334502. PMID 32541017.
  21. ^ Rabut C, Ferrier J, Bertolo A, Osmanski B, Mousset X, Pezet S, et al. (November 2020). "Pharmaco-fUS: Quantification of pharmacologically-induced dynamic changes in brain perfusion and connectivity by functional ultrasound imaging in awake mice". NeuroImage. 222: 117231. doi:10.1016/j.neuroimage.2020.117231. PMID 32795659.
  22. ^ Sieu LA, Bergel A, Tiran E, Deffieux T, Pernot M, Gennisson JL, et al. (September 2015). "EEG and functional ultrasound imaging in mobile rats". Nature Methods. 12 (9): 831–834. doi:10.1038/nmeth.3506. PMC 4671306. PMID 26237228.
  23. ^ Tiran E, Ferrier J, Deffieux T, Gennisson JL, Pezet S, Lenkei Z, et al. (August 2017). "Transcranial Functional Ultrasound Imaging in Freely Moving Awake Mice and Anesthetized Young Rats without Contrast Agent". Ultrasound in Medicine & Biology. 43 (8): 1679–1689. doi:10.1016/j.ultrasmedbio.2017.03.011. PMC 5754333. PMID 28476311.
  24. ^ Demené C, Mairesse J, Baranger J, Tanter M, Baud O (January 2019). "Ultrafast Doppler for neonatal brain imaging". NeuroImage. 185: 851–856. doi:10.1016/j.neuroimage.2018.04.016. PMID 29649559.
  25. ^ Imbault M, Chauvet D, Gennisson JL, Capelle L, Tanter M (August 2017). "Intraoperative Functional Ultrasound Imaging of Human Brain Activity". Scientific Reports. 7 (1): 7304. Bibcode:2017NatSR...7.7304I. doi:10.1038/s41598-017-06474-8. PMC 5544759. PMID 28779069.
  26. ^ Soloukey S, Vincent AJ, Satoer DD, Mastik F, Smits M, Dirven CM, et al. (2019). "Functional Ultrasound (fUS) During Awake Brain Surgery: The Clinical Potential of Intra-Operative Functional and Vascular Brain Mapping". Frontiers in Neuroscience. 13: 1384. doi:10.3389/fnins.2019.01384. PMC 6962116. PMID 31998060.