This is a nice article illustrating the potential life altering pro...
To give a sense of the team's background, there are neurosurgeons (...
More background on the brain-computer interface: First researched ...
TSNE *t-distributed stochastic neighbor embedding* (*t-SNE*) is a ...
> "Together, these results suggest that, even years after paralysis...
> "Next, we tested whether we could decode complete handwritten sen...
Recurrent Neural Networks (RNNs) are a class of neural networks tha...
572 training sentences doesn't sound like a huge amount of training...
A character error rate of .89% is extremely low!
Amazing result! > "To our knowledge, 90characters per minute is t...
> "It is important to recognize that the current system is a proof ...
Nature | Vol 593 | 13 May 2021 | 249
Article
High-performance brain-to-text
communication via handwriting
Francis R. Willett
1,2,3
 ✉
, Donald T. Avansino
1
, Leigh R. Hochberg
4,5,6,7
, Jaimie M. Henderson
2,8,9,12
& Krishna V. Shenoy
1,3,8,9,10,11,12
Brain–computer interfaces (BCIs) can restore communication to people who have
lost the ability to move or speak. So far, a major focus of BCI research has been on
restoring gross motor skills, such as reaching and grasping
1–5
or point-and-click typing
with a computer cursor
6,7
. However, rapid sequences of highly dexterous behaviours,
such as handwriting or touch typing, might enable faster rates of communication.
Here we developed an intracortical BCI that decodes attempted handwriting
movements from neural activity in the motor cortex and translates it to text in real
time, using a recurrent neural network decoding approach. With this BCI, our study
participant, whose hand was paralysed from spinal cord injury, achieved typing
speeds of 90characters per minute with 94.1% raw accuracy online, and greater than
99% accuracy oine with a general-purpose autocorrect. To our knowledge, these
typing speeds exceed those reported for any other BCI, and are comparable to typical
smartphone typing speeds of individuals in the age group of our participant
(115characters per minute)
8
. Finally, theoretical considerations explain why
temporally complex movements, such as handwriting, may be fundamentally easier
to decode than point-to-point movements. Our results open a new approach for BCIs
and demonstrate the feasibility of accurately decoding rapid, dexterous movements
years after paralysis.
Previous BCI studies have shown that the motor intention for gross
motor skills, such as reaching, grasping or moving a computer cur-
sor, remains neurally encoded in the motor cortex after paralysis
1–7
.
However, it is still unknown whether the neural representation for
a rapid and highly dexterous motor skill, such as handwriting, also
remains intact. We tested this by recording neural activity from two
microelectrode arrays in the hand ‘knob’ area of the precentral gyrus
(a premotor area)
9,10
while our BrainGate study participant, T5,
attempted to handwrite individual letters and symbols (Fig.1a). T5
has a high-level spinal cord injury and was paralysed from the neck
down; his hand movements were entirely non-functional and limited
to twitching and micromotion. We instructed T5 to ‘attempt’ to write
as if his hand were not paralysed, while imagining that he was holding
a pen on a piece of ruled paper.
Neural representation of handwriting
To visualize the neural activity (multiunit threshold crossing rates)
recorded during attempted handwriting, we used principal compo-
nents analysis to display the top three neural dimensions that contain
the most variance (Fig.1b). The neural activity appeared to be strong
and repeatable, although the timing of its peaks and valleys varied
across trials, potentially owing to fluctuations in writing speed. We
used a time-alignment technique to remove temporal variability
11
,
which revealed notably consistent underlying patterns of neural activ-
ity that are unique to each character (Fig.1c). To ascertain whether the
neural activity encoded the pen movements that are needed to draw
each character’s shape, we attempted to reconstruct each character by
linearly decoding the pen-tip velocity from the trial-averaged neural
activity (Fig.1d). Readily recognizable letter shapes confirmed that
pen-tip velocity is robustly encoded. The neural dimensions that repre-
sented pen-tip velocity accounted for 30% of the total neural variance.
Next, we used a nonlinear dimensionality reduction method
(t-distributed stochastic neighbour embedding; t-SNE) to produce a
two-dimensional (2D) visualization of each single trial’s neural activity
recorded after the ‘go’ cue was given (Fig.1e). The t-SNE visualization
revealed tight clusters of neural activity for each character and a pre-
dominantly motoric encoding in which characters that are written simi-
larly are closer together. Using a k-nearest-neighbour classifier applied
offline to the neural activity, we could classify the characters with 94.1%
accuracy (95% confidence interval (CI)=[92.6, 95.8]). Together, these
results suggest that, even years after paralysis, the neural representa-
tion of handwriting in the motor cortex is probably strong enough to
be useful for a BCI.
https://doi.org/10.1038/s41586-021-03506-2
Received: 2 July 2020
Accepted: 26 March 2021
Published online: 12 May 2021
Check for updates
1
Howard Hughes Medical Institute at Stanford University, Stanford, CA, USA.
2
Department of Neurosurgery, Stanford University School of Medicine, Stanford, CA, USA.
3
Department of Electrical
Engineering, Stanford University, Stanford, CA, USA.
4
VA RR&D Center for Neurorestoration and Neurotechnology, Rehabilitation R&D Service, Providence VA Medical Center, Providence, RI,
USA.
5
School of Engineering, Brown University, Providence, RI, USA.
6
Carney Institute for Brain Science, Brown University, Providence, RI, USA.
7
Center for Neurotechnology and Neurorecovery,
Department of Neurology, Massachusetts General Hospital, Harvard Medical School, Boston, MA, USA.
8
Wu Tsai Neurosciences Institute, Stanford University, Stanford, CA, USA.
9
Bio-X Institute,
Stanford University, Stanford, CA, USA.
10
Department of Bioengineering, Stanford University, Stanford, CA, USA.
11
Department of Neurobiology, Stanford University, Stanford, CA, USA.
12
These
authors jointly supervised this work: Jaimie M. Henderson, Krishna V. Shenoy.
e-mail: fwillett@stanford.edu
250 | Nature | Vol 593 | 13 May 2021
Article
Decoding handwritten sentences
Next, we tested whether we could decode complete handwritten sen-
tences in real time, thus enabling an individual with tetraplegia to com-
municate by attempting to handwrite their intended message. To do
so, we trained a recurrent neural network (RNN) to convert the neural
activity into probabilities describing the likelihood of each character
being written at each moment in time (Fig.2a, Extended Data Fig.1).
These probabilities could either be thresholded in a simple way to emit
discrete characters, which we did for real-time decoding (‘raw online
output’, Fig.2a), or processed more extensively by a large-vocabulary
language model to simulate an autocorrect feature, which we applied
offline (‘offline output from a language model’, Fig.2a). We used the lim-
ited set of 31characters shown in Fig.1d, consisting of the 26 lower-case
letters of the alphabet, together with commas, apostrophes, question
marks, full stops (written by T5 as a tilde symbol; ‘~’) and spaces (written
by T5 as a greater-than symbol; ‘>’). The ‘~’ and ‘>’ symbols were chosen
to make full stops and spaces easier to detect. T5 attempted to write
each character in print (not cursive), with each character printed on
top of the previous one.
To collect training data for the RNN, we recorded neural activity
while T5 attempted to handwrite complete sentences at his own pace,
following instructions on a computer monitor. Before the first day of
real-time evaluation, we collected a total of 242 sentences across 3 pilot
days that were combined to train the RNN. On each subsequent day of
real-time testing, additional training data were collected to recalibrate
the RNN before evaluation, yielding a combined total of 572 training
sentences by the last day (comprising 7.6 hours and 31,472 characters).
To train the RNN, we adapted neural network methods in automatic
speech recognition
1214
to overcome two key challenges: (1) the time
that each letter was written in the training data was unknown (as T5’s
hand was paralysed), making it challenging to apply supervised learning
Prepare: a Go
Return
Delay
(2.0–3.0 s)
1.0 s
a
PC 1 PC 2
d
m
e
a
b
c
d
e
f
g
h i
j
k l
m n
o
p q
r
s
t
u
v w x
y
z
>
, (comma)
' (apostrophe) ~ (tilde)
?
de
Time (s)
0
1
Trial no.
1
27
PC 1 PC 2
d
m
e
Trial no.
1
27
Time (s)
0
1
b
c
Low
High
Time
warping
Time
Warped
time
PC score
Prepare: m
y
>
x
b
p
k
m
w
h
n
r
v
u
~
d
o
q
a
e
z
c
g
s
t
f
?
'
(apostrophe)
,
(comma)
j
i
l
PC 3
PC 3
Fig. 1 | Neural representation of attempted handwriting. a, To assess the
neural representation of attempted handwriting, participant T5 attempted to
handwrite each character one at a time, following the instructions given on a
computer screen (bottom panels depict what is shown on the screen, following
the timeline).Credit: drawingof the human silhouette created by E. Woodrum.
b, Neural activity in the top 3 principal components (PCs) is shown for three
example letters (d, e and m) and 27 repetitions of each letter (trials). The
colour scale was normalized within each panel separately for visualization.
c, Time-warping the neural activity to remove trial-to-trial changes in writing
speed reveals consistent patterns of activity unique to each letter. In the inset
above c, example time-warping functions are shown for the letter ‘m’ and lie
relatively close to the identity line (the warping function of each trial is plotted
with a different coloured line). d, Decoded pen trajectories are shown for all 31
tested characters. Intended 2D pen-tip velocity was linearly decoded from the
neural activity using cross-validation (each character was held out), and
decoder output was denoised by averaging across trials. Orange circles denote
the start of the trajectory. e, A 2D visualization of the neural activity made using
t-SNE. Each circle is a single trial (27 trials are shown for each of 31 characters).
Nature | Vol 593 | 13 May 2021 | 251
techniques; and (2) the dataset was limited in size compared to typical
RNN datasets, making it difficult to prevent overfitting to the training
data (seeSupplementary Methods, Extended Data Figs.2, 3).
We evaluated the performance of the RNN over a series of 5 days,
each day containing 4 evaluation blocks of 7–10 sentences that the RNN
was never trained on (thus ensuring that the RNN could not overfit to
those sentences). T5 copied each sentence from an on-screen prompt,
attempting to handwrite it letter by letter, while the decoded charac-
ters appeared on the screen in real time as they were detected by the
RNN (Supplementary Videos1, 2, Extended Data Table1). Characters
appeared after they were completed by T5 with a short delay (estimated
to be 0.4–0.7s). The decoded sentences were quite legible (‘raw output’,
Fig.2b). Notably, typing speeds were high, plateauing at 90characters
per minute with a mean character error rate of 5.4% (averaged across
all four blocks on the final day) (Fig.2c). As there was no ‘backspace’
function implemented, T5 was instructed to continue writing if any
decoding errors occurred.
When a language model was used to autocorrect errors offline, error
rates decreased considerably (Fig.2c, Table1). The character error rate
decreased to 0.89% and the word error rate decreased to 3.4% aver-
aged across all days, which is comparable to state-of-the-art speech
recognition systems with word error rates of 4–5%
14,15
, putting it well
within the range of usability. Finally, to probe the limits of possible
decoding performance, we trained a new RNN offline using all available
sentences to process the entire sentence in a non-causal way (com-
parable to other BCI studies
16,17
). Accuracy was extremely high in this
regime (0.17% character error rate), indicating a high potential ceiling
of performance, although this decoder would not be able to provide
letter-by-letter feedback to the user.
Next, to evaluate performance in a less restrained setting, we col-
lected two days of data in which T5 used the BCI to freely type answers
to open-ended questions (Supplementary Video3, Extended Data
Table2). The results confirm that high performance can also be
achieved when the user writes self-generated sentences as opposed
to copying on-screen prompts (73.8characters per minute, with a char-
acter error rate of 8.54% in real time and 2.25% with a language model).
To our knowledge, the previous record for free typing in intracortical
BCIs is 24.4correct characters per minute
7
.
Daily decoder retraining
Following standard practice
1,2,4,5,18
, we retrained our handwriting
decoder each day before evaluating it, with the help of ‘calibration’
data collected at the beginning of each day. Retraining helps to account
Probabilities
Threshold crossing features
RNN
... the paper ...
t
h
n
t
a
e
0.1
0.9
1.0
0.3
0.7
Raw online
output
... tne paper ...
Thresholding
(online)
Character
t
New
a
c
b
1,211
1,218
Trial day
1,220
1,237
1,239
1,211
1,218
Trial day
1,220
1,237
1,239
h
t
x
t
p
t–d
h
e
n
Language model
24 26 28
Time (s)
50
100
150
Electrode no.
24 26 28
0
1
24 26 28
Time (s)
0
1
n
e
>
p a p e r
character
t
h
e
n
t
h
e
n
t
Raw output
With ofine
language model
Offline output
from the
language model
Viterbi search
(offline)
0
Prompt: daisy leaned forward, at once horried and fascinated.
Raw output:
you must be the change you wish to see in the worldPrompt:
Raw output:
10 20 30 40
0 10
Time (s)
Time (s)
20 30
Character error rate (%)
Characters per minute
0
5
10
15
20
0
20
40
60
80
100
Previous intracortical BCIs
Fig. 2 | Neural decoding of attempted handwriting in real time. a, Diagram
of the decoding algorithm. First, the neural activity (multiunit threshold
crossings) was temporally binned and smoothed on each electrode (20-ms
bins). Then, an RNN converted this neural population time series (x
t
) into a
probability time series (p
td
) describing the likelihood of each character and the
probability of any new character beginning. The RNN had a one second output
delay (d), giving it time to observe each character fully before deciding its
identity. Finally, the character probabilities were thresholded to produce the
‘raw online output’ for real-time use (when the ‘new character’ probability
crossed a threshold at time t, the most likely character at time t+0.3s was
emitted and shown on the screen). In an offline retrospective analysis, the
character probabilities were combined with a large-vocabulary language
model to decode the most likely text that the participant wrote (using a custom
50,000-word bigram model).Credit: drawingof the human silhouette created
by E. Woodrum. b, Two real-time example trials are shown, demonstrating the
ability of the RNN to decode readily understandable text on sentences on which
it was never trained. Errors are highlighted in red and spaces are denoted with
‘>’. c, Error rates (edit distances) and typing speeds are shown for 5 days,
with 4 blocks of 7–10 sentences each (each block is indicated with a single circle
and coloured according to the trial day). The speed is more than double that of
the next-fastest intracortical BCI
7
, which is indicated with the dashed line.
252 | Nature | Vol 593 | 13 May 2021
Article
for changes in neural recordings that accrue over time, which might be
caused by neural plasticity or electrode array micromotion. Ideally, to
reduce the burden on the user, minimal or no calibration data would be
required. In a retrospective analysis of the copy-typing data reported
above in Fig.2, we assessed whether high performance could still have
been achieved using fewer than the original 50 calibration sentences
per day (Fig.3a). We found that 10 sentences (8.7min) were enough to
achieve a raw error rate of 8.5% (1.7% with a language model), although
30 sentences were needed to match the raw online error rate of 5.9%.
However, our copy-typing data were collected over a 28-day time
span, possibly allowing larger changes in neural activity to accumulate.
Using further offline analyses, we assessed whether sessions that are
more closely spaced reduce the need for calibration data (Fig.3b). We
found that when only 2–7days passed between sessions, performance
was reasonable with no decoder retraining (11.1% raw error rate, 1.5%
with a language model), as might be expected from previous work show-
ing the short-term stability of neural recordings
19–21
. Finally, we tested
whether decoders could be retrained in an unsupervised manner by
using a language model to error-correct and retrain the decoder, thus
bypassing the need to interrupt the user for calibration (by enabling
automatic recalibration during normal use). Encouragingly, unsuper-
vised retraining achieved a raw error rate of 7.3% (0.84% with a language
model) when sessions were separated by a time span of 7days or less.
Ultimately, whether decoders can be successfully retrained with
minimal recalibration data depends on how quickly the neural activity
changes over time. We assessed the stability of the neural patterns asso-
ciated with each character and found high short-term stability (mean
correlation of 0.85 when 7days apart or less), and neural changes that
seemed to accumulate at a steady and predictable rate (Extended Data
Fig.4). These results are promising for clinical viability, as they suggest
that unsupervised decoder retraining, combined with more-limited
supervised retraining after longer periods of inactivity, may be suf-
ficient to achieve high performance. Nevertheless, future work must
confirm this online, as offline simulations are not always predictive of
online performance.
Temporal variety improves decoding
To our knowledge, 90characters per minute is the highest typ-
ing rate that has yet been reported for any type of BCI (see ‘Discus-
sion’). For intracortical BCIs, the best-performing method has been
point-and-click typing with a 2D computer cursor, which peaks at
40correct characters per minute
7
(see Supplementary Video4 for a
direct comparison). The speed of point-and-click BCIs is limited primar-
ily by decoding accuracy. During parameter optimization, the cursor
gain is increased so as to increase typing rate, until the cursor moves
so quickly that it becomes uncontrollable owing to decoding errors
22
.
We therefore asked how handwriting movements could be decoded
more than twice as fast, with similar levels of accuracy.
We theorize that handwritten letters may be easier to distinguish
from each other than point-to-point movements, as letters have more
variety in their spatiotemporal patterns of neural activity than do
straight-line movements. To test this theory, we analysed the patterns
of neural activity associated with 16 straight-line movements and 16
letters that required no lifting of the pen off the page, both performed
by T5 with attempted handwriting (Fig.4a, b).
First, we analysed the pairwise Euclidean distances between each
neural activity pattern. We found that the nearest-neighbour distances
for each movement were 72% larger for characters compared to straight
lines (95% CI=[60%, 86%]), making it less likely for a decoder to con-
fuse two nearby characters (Fig.4c). To confirm this, we simulated
the classification accuracy for each set of movements as a function
of neural noise (Fig.4d), which showed that characters are easier to
classify than straight lines.
To gain insight into what might be responsible for the relative
increase in nearest-neighbour distances for characters, we examined
the spatial and temporal dimensionality of the neural patterns. Spatial
and temporal dimensionality were estimated using the ‘participation
ratio’ of the principal component analysis (PCA) eigenvalue spectrum,
which quantifies approximately how many spatial or temporal dimen-
sions are required to explain 80% of the variance in the patterns of
neural activity
23
. We found that the spatial dimensionality was only
modestly larger for characters (1.24 times larger; 95% CI=[1.19, 1.30]),
but that the temporal dimensionality was much greater (2.65 times
larger; 95% CI=[2.58, 2.72]), suggesting that the increased variety of
temporal patterns in letter writing drives the increased separability of
each movement (Fig.4e).
To illustrate how increased temporal dimensionality can make move-
ments more distinguishable, we constructed a toy model with four
movements and two neurons, with the neural activity constrained to lie
along a single dimension (Fig.4f, g). Simply by allowing the trajectories
to change in time (Fig.4g), the nearest-neighbour distance between
the neural trajectories can be increased, resulting in an increase in
classification accuracy when noise levels are large enough (Fig.4h).
Although neural noise in the toy model was assumed to be independ-
ent white noise, we found that these results also hold for noise that
No r
etraining
5 sentences
10 sentences
20 sentences
30 sentences
40 sentences
50 sentences
0
10
20
30
Character error rate (%)
Raw
Language model
No retraining
5 sentences
10 sentences
Unsupervised
0
10
20
30
Character error rate (%)
2–7 days
8–14 days
15–37 days
a
b
Fig. 3 | Performance remains high when daily decoder retraining is
shortened (or unsupervised). a, To account for changes in neural activity that
accrue over time, we retrained our handwriting decoder each day before
evaluating it. Here, we simulated offline how decoding performance would
have changed if fewer than the original 50 calibration sentences were used.
Lines show the mean error rate across all data and shaded regions indicate 95%
CIs. b, Copy-typing data from eight sessions were used to assess whether fewer
calibration data are required if sessions occur closer in time. All session pairs
(X, Y) were considered. Decoders were first initialized using training data from
session X and earlier, and then evaluated on session Y under different
retraining methods (no retraining, retraining with limited calibration data, or
unsupervised retraining). Lines show the average raw error rate and shaded
regions indicate 95% CIs.
Table 1 | Mean character and word error rates (with 95% CIs)
for the handwriting BCI across all 5 days
Character error rate
[95% CI]
Word error rate
[95% CI]
Raw online output 5.9% [5.3, 6.5] 25.1% [22.5, 27.4]
Online output+ofline language
model
0.89% [0.61, 1.2] 3.4% [2.5, 4.4]
Ofline bidirectional RNN
+language model
0.17% [0, 0.36] 1.5% [0, 3.2]
‘Raw online output’ is what was decoded online (in real time). ‘Online output+ofline
language model’ was obtained by applying a language model retrospectively to what was
decoded online (to simulate an autocorrect feature). ‘Ofline bidirectional RNN+language
model’ was obtained by retraining a bidirectional (acausal) decoder ofline using all available
data, in addition to applying a language model. Word error rates can be much higher than
character error rates because a word is considered incorrect if any character in that word is
wrong.
Nature | Vol 593 | 13 May 2021 | 253
is correlated across time and neurons (Extended Data Fig.5, Supple-
mentary Note1).
These results suggest that time-varying patterns of movement,
such as handwritten letters, are fundamentally easier to decode
than point-to-point movements. We think this is one—but not neces-
sarily the only—important factor that enabled a handwriting BCI to
go faster than continuous-motion point-and-click BCIs. Other dis-
crete (classification-based) BCIs have also typically used directional
movements with little temporal variety, which may have limited their
accuracy and/or the size of the movement set
24,25
.
More generally, using the principle of maximizing the nearest-
neighbour distance between movements, it should be possible to
optimize a set of movements for ease of classification
26
. We investi-
gated this possibility and designed an alphabet that is theoretically
easier to classify than the Latin alphabet (Extended Data Fig.6). The
optimized alphabet avoids large clusters of redundant letters that are
written similarly (most Latin letters begin with either a downstroke or
a counter-clockwise curl).
Discussion
Locked-in syndrome (paralysis of nearly all voluntary muscles) severely
impairs or prevents communication, and is most frequently caused by
brainstem stroke or late-stage amyotrophic lateral sclerosis (estimated
prevalence of locked-in syndrome: 1 in 100,000
27
). Commonly used BCIs
for restoring communication are either flashing electroencephalogram
(EEG) spellers
18,2832
or intracortical point-and-click BCIs
6,7,33
. EEG spell-
ers based on oddball potentials or motor imagery typically achieve
1–5characters per minute
2832
. EEG spellers that use visually evoked
potentials have achieved speeds of 60characters per minute
18
, but have
notable usability limitations, as they tie up the eyes, are not typically
self-paced and require panels of flashing lights on a screen. Intracorti-
cal BCIs based on 2D cursor movements give the user more freedom
to look around and set their own pace of communication, but have
yet to exceed 40 correct characters per minute in humans
7
. Recently,
speech-decoding BCIs have shown exciting promise for restoring rapid
communication
16,17,34
, but their accuracies and vocabulary sizes require
further improvement to support general-purpose use.
Here, we introduced a new approach for communication BCIs—
decoding a rapid, dexterous motor behaviour in an individual with
tetraplegia—that sets a benchmark for communication rate at 90char-
acters per minute. The real-time system is general (the user can express
any sentence), easy to use (entirely self-paced and the eyes are free to
move) and accurate enough to be useful in the real-world (94.1% raw
accuracy, and greater than 99% accuracy offline with a large-vocabulary
language model). To achieve high performance, we developed decod-
ing methods to work effectively with unlabelled neural sequences in
data-limited regimes. These methods could be applied more gener-
ally to any sequential behaviour that cannot be observed directly (for
example, decoding speech from someone who can no longer speak).
It is important to recognize that the current system is a proof of con-
cept that a high-performance handwriting BCI is possible (in a single
participant); it is not yet a complete, clinically viable system. More
work is needed to demonstrate high performance in additional people,
expand the character set (for example, capital letters), enable text
editing and deletion, and maintain robustness to changes in neural
activity without interrupting the user for decoder retraining. More
broadly, intracortical microelectrode array technology is still maturing,
and requires further demonstrations of longevity, safety and efficacy
before widespread clinical adoption
35,36
. Variability in performance
across participants is also a potential concern (in a previous study, T5
achieved the highest performance of three participants
7
).
Nevertheless, we believe that the future of intracortical BCIs is bright.
Current microelectrode array technology has been shown to retain
functionality for more than 1,000days after implant
37,38
(including here;
see Extended Data Fig.7), and has enabled the highest BCI performance
so far compared to other recording technologies (for example, EEG or
electrocorticography) for restoring communication
7
, arm control
2,5
...
...
Electrodes
Movements
ac ed
Characters
Lines
Time
N1
N2
N2
N1
Time
Distances
Distances
0
1
2
3
1
0
1
2
2
1
0
1
3
2
1
0
0
1.6
1.5
2.2
1.6
0
2.2
1.5
1.5
2.2
0
1.6
2.2
1.5
1.6
0
Time-varying trajectories
Constant trajectories
f
b
Lines
Characters
Mean distance (Hz)
Characters
Lines
0
5
10
15
Characters
Lines
0
10
20
Characters
Lines
0
2
4
Spatial dim.
Characters
Lines
0
2
4
Temporal dim.
P = 1.2 × 10
–7
P = 6.8 × 10
–12
Estimated
true noise
gh
Normalized
time
Nearest-neighbour
distance (Hz)
Classication
accuracy (%)
Noise (V)
0123
0.6
0.8
1.0
Noise (V)
Classication
accuracy (%)
0.6
0.8
1.0
0.2 0.4 0.6 0.8
Fig. 4 | Increased temporal variety can make movements easier to decode.
a, We analysed the spatiotemporal patterns of neural activity corresponding to
16 handwritten characters (1s in duration) versus 16 handwritten straight-line
movements (0.6s in duration). b, Spatiotemporal neural patterns were found
by averaging over all trials for a given movement (after time-warping to align
the trials in time)
11
. Neural activity was resampled to equalize the duration of
each set of movements, resulting in a 192×100 matrix for each movement
(192 electrodes and 100 time steps). c, Pairwise Euclidean distances
between neural patterns were computed for each set, revealing larger
nearest-neighbour distances (but not mean distances) for characters. Each
circle represents a single movement and bar heights show the mean. d, Larger
nearest-neighbour distances made the characters easier to classify than
straight lines. The noise is in units of standard deviations and matches the scale
of the distances in c. e, The spatial dimensionality (dim.) was similar for
characters and straight lines, but the temporal dimensionality was more than
twice as high for characters, suggesting that more temporal variety underlies
the increased nearest-neighbour distances and better classification
performance. Error bars show 95% CIs. Dimensionality was quantified using the
participation ratio. fh, A toy example to give intuition for how increased
temporal dimensionality can make neural trajectories more separable. Four
neural trajectories are depicted (N1 and N2 are two hypothetical neurons, the
activity of which is constrained to a single spatial dimension, the unity
diagonal). Allowing the trajectories to vary in time by adding one bend, which
increases the temporal dimensionality from 1 (f) to 2 (g), enables larger
nearest-neighbour distances and better classification (h).
254 | Nature | Vol 593 | 13 May 2021
Article
and general-purpose computer use
39
. New developments are under
way for implant designs that increase the electrode count by at least
an order of magnitude, which will further improve performance and
longevity
35,36,40,41
. Finally, we envision that a combination of algorithmic
innovations
42–44
and improvements to device stability will continue to
reduce the need for daily decoder retraining. Here, offline analyses
showed the potential promise of more limited, or even unsupervised,
decoder retraining (Fig.3).
Online content
Any methods, additional references, Nature Research reporting sum-
maries, source data, extended data, supplementary information,
acknowledgements, peer review information; details of author contri-
butions and competing interests; and statements of data and code avail-
ability are available at https://doi.org/10.1038/s41586-021-03506-2.
1. Hochberg, L. R. etal. Reach and grasp by people with tetraplegia using a neurally
controlled robotic arm. Nature 485, 372–375 (2012).
2. Collinger, J. L. etal. High-performance neuroprosthetic control by an individual with
tetraplegia. Lancet 381, 557–564 (2013).
3. Alalo, T. etal. Neurophysiology. Decoding motor imagery from the posterior parietal
cortex of a tetraplegic human. Science 348, 906–910 (2015).
4. Bouton, C. E. etal. Restoring cortical control of functional movement in a human with
quadriplegia. Nature 533, 247–250 (2016).
5. Ajiboye, A. B. etal. Restoration of reaching and grasping movements through
brain-controlled muscle stimulation in a person with tetraplegia: a proof-of-concept
demonstration. Lancet 389, 1821–1830 (2017).
6. Jarosiewicz, B. etal. Virtual typing by people with tetraplegia using a self-calibrating
intracortical brain–computer interface. Sci. Transl. Med. 7, 313ra179 (2015).
7. Pandarinath, C. etal. High performance communication by people with paralysis using
an intracortical brain–computer interface. eLife 6, e18554 (2017).
8. Palin, K., Feit, A. M., Kim, S., Kristensson, P. O. & Oulasvirta, A. How do people type on
mobile devices? Observations from a study with 37,000 volunteers. In Proc. 21st
International Conference on Human–Computer Interaction with Mobile Devices and
Services 1–12 (Association for Computing Machinery, 2019).
9. Yousry, T. A. etal. Localization of the motor hand area to a knob on the precentral gyrus. A
new landmark. Brain 120, 141–157 (1997).
10. Willett, F. R. etal. Hand knob area of premotor cortex represents the whole body in a
compositional way. Cell 181, 396–409 (2020).
11. Williams, A. H. etal. Discovering precise temporal patterns in large-scale neural
recordings through robust and interpretable time warping. Neuron 105, 246–259 (2020).
12. Hinton, G. etal. Deep neural networks for acoustic modeling in speech recognition: the
shared views of four research groups. IEEE Signal Process. Mag. 29, 82–97 (2012).
13. Graves, A., Mohamed, A. & Hinton, G. Speech recognition with deep recurrent neural
networks. In 2013 IEEE International Conference on Acoustics, Speech and Signal
Processing 6645–6649 (2013).
14. Xiong, W. etal. The Microsoft 2017 Conversational Speech Recognition System. Preprint
at https://arxiv.org/abs/1708.06073 (2017).
15. He, Y. etal. Streaming end-to-end speech recognition for mobile devices. In 2019 IEEE
International Conference on Acoustics, Speech and Signal Processing 6381–6385 (2019).
16. Anumanchipalli, G. K., Chartier, J. & Chang, E. F. Speech synthesis from neural decoding
of spoken sentences. Nature 568, 493–498 (2019).
17. Makin, J. G., Moses, D. A. & Chang, E. F. Machine translation of cortical activity to text with
an encoder-decoder framework. Nat. Neurosci. 23, 575–582 (2020).
18. Chen, X. etal. High-speed spelling with a noninvasive brain–computer interface. Proc.
Natl Acad. Sci. USA 112, E6058–E6067 (2015).
19. Dickey, A. S., Suminski, A., Amit, Y. & Hatsopoulos, N. G. Single-unit stability using
chronically implanted multielectrode arrays. J. Neurophysiol. 102, 1331–1339 (2009).
20. Eleryan, A. etal. Tracking single units in chronic, large scale, neural recordings for brain
machine interface applications. Front. Neuroeng. 7, 23 (2014).
21. Downey, J. E., Schwed, N., Chase, S. M., Schwartz, A. B. & Collinger, J. L. Intracortical
recording stability in human brain–computer interface users. J. Neural Eng. 15, 046016
(2018).
22. Willett, F. R. etal. Signal-independent noise in intracortical brain–computer interfaces
causes movement time properties inconsistent with Fitts’ law. J. Neural Eng. 14, 026010
(2017).
23. Gao, P. etal. A theory of multineuronal dimensionality, dynamics and measurement.
Preprint at https://doi.org/10.1101/214262 (2017).
24. Musallam, S., Corneil, B. D., Greger, B., Scherberger, H. & Andersen, R. A. Cognitive
control signals for neural prosthetics. Science 305, 258–262 (2004).
25. Santhanam, G., Ryu, S. I., Yu, B. M., Afshar, A. & Shenoy, K. V. A high-performance
brain–computer interface. Nature 442, 195–198 (2006).
26. Cunningham, J. P., Yu, B. M., Gilja, V., Ryu, S. I. & Shenoy, K. V. Toward optimal target
placement for neural prosthetic devices. J. Neurophysiol. 100, 3445–3457 (2008).
27. Pels, E. G. M., Aarnoutse, E. J., Ramsey, N. F. & Vansteensel, M. J. Estimated prevalence of
the target population for brain–computer interface neurotechnology in the Netherlands.
Neurorehabil. Neural Repair 31, 677–685 (2017).
28. Vansteensel, M. J. etal. Fully implanted brain–computer interface in a locked-in patient
with ALS. N. Engl. J. Med. 375, 2060–2066 (2016).
29. Nijboer, F. etal. A P300-based brain–computer interface for people with amyotrophic
lateral sclerosis. Clin. Neurophysiol. 119, 1909–1916 (2008).
30. Townsend, G. etal. A novel P300-based brain–computer interface stimulus
presentation paradigm: moving beyond rows and columns. Clin. Neurophysiol. 121,
1109–1120 (2010).
31. McCane, L. M. etal. P300-based brain–computer interface (BCI) event-related potentials
(ERPs): people with amyotrophic lateral sclerosis (ALS) vs. age-matched controls. Clin.
Neurophysiol. 126, 2124–2131 (2015).
32. Wolpaw, J. R. etal. Independent home use of a brain–computer interface by people with
amyotrophic lateral sclerosis. Neurology 91, e258–e267 (2018).
33. Bacher, D. etal. Neural point-and-click communication by a person with incomplete
locked-in syndrome. Neurorehabil. Neural Repair 29, 462–471 (2015).
34. Mugler, E. M. etal. Direct classiication of all American English phonemes using signals
from functional speech motor cortex. J. Neural Eng. 11, 035015 (2014).
35. Nurmikko, A. Challenges for large-scale cortical interfaces. Neuron 108, 259–269
(2020).
36. Vázquez-Guardado, A., Yang, Y., Bandodkar, A. J. & Rogers, J. A. Recent advances in
neurotechnologies with broad potential for neuroscience research. Nat. Neurosci. 23,
1522–1536 (2020).
37. Simeral, J. D., Kim, S.-P., Black, M. J., Donoghue, J. P. & Hochberg, L. R. Neural control of
cursor trajectory and click by a human with tetraplegia 1000 days after implant of an
intracortical microelectrode array. J. Neural Eng. 8, 025027 (2011).
38. Bullard, A. J., Hutchison, B. C., Lee, J., Chestek, C. A. & Patil, P. G. Estimating risk for future
intracranial, fully implanted, modular neuroprosthetic systems: a systematic review of
hardware complications in clinical deep brain stimulation and experimental human
intracortical arrays. Neuromodulation 23, 411–426 (2020).
39. Nuyujukian, P. etal. Cortical control of a tablet computer by people with paralysis.
PLoS One 13, e0204566 (2018).
40. Musk, E. An integrated brain–machine interface platform with thousands of channels.
J. Med. Internet Res. 21, e16194 (2019).
41. Sahasrabuddhe, K. etal. The Argo: a high channel count recording system for neural
recording invivo. J. Neural Eng. 18, 015002 (2021).
42. Sussillo, D., Stavisky, S. D., Kao, J. C., Ryu, S. I. & Shenoy, K. V. Making brain–machine
interfaces robust to future neural variability. Nat. Commun. 7, 13749 (2016).
43. Dyer, E. L. etal. A cryptography-based approach for movement decoding. Nat. Biomed.
Eng. 1, 967–976 (2017).
44. Degenhart, A. D. etal. Stabilization of a brain–computer interface via the alignment of
low-dimensional spaces of neural activity. Nat. Biomed. Eng. 4, 672–685 (2020).
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional afiliations.
© The Author(s), under exclusive licence to Springer Nature Limited 2021
Reporting summary
Further information on research design is available in theNature
Research Reporting Summary linked to this paper.
Data availability
All neural data needed to reproduce the findings in this study are pub-
licly available at the Dryad repository (https://doi.org/10.5061/dryad.
wh70rxwmv). The dataset contains neural activity recorded during
the attempted handwriting of 1,000 sentences (43,501characters)
over 10.7hours.
Code availability
Code that implements an offline reproduction of the central findings in
this study (high-performance neural decoding with an RNN) is publicly
available on GitHub at https://github.com/fwillett/handwritingBCI.
Acknowledgements We thank participant T5 and his caregivers for their dedicated
contributions to this research, N. Lam, E. Siauciunas and B. Davis for administrative
supportand E. Woodrum for the drawings in Figs.1a,2a. F.R.W. and D.T.A. acknowledge the
support of the Howard Hughes Medical Institute. L.R.H. acknowledges the support of the
Ofice of Research and Development, Rehabilitation R&D Service, US Department of Veterans
Affairs (A2295R, N2864C); the National Institute of Neurological Disorders and Stroke and
BRAIN Initiative (UH2NS095548); and the National Institute on Deafness and Other
Communication Disorders (R01-DC009899, U01-DC017844). K.V.S. and J.M.H. acknowledge
the support of the National Institute on Deafness and Other Communication Disorders
(R01-DC014034, U01-DC017844); the National Institute of Neurological Disorders and Stroke
(UH2-NS095548, U01-NS098968); L. and P. Garlick; S. and B. Reeves; and the Wu Tsai
Neurosciences Institute at Stanford. K.V.S. acknowledges the support of the Simons
Foundation Collaboration on the Global Brain 543045 and the Howard Hughes Medical
Institute (K.V.S. is a Howard Hughes Medical Institute Investigator). The funders had no role in
study design, data collection and interpretation, or the decision to submit the work for
publication.
Author contributions F.R.W. conceived the study, built the real-time decoder, analysed
the data and wrote the manuscript. F.R.W. and D.T.A. collected the data. L.R.H. is the
sponsor-investigator of the multi-site clinical trial. J.M.H. planned and performed T5s array
placement surgery and was responsible for all clinical-trial-related activities at Stanford.
J.M.H. and K.V.S. supervised and guided the study. All authors reviewed and edited the
manuscript.
Competing interests The MGH Translational Research Center has a clinical research support
agreement with Neuralink, Paradromics and Synchron, for which L.R.H. provides consultative
input. J.M.H. is a consultant for Neuralink, and serves on the Medical Advisory Board of Enspire
DBS. K.V.S. consults for Neuralink and CTRL-Labs (part of Facebook Reality Labs) and is on the
scientiic advisory boards of MIND-X, Inscopix and Heal. F.R.W., J.M.H. and K.V.S. are inventors
on patent application US 2021/0064135 A1 (the applicant is Stanford University), which covers
the neural decoding approach taken in this work. All other authors have no competing
interests.
Additional information
Supplementary information The online version contains supplementary material available at
https://doi.org/10.1038/s41586-021-03506-2.
Correspondence and requests for materials should be addressed to F.R.W.
Peer review information Nature thanks Karim Oweiss and the other, anonymous, reviewer(s)
for their contribution to the peer review of this work. Peer reviewer reports are available.
Reprints and permissions information is available at http://www.nature.com/reprints.
Article
Extended Data Fig. 1 | Diagram of the RNN architecture. We used a two-layer
gated recurrent unit (GRU) recurrent neural network architecture to convert
sequences of neural firing rate vectors x
t
(which were temporally smoothed
and binned at 20 ms) into sequences of character probability vectors y
t
and
‘new character’ probability scalars z
t
. The y
t
vectors describe the probability of
each character being written at that moment in time, and the z
t
scalars go high
whenever the RNN detects that T5 is beginning to write any new character. Note
that the top RNN layer runs at a slower frequency than the bottom layer, which
we found improved the speed of training by making it easier to hold
information in memory for long time periods. Thus, the RNN outputs are
updated only once every 100ms. Also, note that we used a day-specific affine
transform to account for day-to-day changes in the neural activity (bottom
row)—this helps the RNN to account for changes in neural tuning caused by
electrode array micromotion or brain plasticity when training data are
combined across multiple days.
Extended Data Fig. 2 | Overview of RNN training methods. a, Diagram of the
session flow for copy-typing and free-typing sessions (each rectangle
corresponds to one block of data). First, single-letter and sentences training
data are collected (blue and red blocks). Next, the RNN is trained using the
newly collected data plus all of the previous days’ data (purple block). Finally,
the RNN is held fixed and evaluated (green blocks). b, Diagram of the data
processing and RNN training process (purple block in a). First, the single-letter
data are time-warped and averaged to create spatiotemporal templates of
neural activity for each character. These templates are used to initialize the
hidden Markov models (HMMs) for sentence labelling. After labelling, the
observed data are cut apart and rearranged into new sequences of characters
to make synthetic sentences. Finally, the synthetic sentences are combined
with the real sentences to train the RNN. c, Diagram of a forced-alignment HMM
used to label the sentence ‘few black taxis drive up major roads on quiet hazy
nights’. The HMM states correspond to the sequence of characters in the
sentence. d, The label quality can be verified with cross-correlation heat maps
made by correlating the single character neural templates with the real data.
The HMM-identified character start times form clear hotspots on the heat
maps. Note that these heat maps are depicted only to qualitatively show label
quality and aren’t used for training (only the character start times are needed to
generate the targets for RNN training). e, To generate new synthetic sentences,
the neural data corresponding to each labelled character in the real data are cut
out of the data stream and put into a snippet library. These snippets are then
pulled from the library at random, stretched or compressed in time by up to
30% (to add more artificial timing variability) and combined into new
sentences.
Article
Extended Data Fig. 3 | The effect of key RNN parameters on performance.
a, Training with synthetic data (left) and artificial white noise added to the
inputs (right) were both essential for high performance. Data are shown from a
grid search over both parameters, and lines show performance at the best value
for the other parameter. Results indicate that both parameters are needed for
high performance, even when the other is at the best value. Using synthetic
data is more important when the size of the dataset is highly limited, as is the
case when training on only a single day of data (blue line). Note that the inputs
given to the RNN were z-scored, so the input white noise is in units of standard
deviations of the input features. b, Artificial noise added to the feature means
(random offsets and slow changes in the baseline firing rate) greatly improves
the ability of the RNN to generalize to new blocks of data that occur later in the
session, but does not help the RNN to generalize to new trials within blocks of
data on which it was already trained. This is because feature means change
slowly over time. For each parameter setting, three separate RNNs were trained
(circles); results show low variability across RNN training runs. c, Training an
RNN with all of the datasets combined improves performance relative to
training on each day separately. Each circle shows the performance on one of
seven days. d, Using separate input layers for each day is better than using a
single layer across all days. e, Improvements in character error rates are
summarized for each parameter. 95% CIs were computed with bootstrap
resampling of single trials (n=10,000). As shown in the table, all parameters
show a statistically significant improvement for at least one condition (CIs do
not intersect zero).
Extended Data Fig. 4 | Changes in neural recordings across days. a, To
visualize how much the neural recordings changed across time, decoded
pen-tip trajectories were plotted for two example letters (m and z) for all 10
days of data (columns), using decoders trained on all other days (rows). Each
session is labelled according to the number of days passed relative to
9December 2019 (day 4). Results show that although patterns of neural activity
clearly change over time, their essential structure is largely conserved (as
decoders trained on past days transfer readily to future days). b, The
correlation (Pearson’s r) between the neural activity patterns of each session
was computed for each pair of sessions and plotted as a function of the number
of days separating each pair. Blue circles show the correlation computed in the
full neural space (all 192 electrodes), whereas red circles show the correlation in
the ‘anchor’ space (top 10 principal components of the earlier session). High
values indicate a high similarity in how characters are neurally encoded across
days. The fact that correlations are higher in the anchor space suggests that the
structure of the neural patterns stays largely the same as it slowly rotates into a
new space, causing shrinkage in the original space but little change in
structure. c, A visualization of how each character’s neural representation
changes over time, as viewed through the top two PCs of the original ‘anchor
space. Each circle represents the neural activity pattern for a single character,
and each x symbol shows that same character on a later day (lines connect
matching characters). Left, a pair of sessions with only two days between them
(day −2 to 0); right, a pair of sessions with 11 days between them (day −2 to 9).
The relative positioning of the neural patterns remains similar across days, but
most conditions shrink noticeably towards the origin. This is consistent with
the neural representations slowly rotating out of the original space into a new
space, and suggests that scaling-up the input features may help a decoder to
transfer more accurately to a future session (by counteracting this shrinkage
effect). d, Similar to Fig.3b, copy-typing data from eight sessions were used to
assess offline whether scaling-up the decoder inputs improves performance
when evaluating the decoder on a future session (when no decoder retraining is
used). All session pairs (X, Y) were considered. Decoders were first initialized
using all data from session X and earlier, then evaluated on session Y under
different input-scaling factors (for example, an input scale of 1.5 means that
input features were scaled up by 50%). Lines indicate the mean raw character
error rate and shaded regions show 95% CIs. Results show that when long
periods of time pass between sessions, input scaling improves performance.
We therefore used an input-scaling factor of 1.5 when assessing decoder
performance in the ‘no retraining’ conditions of Fig.3.
Article
Extended Data Fig. 5 | Effect of correlated noise on the toy model of
temporal dimensionality. a, Example noise vectors and covariance matrix for
temporally correlated noise. On the left, example noise vectors are plotted
(each line depicts a single example). Noise vectors are shown for all 100 time
steps of neuron 1. On the right, the covariance matrix used to generate
temporally correlated noise is plotted (dimensions=200×200). The first
100 time steps describe the noise of neuron 1 and the last 100 time steps
describe the noise of neuron 2. The diagonal band creates noise that is
temporally correlated within each simulated neuron (but the two neurons are
uncorrelated with each other). b, Classification accuracy when using a
maximum likelihood classifier to classify between all four possible trajectories
in the presence of temporally correlated noise. Even in the presence of
temporally correlated noise, the time-varying trajectories are still much easier
to classify. c, Example noise vectors and noise covariance matrix for noise that
is correlated with the signal (that is, noise that is concentrated only in
spatiotemporal dimensions that span the class means). Unlike the temporally
correlated noise, this covariance matrix generates spatiotemporal noise that
has correlations between time steps and neurons. d, Classification accuracy in
the presence of signal-correlated noise. Again, time-varying trajectories are
easier to classify than constant trajectories. See Supplementary Note1 for a
detailed interpretation of this figure.
Extended Data Fig. 6 | An artificial alphabet optimized to maximize neural
decodability. a, Using the principle of maximizing the nearest-neighbour
distance, we optimized for a set of pen trajectories that are theoretically easier
to classify than the Latin alphabet (using standard assumptions of linear neural
tuning to pen-tip velocity). b, For comparison, we also optimized a set of
26 straight lines that maximize the nearest-neighbour distance. c, Pairwise
Euclidean distances between pen-tip trajectories were computed for each set,
revealing a larger nearest-neighbour distance (but not mean distance) for the
optimized alphabet compared to the Latin alphabet. Each circle represents a
single movement and bar heights show the mean. d, Simulated classification
accuracy as a function of the amount of artificial noise added. Results confirm
that the optimized alphabet is indeed easier to classify than the Latin alphabet,
and that the Latin alphabet is much easier to classify than straight lines, even
when the lines have been optimized. e, Distance matrices for the Latin alphabet
and optimized alphabets show the pairwise Euclidean distances between the
pen trajectories. The distance matrices were sorted into seven clusters using
single-linkage hierarchical clustering. The distance matrix for the optimized
alphabet has no apparent structure; by contrast, the Latin alphabet has two
large clusters of similar letters (letters that begin with a counter-clockwise curl,
and letters that begin with a downstroke).
Article
Extended Data Fig. 7 | Example spiking activity recorded from each
microelectrode array. a, Magnetic resonance imaging (MRI)-derived brain
anatomy of participant T5. Microelectrode array locations (blue squares) were
determined by co-registration of postoperative CT images with preoperative
MRI images. b, Example spike waveforms detected during a 10-s time window
are plotted for each electrode (data were recorded on post-implant day 1,218).
Each rectangular panel corresponds to a single electrode and each blue trace is
a single spike waveform (2.1-ms duration). Spiking events were detected using a
−4.5 root mean square (RMS) threshold, thereby excluding almost all
background activity. Electrodes with a mean threshold crossing rate of at least
2Hz were considered to have ‘spiking activity’ and are outlined in red (note that
this is a conservative estimate that is meant to include only spiking activity that
could be from single neurons, as opposed to multiunit ‘hash’). The results show
that many electrodes still record large spiking waveforms that are well above
the noise floor (the yaxis of each panel spans 330μV, whereas the background
activity has an average RMS value of only 6.4μV). On this day, 92 electrodes out
of 192 had a threshold crossing rate of at least 2Hz.
Extended Data Table 1 | Example decoded sentences for one block of copy typing
In the rightmost columns, errors are highlighted in red (extra spaces are denoted with a red square, and omitted letters are indicated with a strikethrough). Note that our language model
substitutes ‘epidermal’ for ‘epidural’, because ‘epidural’ is out of vocabulary. The mean typing rate for this block was 86.47characters per minute and the character error rates were 4.18%
(real-time output) and 1.22% (language model). Sentence prompts were taken from the British National Corpus according to a random selection process (seeSupplementary Methods for
details).
Article
Extended Data Table 2 | Example decoded sentences for one block of free typing
In the rightmost columns, errors are highlighted in red (omitted letters are indicated with a strikethrough). Note that our language model substitutes ‘salt ish’ for ‘sailish’, because ‘sailish’ is out
of vocabulary. The mean typing rate for this block was 73.8characters per minute and the character error rates were 6.82% (real-time output) and 1.14% (language model).

Discussion

Amazing result! > "To our knowledge, 90characters per minute is the highest typ-ing rate that has yet been reported for any type of BCI " This is a nice article illustrating the potential life altering promise of this technology, and profiling some of the users of brain-computer interface technology-> The Man Who Controls Computers With His Mind: https://www.nytimes.com/2022/05/12/magazine/brain-computer-interface.html A character error rate of .89% is extremely low! To give a sense of the team's background, there are neurosurgeons (Jaimie), neurologists, neuroscientists, neuro-engineers, and software engineers that contributed to this research. > "Together, these results suggest that, even years after paralysis, the neural representation of handwriting in the motor cortex is probably strong enough to be useful for a BCI." Recurrent Neural Networks (RNNs) are a class of neural networks that allow previous outputs to be used as inputs while having hidden states. Here is a nice introduction to them from Stanford: https://stanford.edu/~shervine/teaching/cs-230/cheatsheet-recurrent-neural-networks 572 training sentences doesn't sound like a huge amount of training data, however, this is very difficult data to collect and they only have one subject, so seen another way, 31,472 characters is an impressive dataset. The size of their dataset will restrict the methods/machine learning models that they will be able to use, as many machine learning techniques require more data than this to be robustly trained. TSNE *t-distributed stochastic neighbor embedding* (*t-SNE*) is a very popular statistical method for visualizing high-dimensional data by giving each datapoint a location in a lower dimensional map. Background here: [t-distributed stochastic neighbor embedding - Wikipedia](https://en.wikipedia.org/wiki/T-distributed_stochastic_neighbor_embedding) > "Next, we tested whether we could decode complete handwritten sentences in real time, thus enabling an individual with tetraplegia to communicate by attempting to handwrite their intended message." > "It is important to recognize that the current system is a proof of concept that a high-performance handwriting BCI is possible (in a single participant); it is not yet a complete, clinically viable system. More work is needed to demonstrate high performance in additional people, expand the character set (for example, capital letters), enable text editing and deletion, and maintain robustness to changes in neural activity without interrupting the user for decoder retraining. More broadly, intracortical microelectrode array technology is still maturing, and requires further demonstrations of longevity, safety and efficacy before widespread clinical adoption. Variability in performance across participants is also a potential concern (in a previous study, T5 achieved the highest performance of three participants7)" More background on the brain-computer interface: First researched in the 1970s by Jacques Vidal, his 1973 paper was the first to introduce the brain-computer interface to the scientific literature: [Toward Direct Brain-Computer Communication | Annual Review of Biophysics](https://www.annualreviews.org/doi/abs/10.1146/annurev.bb.02.060173.001105) Since then, there has been an array of private and public pursuits of brain-computer interfaces, and recently, these devices have attracted significant attention and funding as the technology becomes more capable. For more on Brain-computer interfaces: [Brain–computer interface - Wikipedia](https://en.wikipedia.org/wiki/Brain%E2%80%93computer_interface)