ArticlePDF AvailableLiterature Review

Heart Rate Variability: Measurement and Clinical Utility

Authors:

Abstract and Figures

Electrocardiographic RR intervals fluctuate cyclically, modulated by ventilation, baroreflexes, and other genetic and environmental factors that are mediated through the autonomic nervous system. Short term electrocardiographic recordings (5 to 15 minutes), made under controlled conditions, e.g., lying supine or standing or tilted upright can elucidate physiologic, pharmacologic, or pathologic changes in autonomic nervous system function. Long-term, usually 24-hour recordings, can be used to assess autonomic nervous responses during normal daily activities in health, disease, and in response to therapeutic interventions, e.g., exercise or drugs. RR interval variability is useful for assessing risk of cardiovascular death or arrhythmic events, especially when combined with other tests, e.g., left ventricular ejection fraction or ventricular arrhythmias.
Content may be subject to copyright.
Heart Rate Variability: Measurement and Clinical Utility
Robert E. Kleiger, M.D.,Phyllis K. Stein, Ph.D.,and J. Thomas Bigger, Jr., M.D.
From the Washington University School of Medicine, St. Louis, MO and Columbia University, New York, NY
Electrocardiographic RR intervals fluctuate cyclically, modulated by ventilation, baroreflexes, and
other genetic and environmental factors that are mediated through the autonomic nervous system.
Short term electrocardiographic recordings (5 to 15 minutes), made under controlled conditions, e.g.,
lying supine or standing or tilted upright can elucidate physiologic, pharmacologic, or pathologic
changes in autonomic nervous system function. Long-term, usually 24-hour recordings, can be used to
assess autonomic nervous responses during normal daily activities in health, disease, and in response
to therapeutic interventions, e.g., exercise or drugs. RR interval variability is useful for assessing risk
of cardiovascular death or arrhythmic events, especially when combined with other tests, e.g., left
ventricular ejection fraction or ventricular arrhythmias. A.N.E. 2005;10(1):88–101
autonomic nervous system
Heart rate responds dynamically to physiologic
perturbations mediated by the autonomic nervous
system via efferent vagal and sympathetic nerve
impulses.1,2Even at rest heart rate fluctuates cycli-
cally. High frequency (HF) cyclic fluctuations are
modulated by ventilation, mediated entirely by
changes in vagal outflow.37Slower fluctuations
occur due to baroreflexes or due to thermoreg-
ulation.37The greatest variation of heart rate
occurs with circadian changes, particularly the dif-
ference between night and day heart rate, mediated
by complex and poorly understood neurohormonal
rhythms.6,8Exercise and emotion also have pro-
found effects on heart rate. Fluctuations in heart
rate reflect autonomic modulation and have prog-
nostic significance in pathological states.945
There are two common settings in which heart
rate variability (HRV) is measured. First, HRV
is assessed under controlled laboratory conditions
with short-term measurements before and after
tilt, drugs, controlled ventilation, or other maneu-
vers selected to challenge the autonomic system.
Secondly, HRV can be determined from 24-hour
electrocardiographic (ECG) recordings made while
subjects perform their usual daily activities.
Twenty-four-hour ECG recordings are particularly
useful for risk stratification in a variety of patholog-
ical entities, but can also be useful for quantifying
autonomic dysfunction.5,12,16,4652
Address for reprints: J. Thomas Bigger, Jr., M.D., Columbia University, 630 West 168th Street, PH 9-103, New York, NY 10032. Fax:
212-305-7141; E-mail: jtb2@columbia.edu
Methods for quantifying HRV are categorized as:
time domain, spectral or frequency domain, geo-
metric, and nonlinear. Baroreflex sensitivity (BRS)
and heart rate turbulence can also be considered
measures of HRV. A short discussion of each will
follow.
TIME DOMAIN MEASURES
OF HEART RATE VARIABILITY
In time domain analysis, the intervals between
adjacent normal R waves (NN intervals) are mea-
sured over the period of recording.53 A variety of
statistical variables can be calculated from the in-
tervals directly and others can be derived from the
differences between intervals (Table 1).5355
SDNN, the standard deviation of all normal RR
(NN) intervals during a 24-hour period, is the most
commonly used time domain measure of HRV. A
major component of SDNN magnitude (approxi-
mately 30–40%) is attributable to day:night differ-
ence in NN intervals. Accurate calculation of SDNN
requires careful editing to exclude ectopic beats,
artifact, and missed beats. Artificially short or long
intervals occurring as a result of these events can ar-
tificially increase SDNN. Most laboratories require
at least 18 hours of usable data to calculate SDNN
in a 24-hour recording.
88
A.N.E. rJanuary 2005 rVol. 10, No. 1 rKleiger, et al. rMeasurement and Clinical Utility r89
Table 1. Time Domain Measures of HRV Calculated over 24 Hours
SDNN Standard deviation of all normal to normal R-R (NN) intervals
SDANN Standard deviation of 5-minute average NN intervals
ASDNN (index) Mean of the standard deviations of all NN intervals for all 5-minute segments in 24 hours
rMSSD Square root of the mean of the squares of successive NN interval differences
NN50 The number of NN intervals differing by >50 ms from the preceding interval
pNN50 The percentage of intervals >50 ms different from preceding interval
Night-day difference Mean night R-R interval minus mean day R-R interval
SDANN, the standard deviation of the 5-minute
average NN intervals, provides a “smoothed out”
version of SDNN, i.e., measures long-term fluc-
tuations.12 SDANN is less subject to editing er-
ror than SDNN because averaging several hundred
NN intervals minimizes the effects of unedited ar-
tifacts, missed beats, and ectopic complexity. As
such, SDANN is also much less affected by abnor-
mal rhythms and may even permit risk stratifica-
tion in atrial fibrillation.
ASDNN (or SDNN index) is the average of the
5-minute standard deviations of NN intervals.53
It reflects the average of changes in NN inter-
vals that occur within 5-minute periods. ASDNN is
significantly correlated with both SDNN and
SDANN, because low and high HRV tend to be
global phenomena, decreasing or increasing all
measures.
The most common variables calculated as differ-
ences between normal R-R intervals are rMSSD,
NN50, and pNN50.56,57 rMSSD is the square root
of the squares of the successive differences be-
tween NN intervals, essentially the average change
in interval between beats.58 NN50 is the absolute
count of differences between successive intervals
>50 ms,17 and pNN50 is the proportion of dif-
ferences >50 ms.12 In the presence of normal si-
nus rhythm and normal AV-nodal function, each of
these measures quantifies parasympathetic modu-
lation of normal R-R intervals driven by ventilation.
All other time domain measures are variants of
those discussed above and correlate highly with one
or more of the previously discussed measures.
SPECTRAL ANALYSIS OF R-R
INTERVALS
Either fast Fourier transformation or autoregres-
sion techniques can be used to quantify cyclic fluc-
tuations of R-R intervals.59 Traditionally, spectral
analysis has been done in short-term laboratory
studies; often standard 5-minute ECG segments are
analyzed. Two peaks are seen in 5-minute R-R in-
terval power spectra, a HF peak between 0.15 and
0.40 Hz and a low frequency (LF) peak between
0.04 and 0.15 Hz (Fig. 1, upper panel).
High frequency power reflects ventilatory mod-
ulation of R-R intervals (respiratory sinus arrhyth-
mia) with the efferent impulses on the cardiac va-
gus nerves, and is abolished by atropine. When the
frequency of ventilation is changed, the center fre-
quency of the HF peak moves with the ventilatory
rate.60,61 The amplitude of the peak, reflecting the
degree to which R-R intervals are affected by ven-
tilation, is similar over normal ventilatory frequen-
cies60,61
Low frequency power is modulated by barore-
flexes with a combination of sympathetic and
parasympathetic efferent nerve traffic to the sinoa-
trial node.1,3,6,37,63,64 Standing or head up tilt typ-
ically causes a modest increase in LF power
and a substantial decrease in HF power.63 At-
ropine almost abolishes the LF peak, and beta
blockade prevents the increase caused by stand-
ing up. Various manipulations of high and LF
power, e.g., normalization or the LF/HF ra-
tio has been applied in an attempt to bet-
ter estimate sympathetic activity. These manip-
ulations are based on a somewhat simplistic
“ying-yang” model of cardiac autonomic function.
Results have been illuminating under some circum-
stances (e.g., tilt table testing) and readily misinter-
preted under others (numerous papers in which in-
creases in the LF/HF ratio due to reductions in HF
power have been interpreted as increased sympa-
thetic activity).
R-R interval power spectra also have been com-
puted using data from 24-hour ECG recordings and
categorized into total power and four mutually ex-
clusive power bands, ultra low, very low, low,
and HF power (Fig. 1, lower panel).9,10 Total and
ultra-low frequency power are best calculated from
a R-R interval periodogram of the entire 24-hour
recording. Instead of computing the 24-hour power
spectrum, spectral analysis often is performed on
90 rA.N.E. rJanuary 2005 rVol. 10, No. 1 rKleiger, et al. rMeasurement and Clinical Utility
Figure 1. R-R interval power spectra. The upper panel
plots log power versus frequency for a 5-minute peri-
odogram and the lower panel plots log power versus fre-
quency for a 24-hour periodogram. In the lower panel,
frequency is plotted on a log scale and the Y axis is
markedly compressed compared with the upper panel.
Note the exponential increase in power as frequency de-
creases below the low frequency band for both graphs.
The two graphs resemble each other, but with much
greater amplitude in the 24-hour plot (lower panel). The
similarity in the graphs is consistent with fractal behavior
for power below the low frequency band.
5-minute segments from 24-hour recordings. HF
and LF power are calculated for each suitable seg-
ment and then averaged. Either method is suit-
able for estimating the average 24-hour HF and LF
power. Unfortunately, commercial Holter systems
sometimes calculate total power in each 5-minute
segment and report its average value over 24 hours.
Because the 5-minute value does not measure fluc-
tuations in R-R intervals with cycles longer than
5 minutes, such as those due to day:night differ-
ences, the 5-minute value is much smaller than to-
tal 24-hour power. The large difference between
5-minute and 24-hour total power can cause confu-
sion; it is the 24-hour value that is more useful for
prognosis (read below).
Most of the power of HRV in a 24-hour record-
ing resides in the frequencies below HF and LF
power which together account for <10% of the to-
tal power over 24 hour. About 12% of power is
accounted for by fluctuations in R-R intervals that
have a period between 20 seconds and 5 minutes
(0.0033–0.04 Hz).10 This spectral band is called very
low frequency (VLF) power. The exact physiologic
mechanism responsible for VLF is a matter of dis-
pute, but, like most other forms of HRV, VLF power
is abolished by atropine, suggesting that it uses a
parasympathetic efferent limb.64,65 Very low fre-
quency power is also reduced by about 20% by
ACE inhibition, suggesting that, at least in part, it
reflects the activity of the renin-aldosterone sys-
tem.66,67 Others have suggested that VLF power
reflects thermoregulation or vasomotor activity.68
Bernardi et al. showed that physical activity can
exert a large effect on VLF power.69 In addition,
sleep-disordered breathing can cause exaggerated
values for VLF power, seen as clear peaks on plots
of the HRV power spectrum during the night.70
The lowest frequency band in the 24-hour R-R
interval power spectrum is ultra low frequency
(ULF) power, which quantifies fluctuations in R-R
intervals with periods between every 5 minutes
and once per 24 hours (ULF <0.003 Hz). Ultra
low frequency power is strongly associated with
SDANN.11
Although the physiologic basis for ULF and VLF
power are far less clear than HF and LF power,
they have proven to be more powerful risk predic-
tors in cardiovascular diseases.10 It is important to
point out that accurate editing, and attention to the
uniformity of beat onset detection, is crucial for 24-
hour spectral analysis. Including nonNN intervals
in the R-R interval time series will substantially
degrade spectral analysis, even more so than for
time domain analysis. Each of the 24-hour spectral
measures has an equivalent time domain variable,
which is highly correlated with it (Table 2) because
both are influenced by the same physiologic inputs
and because of mathematical relationships.11 For
example, SDNN is the square root of the total vari-
ance in normal R-R intervals, whereas total power
is equivalent to total variance. In practice, the
A.N.E. rJanuary 2005 rVol. 10, No. 1 rKleiger, et al. rMeasurement and Clinical Utility r91
Table 2. Highly Correlated Time and Spectral
Measures of HRV
Time Domain Frequency Domain
SDNN Total power
SDANN ULF power
ASDNN VLF power
PNN50, rMSSD HF power
Highly Correlated Time Domain Measures
SDNN SDANN
RMSSD pNN50
correlations between TP and SDNN, ULF power
and SDANN, VLF power, and SDNN index exceed
0.85 and the correlations between ULF power (ap-
proximately 80% of the total power) and TP, SDNN,
SDANN also exceed 0.8. Use of time domain vari-
ables, e.g., SDNN and SDANN rather than the spec-
tral measures for a particular study is a matter of
preference and capability. Because all frequency
domain and some time domain HRV variables have
skewed distributions, the data are usually log trans-
formed for parametric statistical analyses.
GEOMETRIC MEASURES OF R-R
INTERVALS
Heart rate variability triangular index, a geo-
metric measure of HRV, has been used exten-
sively by investigators at St. George’s Hospital in
London.19,37,54 Bedeviled by difficulties in effi-
ciently dealing with ectopic complexes, missed
beats, and noise in analyzing recordings, they cre-
ated histograms of the intervals by sorting them
into 7.8 ms bins. They then fitted a triangle, using
a least squares technique, to the height of each in-
terval. Two measurements were made, the baseline
width of the triangle in milliseconds and the ratio
of the total number of beats divided by the num-
ber of beats in the modal bin. The latter quantity
is called HRV triangular index or just HRV index,
and is essentially the area of the triangle divided by
the area of the modal bin. The calculation of HRV
index minimizes the influence of outlier R-R inter-
vals, i.e., those much longer or shorter than the
usual, thereby substantially reducing the influence
of missed beats, artifact and ectopic complexes.
With accurate editing, HRV index and SDNN are
strongly correlated and both are powerful risk strat-
ifiers after myocardial infarction.19,37,54
NONLINEAR MEASURES OF R-R
INTERVAL FLUCTUATIONS
Although time and frequency domain measures
of HRV quantify HRV on various time scales, non-
linear HRV measures attempt to quantify the struc-
ture or complexity of the R-R interval time series.
For example, a random series of R-R intervals, a
normal series of R-R intervals and a totally peri-
odic series of R-R intervals might have the exact
same SDNN, but their underlying “organization”
would be completely different. A large number of
nonlinear measures of HRV have been studied, but
only a few have shown clear utility in risk stratifi-
cation (Fig. 2). These include the power law slope,
the short- and long-term fractal-scaling exponent,
and SD12, a measure derived from Poincare plots.
Power Law Slope
In normal sinus rhythm, spectral power, mea-
sured over 24 hours, shows a progressive, expo-
nential increase in amplitude with decreasing fre-
quency.71 (Fig. 1b) This relationship can also be
plotted as the log of power (Y axis) versus the log of
frequency (X axis), which transforms the exponen-
tial curve to a line whose slope can be estimated
(Fig. 2, bottom panel). In a log-log plot, the power
law slope between 102and 104Hz is linear with
a negative slope, and reflects the degree to which
the structure of the R-R interval time series is self-
similar over a scale of minutes to hours. Decreased
power law slope has been shown to be a marker for
increased risk of mortality after myocardial infarc-
tion.72
Detrended Fractal Scaling Exponent
This measure, also referred to as α1,iscomputed
from detrended fluctuation analysis (DFA) and is
a measure of the degree to which the R-R interval
pattern is random at one extreme, or correlated at
the other on a scale of 3–11 beats (Fig. 2, middle
panel).73 A totally random R-R interval pattern has
a value for α1of 0.5, whereas a totally correlated
pattern of R-R intervals, i.e., one that is totally pe-
riodic, has a value of 1.5. α1is usually repeatedly
measured within a period of 1000 R-R intervals and
then averaged. Normal values are about 1.05. De-
creased values for α1are strong predictors of out-
come after MI.73,74 Another measure, α2(or DFA2)
can be computed in a similar way on a scale of
92 rA.N.E. rJanuary 2005 rVol. 10, No. 1 rKleiger, et al. rMeasurement and Clinical Utility
Figure 2. Nonlinear Measures of R-R Interval fluctu-
ations. The top panel shows a two-dimensional vector
analysis of a Poincar ´
e plot; the middle panel shows cal-
culation of detrended fluctuation analysis (DFA); and the
bottom panel shows calculation of the power law slope.
The Poincar ´
e plots and DFA analyses are derived from
a 1-hour recording at night in a healthy subject. The
power law slope is derived from a 24-hour recording. Ab-
breviations: SD1, short-term beat-to-beat R-R variability
from the Poincar ´
e plot (width); SD2, long-term beat-to-
beat variability from the Poincar ´
e plot (length); α1,the
short-term fractal scaling exponent for 4–11 beats; α2,
the intermediate-term fractal scaling exponent (11–20
beats), β, power law slope (adapted from Ref.73)
12–20 R-R intervals. α2, however, has not proved
to be especially useful in risk stratification.
The Poincar´
e Plot
The Poincar´
e graph plots each R-R interval as
a function of the next R-R interval (Fig. 2, top
panel) and provides an excellent way to visualize
patterns of R-R intervals.73 Usually, the R-R inter-
val time series is plotted for an entire 24 hours, but
plots of shorter periods, e.g., hourly, can reveal de-
tails obscured in a 24-hour plot that involves about
100,000 points. Poincar´
e plots that reveal abnormal
R-R interval patterns have been characterized as
“complex.” In addition, Poincar´
e plots that reflect
extremely low HRV have also been classified as ab-
normal. SD12 is determined by fitting an ellipse to
the Poincar´
e plot. SD1 is the short axis of this el-
lipse and SD2 is the long axis. SD12 is their ratio. As
the plot becomes more complex, the relative mag-
nitude of SD1 compared to SD2 increases and SD12
becomes larger (Fig. 2, top panel). In addition, if the
plot is small and ball-shaped because of relatively
constant R-R intervals, SD12 also will be large. This
measure has not been used much for risk stratifi-
cation, but has proved useful for detecting editing
problems that significantly influence the calcula-
tion of HRV variables.
Heart Rate Turbulence
Heart rate turbulence is a novel analytic method,
which evaluates the perturbation (shortening then
lengthening) in R-R intervals following premature
ventricular complexes (VPC).75 Two parameters
quantify the response to VPC: turbulence onset
(TO) and turbulence slope (TS). Turbulence onset,
a decrease in the first two normal R-R intervals fol-
lowing a VPC compared with the two normal R-R
intervals just before the VPC, presumably reflects
baroreceptor reflex activity induced by a decreased
stroke volume and blood pressure during the com-
pensatory pause. Normally, the two R-R intervals
after a VPC are shorter than the two normal R-R
intervals immediately preceding the VPC. Turbu-
lence slope quantifies the degree of lengthening of
R-R intervals following the shortening of R-R in-
tervals immediately after a VPC, again reflecting
baroreflex activity.75 It is calculated by determin-
ing the maximum slope of any 5-beat sequence
of normal R-R intervals during the 15–20 R-R
intervals after the VPC. Turbulence onset and
A.N.E. rJanuary 2005 rVol. 10, No. 1 rKleiger, et al. rMeasurement and Clinical Utility r93
turbulence slope are calculated from all single VPC
in a 24-hour recording. Schmidt recommends that
at least 5 VPC be present in a Holter recording, in
order to estimate heart rate turbulence.74 Reduced
heart rate turbulence is strongly associated with in-
creased death rates after MI.7577 Heart rate turbu-
lence will be discussed in detail elsewhere in this
journal.
DIAGNOSTIC USES FOR HEART
RATE VARIABILITY
Analysis of HRV has been used to assess
autonomic function and/or to quantify risk in
a wide variety of both cardiac and noncardiac
disorders. These include, among others, stroke,
multiple sclerosis, end stage renal disease, neonatal
distress, diabetes mellitus, ischemic heart disease,
particularly myocardial infarction, cardiomyopa-
thy, patients awaiting cardiac transplantation,
valvular heart disease, and congestive heart
failure.3,11,12,14,15,1724,2729,32,33,3537,39,43,47,49,50,52
Several authors have reported that HRV analysis
is a more sensitive indicator of autonomic dys-
function in alcoholics and in diabetic subjects than
conventional autonomic tests.7881 Heart rate vari-
ability analysis has also been used to assess the au-
tonomic effects of drugs, including beta-blockers,
calcium blockers, antiarrhythmics, psychotropic
agents, and cardiac glycosides.6567,8291 Drug
effects on HRV can be established with relatively
small numbers of study participants because HRV
measurements are quite stable over the short- and
long-term.92,93
Heart rate variability analysis has had its great-
est cardiologic use in post MI risk stratification
and in assessing risk for arrhythmic events. Wolff
et al. in 1978 first observed that HR variability mea-
sured on admission to the coronary care unit was a
predictor of mortality.94 They calculated the vari-
ance of 30 consecutive R-R intervals taken from
a1minute ECG recording in 176 patients with
acute myocardial infarction. The group of patients
(n =73) with R-R interval variance <32 ms had sig-
nificantly higher hospital mortality than the group
with preserved sinus arrhythmia (n =103). Clin-
ically, patients with low HRV were older, more
likely to have an anterior infarct, and more likely to
have heart failure. It was not clear from this study
whether decreased HRV was an independent pre-
dictor of adverse outcome or if it predicted long-
term risk after myocardial infarction.
The first study that clearly documented the in-
dependent and long-term predictive value of HRV
analysis after myocardial infarction was reported in
1987 by the Multi-Center Post-Infarction Program
(MPIP).28 Eight hundred and eight patients who had
survived acute myocardial infarction had 24-hour
ambulatory electrocardiograms prior to discharge.
Besides Holter variables, which included mean
heart rate, ventricular arrhythmias, and SDNN, pa-
tients were evaluated clinically, had a radionuclide
ejection fraction determined and were evaluated by
a low level exercise test. During a mean follow up
of 31 months, there were 127 deaths (Fig. 3). Forty-
three of these deaths occurred in the group of pa-
tients with SDNN <50 ms (125 patients), approxi-
mately 16% of the total group. Thus, over a third
of these patients died during follow-up and a third
of the deaths occurred in the group with SDNN
<50 ms, establishing a sensitivity and positive pre-
dictive accuracy of about one third (Table 3). The
relative risk of mortality in patients with SDNN
<50 ms versus those with SDNN 50 ms was 2.8.
Reduced SDNN was significantly associated with
low ejection fraction, poor exercise performance,
high New York Heart Association functional class,
and short R-R intervals (Table 4), but the correla-
tions were weak (0.15–0.52). Multivariate analysis
clearly demonstrated that SDNN was an indepen-
dent risk factor for mortality. SDNN also was the
Holter variable with the strongest association with
Figure 3. Kaplan-Meier survival curves from the Multi-
Center Post-Infarction Study demonstrating decreased
survival among patients with SDNN <50 ms (from
Ref.28)
94 rA.N.E. rJanuary 2005 rVol. 10, No. 1 rKleiger, et al. rMeasurement and Clinical Utility
Table 3. SDNN Prediction of Mortality in MPIP
SDNN
<50 ms 50 ms
Number of patients 125 (15.5%) 683 (84.5%)
Deaths, number (%) 43 (34.4%) 84 (12.3%)
Sensitivity 33.9% (43/127)
Specificity 88.0% (599/681)
Positive predictive 34.4% (43/125)
accuracy
False negative rate 12.3% (84/683)
Relative risk =2.8 (34.4%/12.3%)
all-cause mortality, exceeding that of any ventric-
ular arrhythmia measure. Using combinations of
risk variables such as SDNN and ejection fraction
or SDNN and repetitive VPC subgroups of MI pa-
tients could be determined with either very high
(50%) mortality or very low (<2%) 31 months mor-
tality.28,29
Multi-Center Post-Infarction Program data have
been analyzed using other HRV measures. Bigger
et al. evaluated the predictive value of 24-hour
spectral measures.911 Because of the previously
cited physiologic associations of various frequency
bands, it was thought that spectral analysis might
provide mechanistic insight into death and arrhyth-
mias after myocardial infarction. The anticipated
selectivity was not found. All four frequency bands
predicted all-cause and arrhythmic mortality, but
ultra-low frequency power had the strongest associ-
ation with these fatal outcomes. Frequency domain
measures of HRV had similar predictive value for
death of all causes, cardiac death, and arrhythmic
death. The MPIP data also have been analyzed us-
ing heart rate turbulence, which in the MPIP data
Table 4. Correlations of SDNN with Other Variables
in MPIP
rP
Age 0.19 0.0001
Rales in the CCU 0.25 0.0001
Peak BUN 0.15 0.0007
Ejection fraction 0.24 0.0001
Duration of exercise test 0.15 0.0007
Twenty-four-hour average 0.52 0.0001
RR interval
Ln VPC frequency 0.12 0.0004
Ln ventricular paired VPC 0.07 0.04
Ln ventricular runs per hour 0.02 0.57
set is even a stronger risk predictor than conven-
tional time or frequency domain variables.75
Multi-Center Post-Infarction Program was done
in the late 1970s, prior to the institution of much of
what is standard therapy today. Few of the patients
received aspirin or B-blockers and none had reper-
fusion therapy, thrombolytics, angioplasty, or coro-
nary artery bypass graft surgery. Thus, the ques-
tion arose as to whether the MPIP results apply
in the era of reperfusion. Multiple studies since
MPIP have confirmed the power of HRV analysis
in risk stratification post infarction. Some of these
are summarized in Table 5.
Some of the most important of these studies were
performed at St. George’s hospital in London by
Camm, Malik and co-investigators.19,33,34,37,54 Far-
rell reported 68 patients with acute myocardial
infarction who had both baroreceptor sensitivity
and HRV determined before discharge from hos-
pital.95 The latter was measured using HRV trian-
gular index. Both BRS and HRV index were deter-
mined to be good risk stratifiers for survival; BRS
was superior. Subsequently, these investigators ex-
tended their studies to over 400 survivors of my-
ocardial infarction. In all these studies, they uti-
lized HRV index.19,37,54 Approximately 60% of their
patients received thrombolytic therapy or angio-
plasty. Besides HRV index, late potentials, ejection
fraction, clinical variables, and ventricular arrhyth-
mias were measured. In addition, the mechanisms
of death, arrhythmic or nonarrhythmic, and ma-
lignant ventricular rhythms were adjudicated. De-
creased HRV index best predicted both total cardiac
mortality and malignant arrhythmias better than
decreased ejection fraction, abnormal late poten-
tials, or increased frequency of ventricular ectopy
in 24-hour Holter ECG recordings. Furthermore,
combining decreased HRV index with another risk
variable, such as decreased ejection fraction or ab-
normal late potentials, created subgroups of post
MI patients with high risk as well as subgroups
with very low risk of death or malignant ventric-
ular arrhythmias.19
The GISSI study of thrombolytic therapy in acute
myocardial infarction evaluated HRV.52 In GISSI,
all 12,490 patients were treated with streptokinase.
A subset of 567 patients had a valid 24-hour ambu-
latory ECG recording and 52 of them died during
a 1000-day follow-up. Time domain analysis utiliz-
ing SDNN, NN50+, and rMSSD identified high risk
groups comprising 16–18% of the subset with mor-
talities ranging from 20.8 to 24.2% in the high risk
A.N.E. rJanuary 2005 rVol. 10, No. 1 rKleiger, et al. rMeasurement and Clinical Utility r95
Table 5. Representative Post-MPIP Confirmatory Studies of HRV as a Predictor of All-Cause or Cardiac Mortality After MI. Number in Parenthesis
Refers to Reference List. Others Referenced Below Table
Source Number of HRV Measure
(Study Name) Patients (Events) When Obtained Follow-Up HRV Predictors/ Endpoints
Bigger et al. (CAPS)96 N=331 (30 deaths) 24-hour, 1 year after
enrolling in CAPS and
1 week after stopping
meds
3 years ULF, VLF, LF, HF all significant,
univariate predictors of all-cause
mortality. After adjustment for
covariates, VLF was the strongest
predictor
Copie et al.97 N=579, (54 deaths, 42
cardiac, 26 sudden)
24-hour, before
discharge (median 7
days after MI)
>2 years HRV index better predictor than mean
RR interval for sensitivity <40%. For
sensitivity 40% mean R-R interval
and HRV index equal. Mean R-R
interval <700 ms predicted cardiac
death (45% sensitivity, 85%
specificity, 20% PPA) and predicted
all-cause, cardiac and sudden death
better than LVEF
Fei et al.98 N=700 (45 cardiac
deaths, 24 sudden)
24-hour, 5-minute
period, 5–8 days
before discharge
1 year SDNN for 5-minutes had lower PPA
than HRV index, but could preselect
those who require 24-hour Holter
ECG for risk stratification
Huikuri et al. (DIAMOND-MI)74 N=446 with LVEF
0.35, 114 deaths, 75
arrhythmic, 28
nonarrhythmic
24-hour, predischarge,
traditional and
nonlinear HRV
685 ±360 days α1<0.75 RR 3.0, 95% CI 2.5–4.2 for all
cause mortality, independent
predictor after adjustment. Predicted
by arrhythmic and nonarrhythmic
death
La Rovere et al. (ATRAMI)99 N=1284 (44 cardiac
deaths, 5 nonfatal
sudden)
24-hour, <28 days after
MI
21 ±8 months SDNN <70 ms vs SDNN 70 ms
M¨
akikallio et al. (TRACE)100 N=159 with LVEF 35,
72 deaths
24-hour, traditional and
nonlinear
4 years α1<0.85 best univariate predictor of
mortality (RR 3.17, 95% CI
1.96–5.15), PPA 65% and NPA 86%.
Remained significant after
adjustment
Odemuyiwa et al.101 N=433 (first MI), (46
total deaths, 15 sudden
deaths)
24-hour, before
discharge
4 weeks to 5
years
HRV index <20 univariate predictor of
mortality for whole follow-up but
independent predictor of total cardiac
mortality for first 6 months only
Odemuyiwa et al.37 N=385 (44 deaths, 14
sudden)
24-hour, before
discharge
151–1618 days HRV index <39 sensitivity 75%,
specificity 52% compared with LVEF
40% which had specificity of 40%
for all-cause mortality. HRV +LVEF
better specificity for sensitivity <60%
Continued
96 rA.N.E. rJanuary 2005 rVol. 10, No. 1 rKleiger, et al. rMeasurement and Clinical Utility
Table 5. Continued.
Source Number of HRV Measure
(Study Name) Patients (Events) When Obtained Follow-Up HRV Predictors/ Endpoints
Quintana et al.102 N=74 (18 deaths 9
nonfatal MI), 24 normal
controls
24-hour, mean 4 days
after MI
36 ±15 months LnVLF <5.99 independent predictor of
all-cause mortality (RR =1.9) or
mortality/nonfatal infarction (RR =2.2)
Tapanainen et al.103 N=697, 49 deaths 24-hour, 2–7 days after
MI
18.4 ±6.5
months
α1<0.65 most powerful predictor of
mortality RR 5.05, 95% CI 2.87–8.89).
After adjustment, α(1) remained
independently associated with mortality
(RR =3.90, 95% CI 2.03–7.49)
Touboul et al. (GREPI)104 N=471 (26 deaths for 1
year FU, 39 for
long-term FU, 9 sudden)
45% had thrombolysis
24-hour HRV, 10 days
after MI
1 year and long
term (median
31.4 months)
Nighttime AVGNN <750 ms (RR =3.2),
daytime SDNN <100 ms (RR =2.6)
Viashnav et al.105 N=226 (19 cardiac
deaths)
24-hour, mean 83 hours
after MI
Mean 8 months Cox regression not performed Decreased
SDNN, SDANN, ASDNN, LF, HF, LF/HF
among nonsurvivors, but rMSSD and
pNN50 not different
Voss et al.106 N=572 (43 all-cause, 14
sudden arrhythmic, 22
sudden, 34 cardiac, 13
nonfatal VT/VF)
24-hour, 5–8 days after
MI, standard, nonlinear
HRV
2 years For best combination of predictors maximum
specificity at 70% sensitivity where PPA for
endpoints was 16–29% compared with
6–17% for HRV alone
Zabel et al.107 N=250 (30 endpoints) 24-hour HRV, stable,
before discharge
Mean 32 months SDNN significantly higher in event-free (no
VT, resuscitated VF, or death)
Zuanetti et al. (GISSI)52 N=567 males treated
with thrombolysis (52
total deaths, 44 cardiac
deaths)
24-hours at discharge
(median 13 days)
1000 days Independent predictors of all-cause
mortality: NN50 +(RR =3.5), SDNN (RR
=3.0), rMSSD (RR =2.8)
ATRAMI =Autonomic Tone and Reflexes after Myocardial Infarction; CAPS =Cardiac Arrhythmia Pilot Study; DIAMOND =Danish Investigations of Arrhythmia and
Mortality on Dofetilide; GISSI =Grupo Italiano per lo Studio della Sopravvivenza nell-Infarto Miocardico; GREPI =Groupe d’Etude du Pronostic de l’Infarctus du
Myocarde; FU =follow-up; HRV =heart rate variability; LVEF =left ventricular ejection fraction; MI =myocardial infarction; PPA =positive predictive accuracy; RR =
relative risk; TRACE =TRAndolapril Cardiac Evaluation; VF =ventricular fibrillation; VT =ventricular tachycardia.
A.N.E. rJanuary 2005 rVol. 10, No. 1 rKleiger, et al. rMeasurement and Clinical Utility r97
group versus 6.0–6.8% in the low risk group de-
fined by HRV analysis. The relative risk of mortal-
ity was approximately 3.0 for the low HRV groups.
This study not only confirmed the ability of HRV
analysis to risk stratify patients in the modern era
of myocardial infarction treatment but also demon-
strated that HRV analysis alone has limited positive
predictive accuracy.52
THERAPEUTIC IMPLICATIONS OF
HEART RATE VARIABILITY
ANALYSIS
The therapeutic implications to be derived from
HRV analysis after myocardial infarction are un-
clear. Animal models of experimental ischemia and
myocardial infarction show a strong association
between decreased HRV and spontaneous ven-
tricular fibrillation, decreased ventricular fibrilla-
tion threshold and mortality. Furthermore, proce-
dures that increase HRV such as β-adrenergic re-
ceptor blockade, exercise conditioning, low dose
atropine, or scopolamine administration reduce
mortality rates, increase VF threshold, and de-
crease spontaneous, exercise induced, or ischemia
induced ventricular fibrillation in animal mod-
els.15,24,26,40,6567,69,8284,108110 In human studies,
β-blockade increase HRV in both healthy per-
sons and patients who have had myocardial in-
farction,8284 as does scopolamine.109 Type 1C
antiarrhythmic drugs decrease HRV.86,87 Scopo-
lamine in animal models and β-blockers in both
animal models and humans improve survival af-
ter myocardial infarction, whereas 1C antiarrhyth-
mic drugs increase mortality rates; however, how
these effects are related to HRV is not estab-
lished. Thus, diminished HRV is associated with
increased sympathetic and decreased vagal modu-
lation, and these autonomic changes have been as-
sociated with an increase in malignant ventricular
arrhythmias.108,110,111
Perhaps the greatest potential therapeutic use
for HRV analysis in patients after myocardial in-
farction is risk stratification for antiarrhythmic
therapy. The European Myocardial Infarct Amio-
darone Trial (EMIAT) randomized 1486 post MI
patients with ejection fraction 0.40, age 75 to
amiodarone or placebo therapy.112 The 743 pa-
tients on amiodarone had exactly the same mor-
tality as those on placebo with superimposable
Kaplan-Meier mortality curves, but the mecha-
nisms of death was predominantly arrhythmic in
the placebo group and nonarrhythmic in the amio-
darone treated group. However, in those patients
with a low HRV defined as SDNN <50 ms or HRV
index 20 units, there was both a 66% reduction
in arrhythmic death and a 24%, borderline signifi-
cant, decrease in total mortality in the group treated
with amiodarone.113 These results need confirma-
tion but suggest that HRV analysis may be useful in
determining which patients with low ejection frac-
tions after myocardial infarction might most benefit
from ICD implantation.
The Autonomic Tone and Reflexes after Myocar-
dial Infarction (ATRAMI) epidemiological study fol-
lowed 1071 patients after myocardial infarction to
evaluate the predictive value of LVEF, BRS, and
SDNN after myocardial infarction.99 Therapy in
this study was modern, with 63% of the patients
receiving reperfusion therapy. The patients were
low risk because those with heart failure or angina
were excluded. The average follow up was 21 ±
8 months. There were 43 cardiac deaths, 5 patients
had nonfatal cardiac arrest, and 30 patients had
sudden death and/or sustained ventricular tachy-
cardia. LVEF <0.35, SDNN <70 ms, and BRS
<3.0 were all associated with cardiac death, sud-
den death, and nonfatal cardiac arrest. Both BRS
and SDNN predicted mortality during follow-up af-
ter infarction. In patients under age 65, BRS was a
slightly better predictor than SDNN. However, for
those over age 65, SDNN predicted death much bet-
ter than BRS. Baroreflex sensitivity and heart rate
variability had independent predictive value, al-
though they were significantly associated. The sub-
group that had both low BRS and low HRV had an
18% mortality versus <2% for the group with high
values for both variables. It is clear from these data
that depressed HRV remains a statistically power-
ful predictor of death despite modern treatment for
myocardial infarction.114 In ATRAMI, an ejection
fraction <0.35 and decreased BRS or HRV defined,
even in this low risk group of MI patients, a sub-
group with a high, approximately 20%, 2-year mor-
tality.99
LIMITATIONS OF HEART RATE
VARIABILITY AS A RISK
STRATIFIER AFTER MYOCARDIAL
INFARCTION
Although decreased HRV is the most pow-
erful ambulatory ECG predictor of cardiac
mortality and malignant arrhythmias following
98 rA.N.E. rJanuary 2005 rVol. 10, No. 1 rKleiger, et al. rMeasurement and Clinical Utility
myocardial infarction, and in some studies is a
more powerful predictor than ejection fraction,
late potentials, and clinical variables, it has sev-
eral significant limitations. It requires normal si-
nus rhythm and reasonable signal quality. Atrial
fibrillation, sinoatrial dysfunction, and >20% ec-
topic complexes preclude its use. For heart rate tur-
bulence analysis, five or more VPC are needed in
addition to sinus rhythm. To get useful recordings
requires great care in applying electrodes to reduce
artifact and careful editing of the recording to ex-
clude ectopic complexes and artifacts from the cal-
culations. The best predictors require rather long
recording periods in order to include both night-
time and daytime periods. The optimal time after
myocardial infarction to measure HRV is not cer-
tain. There is considerable recovery of HRV in the
3–6 months after myocardial infarction, but recov-
ery values of HRV are, on average, well below
normal age and gender matched healthy individ-
uals.72,115 Although most studies have been per-
formed in the subacute phase of infarction, some
have been performed as late as a year post infarct
and HRV remained significantly associated with
subsequent mortality.96
The best HRV variable to measure is unclear;
conventional time domain, BRS, heart rate turbu-
lence, spectral measures, geometric measures, and
a variety of nonlinear variables reflect different as-
pects of HRV and have all been significantly as-
sociated with outcome without clear, consistent
superiority for any. Moreover, isolated HRV mea-
surements have limited predictive accuracy. As a
univariate predictor, HRV has low sensitivity and
low positive predictive accuracy. Thus, the ther-
apeutic implications of abnormal HRV are uncer-
tain. Yet it has clearly been demonstrated that com-
bining HRV with other risk variables, such as ejec-
tion fraction, BRS, late potentials, exercise testing,
or ventricular arrhythmias can define subgroups
of patients with both very low and very high total
cardiac and arrhythmic mortality after myocardial
infarction. HRV in combination with other vari-
ables, e.g., left ventricular ejection fraction, may be
a very useful clinical tool to better define patients
likely or unlikely to benefit from prophylactic ICD
implantation.
REFERENCES
1. Appel ML, Berger RD, Saul JP, et al. Beat to beat variabil-
ity in cardiovascular variables: Noise or music? J Am Coll
Cardiol 1989;14:1141.
2. Bernardi L, Keller F, Sanders M, et al. Respiratory sinus
arrhythmia in the denervated human heart. J Appl Physiol
1989;67:1447.
3. Billman GE, Dujardin JP. Dynamic changes in cardiac va-
gal tone as measured by time-series analysis. Am J Physiol
1990;258:H896.
4. Conway J, Boon N, Davies C, et al. Neural and humoral
mechanisms involved in blood pressure variability. J Hy-
pertens 1984;2:203.
5. Eckberg DW. Parasympathetic cardiovascular control in
human disease: A critical review of methods and results.
Am J Physiol 1980;241:H581.
6. Furlan R, Guzzetti S, Crivellaro W, et al. Continuous 24-
hour assessment of the neural regulation of systemic ar-
terial pressure and RR variabilities in ambulant subjects.
Circulation 1990;81:537.
7. Grossman P, Van Beek J, Wientjes C. A comparison of
three quantification methods for estimation of respiratory
sinus arrhythmia. Psychophysiology 1990;29:702.
8. Molgaard H, Sorensen KE, Bjerregaard P. Circadian varia-
tion and influence of risk factors on heart rate variability
in healthy subjects. Am J Cardiol 1991;68:777.
9. Bigger JT, Albrecht P, Steinman RC, et al. Comparison of
time- and frequency domain-based measures of cardiac
parasympathetic activity in Holter recordings after my-
ocardial infarction. Am J Cardiol 1989;64:538.
10. Bigger JT, Fleiss J, Steinman RC, et al. Frequency domain
measures of heart period variability and mortality after
myocardial infarction. Circulation 1992;85:164.
11. Bigger JT, Fleiss JL, Steinman RC, et al. Correlations among
time and frequency domain measures of heart period vari-
ability two weeks after acute myocardial infarction. Am J
Cardiol 1992;69:891.
12. Bigger JT, Kleiger RE, Fleiss JL, et al. Components of heart
rate variability measured during healing of acute myocar-
dial infarction. Am J Cardiol 1988;61:208.
13. Billman GE, Schwartz PJ, Stone HL. Baroreceptor reflex
control of heart rate: A predictor of sudden cardiac death.
Circulation 1982;66:874.
14. Casolo G, Balli E, Taddei T, et al. Decreased spontaneous
heart rate variability in congestive heart failure. Am J Car-
diol 1989;64:1162.
15. Corr PB, Yamada KA, Witkowski FX. Mechanisms con-
trolling cardiac autonomic function and their relation to
arrhythmogenesis. In Fozzard HA, Haber E, Jenning RB,
Katz AM (eds.): The Heart and Cardiovascular System.
New York, Raven Press, 1986, p. 1343.
16. Deanfield JE. Holter monitoring in assessment of angina
pectoris. Am J Cardiol 1987;59:18C.
17. Ewing DJ. Heart rate variability: An important new risk
factor in patients following myocardial infarction. Clin
Cardiol 1991;14:683.
18. Ewing DJ, Campbell IW, Clarke BF. The natural history
of diabetic autonomic neuropathy. Quart J Med 1980;49:
95.
19. Farrell TG, Bashir Y, Cripps T, et al. Risk stratification for
arrhythmic events in postinfarction patients based on heart
rate variability, ambulatory electrocardiographic variables
and the signal-averaged electrocardiogram. J Am Coll Car-
diol 1991;18:687.
20. Gaziano EP, Freeman DW. Analysis of heart rate patterns
preceding fetal death. Obstet Gynecol 1977;50:578.
21. Gunderson HJC, Neubauer B. A long term diabetic auto-
nomic nervous abnormality. Reduced variations in resting
heart rate measured by a simple and sensitive method. Di-
abetologia 1977;13:137.
22. Higgins CB, Vatner SF, Eckberg DL, et al. Alterations in the
baroreceptor reflex in conscious dogs with heart failure. J
Clin Invest 1972;251:715.
A.N.E. rJanuary 2005 rVol. 10, No. 1 rKleiger, et al. rMeasurement and Clinical Utility r99
23. Hinkle LE, Carver ST, Plakun A. Slow heart rates and in-
creased risk of cardiac death in middle-aged men. Arch
Intern Med 1972;129:732.
24. Hull SS, Adamson P, Reynolds C, et al. Low sodium diet
increases risk for sudden death in conscious dogs after my-
ocardial infarction. Circulation 1990;82(Suppl):III-334.
25. Imaizumi T, Takeshita A, Makino N, et al. Impaired
baroreflex control of vascular resistance and heart rate in
acute myocardial infarction. Br Heart J 1984;52:418.
26. Kent KM, Smith ER, Redwood DR, et al. Electrical stability
of acutely ischemic myocardium. Influence of heart rate
and vagal stimulation. Circulation 1973;47:291.
27. Kitney RI, Byrne S, Edmonds ME, et al. Heart rate variabil-
ity in the assessment of autonomic diabetic neuropathy.
Automedica 1982;4:155.
28. Kleiger RE, Miller JP, Bigger JT, et al. Decreased heart
rate variability and its association with increased mortality
after acute myocardial infarction. Am J Cardiol 1987;59:
256.
29. Kleiger RE, Miller JP, Krone RJ, et al. The independence of
cycle length variability and exercise testing on predicting
mortality of patients surviving acute myocardial infarction.
Am J Cardiol 1990;65:408.
30. Kolman BS, Verrier RL, Lown B. The effects of vagus nerve
stimulation upon the vulnerability of the canine ventricle:
Role of sympathetic-parasympathetic interactions. Am J
Cardiol 1976;37:1041.
31. Lown B. Sudden cardiac death: The major challenge
confronting contemporary cardiology. Am J Cardiol
1979;43:313.
32. Magid NM, Martin GJ, Kehoe RF, et al. Diminished heart
rate variability in sudden cardiac death. (abstract) Circula-
tion 1985;72:III-241.
33. Malik M, Cripps T, Farrell T, et al. Prognostic value of
heart rate variability after myocardial infarction: A com-
parison of different data-processing methods. Med Biol En-
gin Comput 1989;29:603.
34. Malik M, Farrell TG, Camm AJ. Evaluation of receiver
operator characteristics. Optimal time of day for the as-
sessment of heart rate variability after acute myocardial
infarction. Int J Biomed Comput 1991;29:175.
35. Malliani A, Schwartz PJ, Zanchetti MD. Neural mech-
anisms in life-threatening arrhythmias. Am Heart J
1980;100:705.
36. Niklasson U, Olofsson BO, Bjerle P. Autonomic neuropa-
thy in familial amyloidotic polyneuropathy. A clinical
study based on heart rate variability. Acta Neurol Scand
1989;79:182.
37. Odemuyiwa O, Malik M, Farrell T, et al. Comparison of the
predictive characteristics of heart rate variability index and
left ventricular ejection fraction for all-cause mortality, ar-
rhythmic events and sudden death after acute myocardial
infarction. Am J Cardiol 1991;68:434.
38. Rothschild M, Rothschild A, Pfeifer M. Temporary de-
crease in cardiac parasympathetic tone after acute myocar-
dial infarction. Am J Cardiol 1988;62:637.
39. Schechtman VL, Harper RM, Kluge KA, et al. Heart rate
variation in normal infants and victims of the sudden in-
fant death syndrome. Early Human Dev 1989;19:167.
40. Schwartz PJ. Manipulation of the autonomic nervous sys-
tem in the prevention of cardiac sudden death. In Brugada
PW, Wellens HJJ (eds.): Cardiac Arrhythmias Where Do
We Go from Here? New York, Futura Publishing, 1987,
p. 741.
41. Singer DH, Martin GL, Magid N, et al. Low heart rate
variability and sudden cardiac death. J Electrocardiol
1988;21:S46.
42. Tibblin G, Eriksson CG, Bjuro T, et al. Heart rate and
heart rate variability a risk factor for the development of
ischaemic heart disease (IHD) in the men born in 1913
study—a ten years follow-up. IRCS medical science: Car-
diovascular system. Soc Occupational Med 1975;3:95.
43. Valimaki IA, Nieminen T, Antila KJ, et al. Heart rate vari-
ability and SIDS. Examination of heart rate patterns using
an expert system generator. Ann NY Acad Sci 1988;533:
248.
44. Webb SW, Adgey AAJ, Pantridge JF. Autonomic distur-
bance at onset of acute myocardial infarction. Br Med J
1972;3:89.
45. Zipes DP. Influence of myocardial ischemia and in-
farction on autonomic innervation of heart. Circulation
1990;82:1095.
46. Billman GE, Hoskins RS. Time-series analysis of heart rate
variability during submaximal exercise. Evidence for re-
duced cardiac vagal tone in animals susceptible to ventric-
ular fibrillation. Circulation 1986;80:146.
47. Ewing DJ. Practical bedside investigation of diabetic au-
tonomic failure. In Bannister R (ed.): Autonomic Failure.
Oxford, Oxford University Press, 1983, p. 371.
48. Ewing DJ, Hume L, Campbell IW, et al. Autonomic mech-
anisms in the initial heart rate response to standing. J Appl
Physiol 1980;49:809.
49. Ewing DJ, Martyn CN, Young RJ, et al. The value of car-
diovascular autonomic function tests: 10 years experience
in diabetes. Diabetes Care 1985;8:491.
50. Guzetti S, Iosa D, Pecis M, et al. Effects of sympathetic
activation on heart rate variability in chagas’ patients. J
Auton Nerv Syst 1990;30(Suppl):S79.
51. White CW. Abnormalities in baroreflex control of heart
rate in canine heart failure. Am J Physiol 1981;240:H793.
52. Zuanetti G, Neilson JM, Latini R, et al. On behalf of the
GISSI-2 investigators. Prognostic significance of heart rate
variability in post myocardial infarction patients in the
fibrinolytic era. The GISSI 2 results. Circulation 1996;94:
432.
53. Kleiger RE, Stein P, Bosner M, et al. Time domain mea-
surements of heart rate variability. Cardiology Clinics
1992;10:487.
54. Malik M, Camm J. Heart rate variability. Clin Cardiol
1990;13:570.
55. Modanlou HD, Freeman RK, Braly P. A simple method of
fetal and neonatal heart rate beat-to-beat variability quan-
tification. Am J Obstet Gynecol 1977;129:861.
56. Jennett S, Lamb JF, Travis P. Sudden large and periodic
changes in heart rate in healthy young men after short
periods of exercise. Br Med J 1982;285:1154.
57. Pagani M, Furlan R, Dell’Orto S, et al. Simultaneous
analysis of beat by beat systemic arterial pressure and
heart rate variabilities in ambulatory patients. J Hypertens
1985;3:S83.
58. Von Neumann J, Kent RH, Bellinson HR, et al. The mean
square successive difference. Ann Math Stat 1941;12:153.
59. Task Force of the European Society of Cardiology and
the North American Society of Pacing and Electrophysi-
ology. Heart rate variability. Standards of measurement,
physiological interpretation, and clinical use. Circulation
1996;93:1043.
60. Brown RB, Beightol LA, Koh J, et al. Important influence of
respiration on human R-R interval power spectra is largely
ignored. J Appl Physiol 1993;75:2310.
61. Bloomfield DM, Magnano A, Bigger JT Jr, et al. Com-
parison of spontaneous versus metronome-guided breath-
ing on the assessment of vagal modulation using RR vari-
ability. Am J Physiol: Heart and Circ Physiol 2001;280:
H1145.
62. Eckberg DL. The human respiratory gate. J Physiol Lond
2003;548:339.
63. Bloomfield DM, Kaufman ES, Bigger JT Jr, et al. Passive
100 rA.N.E. rJanuary 2005 rVol. 10, No. 1 rKleiger, et al. rMeasurement and Clinical Utility
head-up tilt and actively standing up produce similar over-
all autonomic balance. Am Heart J 1997;134:136.
64. Taylor JA, Carr DL, Myers CW, et al. Mechanisms un-
derlying very-low-frequency RR-interval oscillations in hu-
mans. Circulation 1998;98:547.
65. Campbell BC, Sturani A, Reid JL. Evidence of parasym-
pathetic activity of the angiotensin converting enzyme
inhibitor, captopril in normotensive man. Clin Sci
1985;68:49.
66. Bonaduce D, Marciano F, Petretta M, et al. Effects of
converting enzyme inhibition on heart period variability
in patients with acute myocardial infarction. Circulation
1994;90:108–113.
67. Lombardi F, Gnocchi-Ruscone T, Montano N, et al. Re-
straining effect of captopril on cardiovascular sympathetic
efferent neural activity. J Hypertension 1989;7(Suppl.
7):S55.
68. Kitney RI. An analysis of the thermoregulatory influences
on heart-rate variability. In Kitney RI, Rompelman O
(eds.): The Study of Heart-Rate Variability. Oxford, Claren-
don Press, 1980, pp. 81–106.
69. Bernardi L, Valle F, Coco M, et al. Physical activity
influences heart rate variability and very-low-frequency
components in Holter electrocardiograms. Cardiovasc Res
1996;32:234.
70. Shiomi T, Guilleminault C, Sasanabe R, et al. Augmented
very low frequency component of heart rate variability
during obstructive sleep apnea. Sleep 1996;19:370.
71. Saul JP, Albrecht P, Berger RD, et al. Analysis of long term
heart rate variability: Methods, l/f scaling and implications.
Comput Cardiol 1987;14:419–422.
72. Bigger JT Jr, Steinman RC, Rolnitzky LM, et al. Power
law behavior of RR-interval variability in healthy middle-
aged persons, patients with recent myocardial infarction,
and patients with heart transplants. Circulation 1996;93:
2142.
73. Makikallio TH, Perkiomaki JS, Huikuri HV. Nonlinear
dynamics of RR intervals. Dynamic Electrocardiography,
Chapter 3, Futura 2004, pp. 22–30.
74. Huikuri HV, M¨
akikallio TH, Peng C-K, et al. for the di-
amond study group. Fractal correlation properties of R-R
interval dynamics and mortality in patients with depressed
left ventricular function after an acute myocardial infarc-
tion. Circulation 2000;101:47.
75. Schmidt G, Malik M, Barthel P, et al. Heart rate tur-
bulence after ventricular premature beats as a predic-
tor of mortality after acute myocardial infarction. Lancet
1999;353:1390.
76. Davies LC, Francis DP, Ponikowski P, et al. Relation of
heart rate and blood pressure turbulence following pre-
mature ventricular complexes to baroreflex sensitivity in
chronic congestive heart failure. Am J Cardiol 2001;87:
737.
77. Bauer A, Schmidt G. Heart rate turbulence in ischemic
heart disease. Dynamic Electrocardiography, Chapter 23,
Futura 2004, pp. 211–213.
78. Malpes SC, Maling TB. Heart rate variability and cardiac
autonomic function in diabetes. Diabetes 1990;41:1177.
79. Rothschild AH, Weinberg CR, Halter JB, et al. Sensitivity
of RR variation and valsalva ratio in assessment of car-
diovascular diabetic autonomic neuropathy. Diabetes Care
1987;10:735.
80. Weise F, Krell D, Brinkhoff N. Acute alcohol inges-
tion reduces heart rate variability. Drug Alcohol Depend
1986;17:89.
81. Malpes SC, Whiteside EA, Maling TJB. Heart rate variabil-
ity and cardiac autonomic function in men with chronic
alcohol dependence. Br Heart J 1991;65:84.
82. Cook JR, Bigger JT, Kleiger RE, et al. Effect of atenolol and
diltiazem on heart rate variability in normal persons. J Am
Coll Cardiol 1991;17:480.
83. Sanderson JE, Chan SK, Yip G, et al. Beta-blockade in heart
failure: A comparison of carvedilol with metoprolol. J Am
Coll Cardiol 1999;34:1522.
84. Goldsmith RL, Bigger JT Jr, Bloomfield DM, et al. Long-
term carvedilol therapy increases parasympathetic ner-
vous system activity in chronic congestive heart failure.
Am J Cardiol 1997;80:1101.
85. Pinar E, Garcia-Alberola A, Llamas C, et al. Effects of ver-
apamil on indexes of heart rate variability after acute my-
ocardial infarction. Am J Cardiol 1998;81:1085.
86. Zuanetti G, Latini R, Neilson JM, et al. Heart rate variabil-
ity in patients with ventricular arrhythmias: Effect of an-
tiarrhythmic drugs. Antiarrhythmic drug evaluation group
(ADEG). J Am Coll Cardiol 1991;17:604.
87. Bigger JT Jr, Fleiss J, Rolnitzky LM, et al. Predicting mor-
tality after myocardial infarction from the response of RR
variability to antiarrhythmic drug therapy. J Am Coll Car-
diol 1994;23:733.
88. Roose SP, Laghrissi-Thode F, Kennedy JS, et al. Compar-
ison of paroxetine and nortriptyline in depressed patients
with ischemic heart disease. J Am Med Assn 1998;279:287.
89. Walsh BT, Greenhill LL, Giardina EG, et al. Effects of de-
sipramine on autonomic input to the heart. J Am Acad
Child Adolesc Psychiatry 1999;38:1186.
90. Kaufman ES, Bosner MS, Bigger JT Jr, et al. Effects of
digoxin and enalapril on heart period variability and re-
sponse to head-up tilt in normal subjects. Am J Cardiol
1993;72:95.
91. Krum H, Bigger JT Jr, Goldsmith RL, et al. The effect of
chronic digoxin therapy on autonomic function in patients
with chronic heart failure. J Am Coll Cardiol 1995;25:289.
92. JKleiger RE, Bigger JT, Bosner MS, et al. Stability over time
of variables measuring heart rate variability in normal sub-
jects. Am J Cardiol 1991;68:626.
93. Bigger JT Jr, Fleiss JL, Rolnitzky LM, et al. Stability over
time of heart period variability in patients with previous
myocardial infarction and ventricular arrhythmias. Am J
Cardiol 1992;69:718.
94. Wolff MW, Varigos GA, Hunt D, et al. Sinus arrhythmia
in acute myocardial infarction. Med J Aust 1978;2:52.
95. Farrell TG, Paul V, Cripps TR, et al. Baroreflex sensitivity
and electrophysiological correlates in patients after acute
myocardial infarction. Circulation 1991;83:945–952.
96. Bigger JR Jr, Fleiss J, Rolnitzky LM, et al. Frequency do-
main measures of heart period variability to assess risk late
after myocardial infarction. J Am Coll Cardiol 1993;21:729.
97. Copie X, Hnatkova K, Staunton A, et al. Predictive power
of increased heart rate versus depressed left ventricular
ejection fraction and heart rate variability for risk strati-
fication after myocardial infarction. Results of a two-year
follow-up study. J Am Coll Cardiol 1996;27:270.
98. Fei L, Copie X, Malik M, et al. Short- and long-term assess-
ment of heart rate variability for risk stratification after
acute myocardial infarction. Am J Cardiol 1996;77:681.
99. La Rovere MT, Bigger JT Jr, Marcus FI, et al, for the
ATRAMI Investigators. Baroreflex sensitivity and heart-
rate variability in prediction of total cardiac mortality after
myocardial infarction. Lancet 1998;351:478.
100. Makikallio TH, Hoiber S, Kober L, et al. Fractal analysis
of heart rate dynamics as a predictor of mortality in pa-
tients with depressed left ventricular function after acute
myocardial infarction. Trace Investigators. TRAndolapril
Cardiac Evaluation. Am J Cardiol 1999;83:836.
101. Odemuyiwa O, Poloniecki J, Malik M, et al. Temporal in-
fluences on the prediction of post infarction mortality by
heart rate variability: A comparison with the left ventric-
ular ejection fraction. Br Heart J 1994;71:521.
A.N.E. rJanuary 2005 rVol. 10, No. 1 rKleiger, et al. rMeasurement and Clinical Utility r101
102. Quintana M, Storck N, Lindblad LE, et al. Heart rate vari-
ability as a means of assessing prognosis after acute my-
ocardial infarction. A 3-year follow-up study. Eur Heart J
1994;18:789.
103. Tapanainen JM, Thomsen PE, Kober L, et al. Fractal anal-
ysis of heart rate variability and mortality after an acute
myocardial infarction. Am J Cardiol 2002;90:347.
104. Touboul P, Andre-Fouet X, Leizorovicz A. Risk stratifi-
cation after myocardial infarction. A reappraisal in the
era of thrombolysis. The Groupe d’Etude du Pronostic de
l’Infarctus du Myocarde (GREPI). Eur Heart J 1997;18:99.
105. Vaishnav S, Stevenson R, Marchant B, et al. Relation
between heart rate variability early after acute myocar-
dial infarction and long-term mortality. Am J Cardiol
1994;73:653.
106. Voss A, Hnatkova K, Wessel N, et al. Multi-parametric
analysis of heart rate variability used for risk stratifica-
tion among survivors of acute myocardial infarction. PACE
1998;21:186.
107. Zabel M, Klingenheben T, Franz MR, et al. Assessment of
QT dispersion for prediction of mortality or arrhythmic
events after myocardial infarction: Results of a prospec-
tive, long-term follow-up study. Circulation 1998;97:
2543.
108. Lown B, Desilva R, Lenson R. Roles of psychologic stress
and autonomic nervous system changes in provocation of
premature complexes. Am J Cardiol 1978;41:979.
109. Vybiral T, Bryg RJ, Maddens ME, et al. Effects of transder-
mal scopolamine on heart rate variability in normal sub-
jects. Am J Cardiol 1990;65:604.
110. Schwartz PJ, Brown AM, Malliani A, et al. (eds.): Neural
Mechanisms in Cardiac Arrhythmias. New York, Raven
Press, 1978, p. 75.
111. Schwartz PJ, Stone HL. The analysis and modulation of au-
tonomic reflexes in the prediction and prevention of sud-
den death. In Zipes DP, Jalife J (eds.): Cardiac Electrophys-
iology and Arrhythmias. Florida, Grune & Stratton, 1985,
p. 167.
112. Julian DG, Camm AJ, Frangin G, et al. Randomized trial
of effect of amiodarone on mortality in patients with left-
ventricular dysfunction after recent myocardial infarction:
Emiat. Lancet 1997;349:667.
113. Malik M, Camm AJ, Janse MJ, et al. Depressed heart rate
variability identifies post infarction patients who might
benefit from prophylactic treatment with amiodarone. J
Am Coll Cardiol 2000;35:1263.
114. La Rovere MT, Pinna GD, Hohnloser SH, et al on behalf of
the Atrami Investigators. Baroreflex sensitivity and heart
rate variability in the identification of patients at risk for
life-threatening arrhythmias. Implications for clinical tri-
als. Circulation 2001;103:2072.
115. Bigger JT Jr, Fleiss JL, Rolnitzky LM, et al. Time course
of recovery of heart period variability after myocardial in-
farction. J Am Coll Cardiol 1991;18:1643.
... Cardiovascular disease is often associated with chronic sympathetic hyperactivity and autonomic cardiac nervous system dysfunction with decreased baroreflex sensitivity (BRS) and blunted heart rate variability (HRV). [3,4] Autonomic cardiovascular dysregulation in PVD is less well described but PVD seems to be associated with increased sympathetic nervous activity . [5] Blunted HRV during wakefulness is considered a marker of cardiovascular autonomic impairment and low HRV has been shown to be associated with adverse cardiovascular outcome. ...
... [5] Blunted HRV during wakefulness is considered a marker of cardiovascular autonomic impairment and low HRV has been shown to be associated with adverse cardiovascular outcome. [3] Short-term HRV can be quantified in the time and in the frequency domain from 5-minute segments of the electrocardiogram (ECG) and provides information about the modulation of heart rate (HR), which is regulated by the interaction of the sympathetic and parasympathetic nervous system. [6] There are no validated standard values and there are differences in HRV based on different ECG lengths (5-min vs. 24h) and awake vs. asleep and during different sleep stages, which must be considered when interpreting. ...
... HRV analysis is a simple method to assess autonomic cardiovascular changes and predict cardiovascular risk and mortality. [3,7] HRV analysis is divided into the time and frequency domains; ...
Article
Full-text available
Background Hypoxia is a trigger for sympathetic activation and autonomic cardiovascular dysfunction. Pulmonary vascular disease (PVD) is associated with hypoxemia, which increases with altitude. The aim was to investigate how exposure of patients with PVD to hypobaric hypoxia at altitude affects autonomic cardiovascular regulation. Methods In a randomised cross-over study, patients with PVD were studied for one day and one night at an altitude of 2500 m (hypobaric hypoxia) and low altitude at 470 m in random order. Outcomes were heart rate variability (HRV) in the time domain and in the frequency domain (low/high frequency, LF/HF) and heart rate (HR) during day and night, and baroreflex sensitivity (BRS). Results In 25 patients with PVD (72% pulmonary arterial hypertension, 28% distal chronic thromboembolic pulmonary hypertension; mean± sd age 60.7±13.6 years), exposure to altitude resulted in significant increases in awake HR by 9.4 bpm (95%CI 6.3 to 12.4, p<0.001) and nocturnal HR by 9.0 bpm (95%CI 6.6 to 11.4, p<0.001) and significant changes in awake and particularly nocturnal HRV indicating decreasing parasympathetic and increasing sympathetic activity (change in daytime LF/HF 1.7 (95%CI 0.6 to 2.8), p=0.004; nocturnal LF/HF 1.9 (95%CI 0.3 to 3.4), p=0.022), and a significant decrease in BRS (−2.4 mmHg ⁻¹ (95%CI - 4.3 to −0.4, p=0.024)). Interpretation Exposure of PVD patients to altitude resulted in a significant change in HRV indicating an increase in sympathetic activity and a decrease in BRS. The relative change in HRV at altitude was more pronounced during sleep.
... For example, high-frequency HRV as well as the time-domain measure root mean square of successive differences (RMSSD) reflect parasympathetic activity and hence regulation of the heart by the central autonomic network [36]. In contrast, total power and the standard deviation of N-N intervals (SDNN) are measures of the total variance in the heartrate signal [37] and thus represent overall autonomic activity. At rest, higher levels of HF-HRV and RMSSD as well as SDNN and total power are desirable with lower levels of these indices typically observed in clinical populations [38]. ...
... SDNN reflects overall autonomic nervous system activity and hence is highly correlated with total power [37]. A summary of the methodological characteristics of all the studies using this outcome measure can be found in Table 6. ...
... TP is the sum of all energy across HRV frequency bands and hence higher TP at rest and therefore serves as an indicator of overall autonomic activity [37]. A summary of the methodological characteristics of studies employing this measure can be found in Table 8. ...
Article
Full-text available
Interoceptive dysfunctions are increasingly implicated in a number of physical and mental health conditions. Accordingly, there is a pertinent need for therapeutic interventions which target interoceptive deficits. Heartrate and heartrate variability biofeedback therapy (HR(V)-BF), interventions which train individuals to regulate their cardiovascular signals and constrain these within optimal parameters through breathing, could enhance the functioning of interoceptive pathways via stimulation of the vagus nerve. Consequently, this narrative systematic review sought to synthesise the current state of the literature with regard to the potential of HR(V)-BF as an interoceptive intervention across behavioural, physiological and neural outcome measures related to interoception. In total, 77 papers were included in this review, with the majority using physiological outcome measures. Overall, findings were mixed with respect to improvements in the outcome measures after HR(V)-BF. However, trends suggested that effects on measures related to interoception were stronger when resonance frequency breathing and an intense treatment protocol were employed. Based on these findings, we propose a three-stage model by which HR(V)-BF may improve interoception which draws upon principles of interoceptive inference and predictive coding. Furthermore, we provide specific directions for future research, which will serve to advance the current knowledge state.
... As per the literature, the HRV parameters that shows a decrease in value during mental stressful intervention is PNS Index [38], SDNN [34,39], SDHR and MaxHR [35], DiffHR [6], RMSSD [27,39,40], NN50 and pNN50 [34,41,42], HTI [3,34], TINN [34,43], DC and DCmod [44,45], HFap and HFrp [6,[46][47][48][49], SD1 [50,51], DFAα1 [52,53], D2 [53,54], MeanLL [55,56], MaxLL [56], REC [55,56], DET [56], and ShEn [55][56][57]. ...
... ms) of RMSSD computed is comparable to the nominal values of 19 ms-75 ms, as reported in the literature [70]. RMSSD is known to correlate with HF in the FD parameters of HRV (see also Fig. 3(a) and (b)) [40]. This is also evident from the statistical significance of HF. ...
... This finding may be because RMSSD is comparatively less influenced by respiration when compared to HF band parameters [114]. In the literature, a decrease in RMSSD may indicate a decrease in vagally mediated changes reflected in HRV [39,40]. However, in our results, the enhancement in the RMSSD was primarily due to the heightened amplitude of the cardiac coherence (RSA) due to 0.1 Hz respiration rate, and it may not directly indicate the vagal dominance. ...
... This prior work is based on HRV measures from ECG recordings taken over several minutes in clinical settings or 24 h recordings taken in a freeliving environment. Longer ECG recordings might be more accurate in representing an individual's HRV, particularly in response to the environment [7,15]. To date, no studies have evaluated associations between HRV from long-term ECG recordings and cognitive function. ...
... Previous works using short-term measurements of HRV suggest that higher values of certain indices are associated with better cognition, but other studies report null results [9,12,40]. The non-significant findings in our study might be due to the measurement of free-living HRV over a longer ECG monitoring period of 14 days which more accurately represents an individual's HRV compared with shorter ECG recordings of several minutes to 24 h [15]. Further, our study focused on only two time-domain measures that are relatively constant over time [7,41]. ...
Article
Full-text available
Low physical activity (PA) measured by accelerometers and low heart rate variability (HRV) measured from short-term ECG recordings are associated with worse cognitive function. Wearable long-term ECG monitors are now widely used, and some devices also include an accelerometer. The objective of this study was to evaluate whether PA or HRV measured from long-term ECG monitors was associated with cognitive function among older adults. A total of 1590 ARIC participants had free-living PA and HRV measured over 14 days using the Zio® XT Patch [aged 72–94 years, 58% female, 32% Black]. Cognitive function was measured by cognitive factor scores and adjudicated dementia or mild cognitive impairment (MCI) status. Adjusted linear or multinomial regression models examined whether higher PA or higher HRV was cross-sectionally associated with higher factor scores or lower odds of MCI/dementia. Each 1-unit increase in the total amount of PA was associated with higher global cognition (β = 0.30, 95% CI: 0.16–0.44) and executive function scores (β = 0.38, 95% CI: 0.22–0.53) and lower odds of MCI (OR = 0.38, 95% CI: 0.22–0.67) or dementia (OR = 0.25, 95% CI: 0.08–0.74). HRV (i.e., SDNN and rMSSD) was not associated with cognitive function. More research is needed to define the role of wearable ECG monitors as a tool for digital phenotyping of dementia.
... This study provides a comprehensive assessment of vagal and sympathetic modulations, indicating the regulation of the autonomic nervous system, by recording several HRV values, such as the RMSSD, SDNN, SI, and Alpha 1. The short-term index Alpha 1 of the trend-adjusted fluctuation analysis is, next to the others, an established parameter that is prognostically relevant for characterizing the regulatory capacity of the ANS [30]. ...
Article
Full-text available
(1) Background: Remote ischemic preconditioning (RIPC) is an intervention involving the application of brief episodes of ischemia and reperfusion to distant tissues to activate protective pathways in the heart. There is evidence suggesting the involvement of the autonomic nervous system (ANS) in RIPC-induced cardioprotection. This study aimed to investigate the immediate effects of RIPC on the ANS using a randomized controlled trial. (2) Methods: From March 2018 to November 2018, we conducted a single-blinded randomized controlled study involving 51 healthy volunteers (29 female, 24.9 [23.8, 26.4] years). Participants were placed in a supine position and heart rate variability was measured over 260 consecutive beats before they were randomized into either the intervention or the SHAM group. The intervention group underwent an RIPC protocol (3 cycles of 5 min of 200 mmHg ischemia followed by 5 min reperfusion) at the upper thigh. The SHAM group followed the same protocol but on the right upper arm, with just 40 mmHg of pressure inflation, resulting in no ischemic stimulus. Heart rate variability measures were reassessed afterward. (3) Results: The intervention group showed a significant increase in RMSSD, the possible marker of the parasympathetic nervous system (IG: 14.5 [5.4, 27.5] ms vs. CG: 7.0 [−4.3, 23.1 ms], p = 0.027), as well as a significant improvement in Alpha 1 levels compared to the control group (IG: −0.1 [−0.2, 0.1] vs. CG: 0.0 [−0.1, 0.2], p = 0.001). (4) Conclusions: Our results hint that RIPC increases the RMSSD and Alpha 1 parameters showing possible immediate parasympathetic modulations. RIPC could be favorable in promoting cardioprotective or/and cardiovascular effects by ameliorating ANS modulations.
... El Holter es un estudio que registra la actividad cardiaca a través de trazados electrocardiográficos guardados electrónicamente en un dispositivo adosado al cuerpo, tiene baterías y electrodos adheridos en el tórax y abdomen superior (15,16,12,10) . ...
Article
Full-text available
El estudio de la regularidad de la Frecuencia Cardiaca, a través del Holter de 24 horas se hace desde la década de los años 60 y es bastante efectivo. Sin embargo, desde los años noventa comenzaron a efectuarse estudios cortos de Holter en pacientes sospechados de tener fallas autonómicas de control de la frecuencia cardiaca, especialmente en pacientes con comorbilidades tales como Hipertensión, Diabetes Mellitus, Aterosclerosis etc. De aquí la importancia de realizar un test de Holter de diez minutos, divididos en dos tiempos de 5 minutos, primero en decúbito dorsal y luego en bipedestación, especialmente en pacientes de más de cincuenta años o con comorbilidades presentes. Los resultados se presentan luego en gráficos de Poincare, que incluye el programa operativo del dispositivo, que permite un vistazo de la elipse con sus dos ejes, que representan las acciones simpáticas y parasimpáticas sobre la frecuencia cardiaca. Una variabilidad anormal de la frecuencia cardiaca debe ser luego estudiada más profundamente a fin de reafirmar el diagnóstico y ulteriores pasos en el tratamiento.
... Dr. Dickinson et al. accurately emphasize the importance of controlled conditions in HRV assessment, as also outlined by Kleiger et al. [1] Accordingly, our study was meticulously designed to adhere to stringent protocols, thereby mitigating the influence of variable conditions on HRV readings [2]. Furthermore, we acknowledge that individual differences such as genetic, lifestyle, and environmental factors, can significantly influence HRV measurements. ...
... Some indices of HRV, such as the root mean square of successive differences (RMSSD), are considered a non-invasive, surrogate measure of vagal tone in the heart ). Functionally, high vagally-mediated HRV (vmHRV) is generally considered to reflect a positive health state or reduced health risk, indicating a high cardiac control (Kleiger et al., 2005;Malik, 1996;Porges, 2007;Thayer & Sternberg, 2006). Moreover, vmHRV is stronger than other biomarkers related to inflammation, blood glucose, and blood lipids to predict subjective health indices which have consistently been found to predict mortality, morbidity, and other health outcomes (Benyamini & Idler, 1999;Jarczok et al., 2015;Pinquart, 2001). ...
Preprint
The neurovisceral integration model proposes that information flows bidirectionally between the brain and the heart via the vagus nerves and vagally-mediated heart rate variability (vmHRV) can be used to index heart-brain interaction. Recent research has shown that voluntary reduction of breathing rate (slow-paced breathing, SPB) can enhance cardiac vagal control. Additionally, prefrontal transcranial direct current stimulation (tDCS) can modulate the excitability of the prefrontal region and influence the vagus nerve. However, fundamental research on the combination of SPB and prefrontal tDCS to increase vmHRV and other physiological indices of the autonomic nervous system is scarce. Therefore, 200 healthy participants were assigned to four experimental groups. Each group received either 20 min of active or sham tDCS combined with 5.5 breath per minute (BPM) or 15 BPM breathing. Regardless of the tDCS condition, the SPB group showed a significant increase in vmHRV over 20 minutes, suggesting an increase in parasympathetic activity. In addition, a significant decrease in HR at the first and second 5-minute epochs of the intervention. Regardless of breathing condition, the active tDCS group exhibited higher HR at the fourth 5-minute epoch of the intervention compared to the sham tDCS group, suggesting more sympathetic arousal. However, there was no combined effect on vmHRV, HR, skin conductance, or blood pressure. SPB is a robust technique for increasing vmHRV, whereas prefrontal tDCS may produce effects that counteract those of SPB. More research is necessary to test whether and how top-down and bottom-up approaches can be combined to improve vagal control.
Article
Full-text available
This Committee Report provides methodological, interpretive, and reporting guidance for researchers who use measures of heart rate (HR) and heart rate variability (HRV) in psychophysiological research. We provide brief summaries of best practices in measuring HR and HRV via electrocardiographic and photoplethysmographic signals in laboratory, field (ambulatory), and brain‐imaging contexts to address research questions incorporating measures of HR and HRV. The Report emphasizes evidence for the strengths and weaknesses of different recording and derivation methods for measures of HR and HRV. Along with this guidance, the Report reviews what is known about the origin of the heartbeat and its neural control, including factors that produce and influence HRV metrics. The Report concludes with checklists to guide authors in study design and analysis considerations, as well as guidance on the reporting of key methodological details and characteristics of the samples under study. It is expected that rigorous and transparent recording and reporting of HR and HRV measures will strengthen inferences across the many applications of these metrics in psychophysiology. The prior Committee Reports on HR and HRV are several decades old. Since their appearance, technologies for human cardiac and vascular monitoring in laboratory and daily life (i.e., ambulatory) contexts have greatly expanded. This Committee Report was prepared for the Society for Psychophysiological Research to provide updated methodological and interpretive guidance, as well as to summarize best practices for reporting HR and HRV studies in humans.
Article
This report deals with the often neglected relationship between heart rate and heart rate variability and the later development of death in IHD and death in general. In 1963 a prospective population study of 855 randomly selected men, all born in 1913 and living in Goteborg, Sweden, was started. 24 Subjects were excluded due to failure to measure 10 beats on the ECG, fibrillation, etc. Heart rate (HR) was calculated from a 10 beat ECG during rest. The length of each RR interval in this 10 beat ECG was measured in ms and plotted beat by beat in a diagram. A line was drawn between the beats and the length of this line was chosen as an indicator of the degree of heart rate variability (HRV); the shorter the line the less heart rate variability and vice versa. This index is analogous to the undulation index used to characterize variability in blood pressure. In order to take HR into consideration, the HRV was calculated as SD units within different HR levels. Negative figures were avoided by adding the units. The Kolmogorov Swirnov one sample test was used due to the small numbers expected. There is a relationship between lack of HRV and IHD death. Surviving myocardial infarction patients show no such relationship. A strong, but nonlinear, relationship was observed between death in other causes and HRV, as well as a surprisingly strong relationship between total death and high HR. No relation between IHD death and HR per se was found.
Article
THE AUTHORS HAVE MEASURED RESPIRATORY SINUS ARRHYTHMIA (RSA) IN NORMAL SUBJECTS AND NEUROPATHS OVER A RANGE OF STEADY STATE BREATHING, ALTERING BOTH TIDAL VOLUME AND FREQUENCY OF BREATHING. THE PRIMARY PURPOSE OF THE STUDY WAS TO QUANTIFY THE WAY IN WHICH HRV (HEART RATE VARIABILITY) CHANGES IN BOTH GROUPS OF SUBJECTS FOR DIFFERENT BREATHING STATES. THE RESULTS OF THESE INVESTIGATIONS CONFIRM THE REDUCTION IN HRV FOUND IN THE PRESENCE OF DIABETIC AUTONOMIC NEUROPATHY. THE REDUCTION IN VARIABILITY NECESSITATES A HIGHER RATE OF DATA SAMPLING THAN REQUIRED FOR NORMAL SUBJECTS IF FREQUENCY SPECTRA ARE TO BE OBTAINED. THE RESULTS WOULD INDICATE THAT ANY INVESTIGATIONS WHERE THE SAMPLING WAS MADE AT LESS THAN 200 HZARE LIKELY TO BE IN ERROR. USE OF THE INTERBEAT FREQUENCY METHOD HAS ADVANTAGES IN TERMS OF COMPUTATION TIME AND SHOULDENABLE THESE STUDIES TO BE CARRIED OUT ON A MICROPROCESSOR-BASED DATA ANALYSIS SYSTEM WHICH WOULD BE SUITABLE FOR USE IN A CLINICAL ENVIRONMENT. THE SHIFT IN THE PEAK OF THE HRV FREQUENCY RESPONSE WHICH OCCURS IN NEUROPATHY APPEARS TO BE DUE TO AN INCREASE IN THE LOOP TIME DELAY OF THE REFLEX, AND MAY PROVIDE THE BASIS FOR A NONINVASIVE CLINICAL TEST TO ASSESS THE PROGRESS OF THE DISEASE.
Article
To determine the short- and long-term reproducibility of heart rate variability measurements in type I diabetics, we examined 23 patients over 1 year. Using 24-hour ambulatory recordings, we demonstrated that heart rate variability is abnormal and reproducible for this group of patients. Little variation of all heart rate variability measurements, especially those reflecting parasympathetic activity, occurred during the study period. We also noted a relationship between heart rate variability and elevated serum cholesterol levels, and differences in diurnal rhythm.