Skip to main content
Erschienen in: Sleep and Breathing 2/2021

Open Access 22.07.2020 | Neurology • Original Article

Effects of central apneas on sympathovagal balance and hemodynamics at night: impact of underlying systolic heart failure

verfasst von: Jens Spiesshoefer, Nora Hegerfeld, Malte Frank Gerdes, Sören Klemm, Martha Gorbachevski, Robert Radke, Izabela Tuleta, Claudio Passino, Xiaoyi Jiang, Paolo Sciarrone, Winfried Randerath, Michael Dreher, Matthias Boentert, Alberto Giannoni

Erschienen in: Sleep and Breathing | Ausgabe 2/2021

Abstract

Background

Increased sympathetic drive is the key determinant of systolic heart failure progression, being associated with worse functional status, arrhythmias, and increased mortality. Central sleep apnea is highly prevalent in systolic heart failure, and its effects on sympathovagal balance (SVB) and hemodynamics might depend on relative phase duration and background pathophysiology.

Objective

This study compared the effects of central apneas in patients with and without systolic heart failure on SVB and hemodynamics during sleep.

Methods

During polysomnography, measures of SVB (heart rate and diastolic blood pressure variability) were non-invasively recorded and analyzed along with baroreceptor reflex sensitivity and hemodynamic parameters (stroke volume index, cardiac index, total peripheral resistance index). Data analysis focused on stable non-rapid eye movement N2 sleep, comparing normal breathing with central sleep apnea in subjects with and without systolic heart failure.

Results

Ten patients were enrolled per group. In heart failure patients, central apneas had neutral effects on SVB (all p > 0.05 for the high, low, and very low frequency components of heart rate and diastolic blood pressure variability). Patients without heart failure showed an increase in very low and low frequency components of diastolic blood pressure variability in response to central apneas (63 ± 18 vs. 39 ± 9%; p = 0.001, 43 ± 12 vs. 31 ± 15%; p = 0.002). In all patients, central apneas had neutral hemodynamic effects when analyzed over a period of 10 min, but had significant acute hemodynamic effects.

Conclusion

Effects of central apneas on SVB during sleep depend on underlying systolic heart failure, with neutral effects in heart failure and increased sympathetic drive in idiopathic central apneas.
Hinweise

Electronic supplementary material

The online version of this article (https://​doi.​org/​10.​1007/​s11325-020-02144-8) contains supplementary material, which is available to authorized users.
First authors J.S. and N.H. contributed equally to this work.
Senior authors M.B. and A.G. contributed equally to this work.
Comments
Spiesshoefer et al. describe different autonomic changes during central apneas in patients with systolic heart failure compared to central apneas in patients with idiopathic central apneas. These findings are interesting and create the basis for future investigations. Further detailed description of central respiratory events and autonomic challenge testing may provide further information to assess, whether differences in central respiratory events or differences in autonomic responses to comparable central respiratory events contribute to the observation described in this manuscript. Additionally, further sleep and autonomic intervention studies are required to finally show, whether central sleep apneas in heart failure patients are really modifiable risk factors or a marker of underlying disease and autonomic imbalance.
Dominik Linz
Mainz, Germany

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Introduction

Patients with systolic heart failure (HF) often present with an unstable respiratory pattern characterized by alternating phases of hyperventilation and central apneas (CA), also named Cheyne-Stokes respiration (CSR) [13]. CSR is mainly caused by increased chemosensitivity to either hypoxia or hypercapnia [4] and delayed chemoresponse due to the prolonged circulatory time that directly results from decreased cardiac output [5]. While the former seems to influence CSR severity, the latter seems to rather influence CSR cycle duration, and the duration of the hyperventilation phase in particular [5]. It has been hypothesized that CSR alters sympathovagal balance (SVB) in favor of increased sympathetic drive and thus has detrimental effects in HF when left untreated [3, 6]. This may be mainly due to an altered pulmonary stretch reflex and blunted vagal stimulation [6, 7], but also due to chemoreflex-mediated sympathetic overactivity in response to apnea-related hypoxemia and hypercapnia [7, 8]. SVB is known to play a key role in determining HF progression because medications that decrease sympathetic drive have been shown to have favorable effects on outcome [9, 10]. For this reason, CA has been considered a potential therapeutic target in HF and has been treated using a variety of options, including oxygen therapy [11], carbon dioxide (CO2) rebreathing [11, 12], acetazolamide [1315], and nocturnal mask-based therapies such as continuous positive airway pressure [16] or adaptive servo-ventilation [17, 18].
The results of the Treatment of Sleep-Disordered Breathing with Predominant Central Sleep Apnea by Adaptive Servo Ventilation in Patients with Heart Failure (SERVE-HF trial) [19] have challenged this hypothesis, because patients with systolic HF (left ventricular ejection fraction [LVEF] ≤ 45%) and CSR treated with adaptive servo-ventilation unexpectedly showed increased mortality compared with controls who received optimal medical therapy alone [20, 21]. This finding has led to controversy about the effect of CSR in HF as potentially compensatory rather than harmful, at least in a subset of patients [20, 21]. Potential favorable effects of CSR are thought to include an increase in venous return and cardiac output, and positive effects of hypocapnia on SVB during the hyperventilation phase [22, 23].
The net effect of CSR, which is characterized by alternating cycles of hyperventilation and CA may be influenced by several factors. These include the relative duration of the hyperventilation and CA phases, as well as background cardiac hemodynamics and/or feedback resetting. CA is also seen, albeit less frequently, in subjects in whom no overt cardiac or neurological cause can be identified, where it is called idiopathic central apnea (ICA) [24]. In patients with ICA, apneic events are usually associated with increased sensitivity of the chemoreflex and are characterized by shorter cycle duration resulting from normal cardiac output and circulation time [25]. ICA may also potentially alter SVB, either directly or indirectly (by altering sleep quality), but this (along with its potential hemodynamic effects) has not yet been investigated.
This study investigated the immediate effects of CA on SVB and hemodynamic parameters during stage 2 non-rapid eye movement (NREM) sleep (N2) in patients with HF-related CA and ICA. This approach allows evaluation of whether the effects of CA on both the autonomous nervous system and the cardiovascular system during sleep differ between patients with and without systolic HF. Thus, it would theoretically be possible to identify a mechanism by which CA potentially exerts beneficial effects in HF patients.

Materials and methods

Study design and study population

Patients with CA (apnea-hypopnea index [AHI] ≥ 15/h with > 50% central events and a central AHI of ≥ 10/h [i.e., SERVE-HF-like inclusion criteria]) due to systolic HF (LVEF < 50% [19], N-terminal pro B-type natriuretic peptide [NT-proBNP] level ≥ 125 ng/L, but by inclusion criterion in sinus rhythm to have valid data on HRV and dBPV) and heart-healthy subjects with ICA (same CA criteria as HF patients, but with LVEF ≥ 55%, sinus rhythm, and serum NT-proBNP < 125 ng/L) from the University Hospital of Muenster were consecutively recruited from April 2018 to February 2019. Diagnosis of CA had previously been established by diagnostic polysomnography (PSG). The study protocol conforms to the 1975 Declaration of Helsinki and was approved by the Institution’s human research committee (Ethikkommission der Ärztekammer Westfalen-Lippe und der Medizinischen Fakultät der Westfälischen Wilhelms-Universität Münster, AZ. 2017-188-f-S). All participants gave written informed consent to participate in the study, and the project had been registered under the German Clinical Trials Registry (drks.de identifier: DRKS00013883).

Assessments

All study participants underwent standard 2-dimensional Doppler echocardiography (LOGIQ S8-XD clear, GE Healthcare, London, UK), standard lung function tests (Vitalograph ® 3000, Vitalograph, Hamburg, Germany) performed according to current recommendations, and measurement of serum NT-proBNP [26]. Furthermore, study participants underwent non-invasive hemodynamic and respiratory monitoring during sleep, as detailed below.

Non-invasive hemodynamic and respiratory monitoring

Parameters reflecting SVB and hemodynamic parameters were measured during sleep using a non-invasive monitor device (Task Force Monitor®, CNSystems, Graz, Austria) as previously validated (Fig. 1a) [2729]. For the assessment of SVB, heart rate variability (HRV) and diastolic blood pressure variability (dBPV) were analyzed using a 3-lead electrocardiogram (sampling rate 1000 Hz) and a continuous non-invasive arterial blood pressure signal (CNAP® technology) with a sampling rate of 100 Hz. HRV and dBPV were computed by frequency domain analysis of continuous recordings of RR intervals (for HRV) or blood pressure changes (for dBPV). HRV (expressed as ms2) and dBPV (expressed as mmHg2) were presented as the high frequency component (HF; 0.15–0.4 Hz), low frequency component (LF; 0.04–0.15 Hz), their relative ratio (LF/HF), and the very low frequency component (VLF; 0.0–0.04 Hz) [27, 30, 31]. Given that total power spectra of HRV and dBPV are significantly reduced in patients with HF compared with heart-healthy individuals, LF, HF, and VLF data for both HRV and dBPV were reported as normalized unit (n.u.) for total power spectra [32].
Baroreceptor reflex sensitivity (BRS) was measured using the sequence method [33, 34].
Non-invasive hemodynamic measurements included beat-to-beat systolic (sBP) and diastolic (dBP) blood pressure; these were validated against periodic measurements obtained every 60 min by oscillometric recording from the upper contralateral arm. Transthoracic impedance measurements were used to estimate cardiac stroke volume index (SVI), cardiac index (CI), and systemic vascular resistance (SVR), from which the total peripheral resistance index (TPRI) was calculated (mean blood pressure divided by CI). Bioimpedance-based measurements have been previously validated against invasive hemodynamic monitoring [2729]. All signals were simultaneously acquired and displayed in real time on a personal computer running DOMINO 2.9.0 software.
Respiration was monitored using a polysomnograph device (SOMNO HD™, SOMNOmedics GmbH, Randersacker, Germany), recording airflow through a nasal cannula thermistor (TerniMed, Bielefeld, Germany), thoracic and abdominal efforts through inductance belts, and an electroencephalogram (EEG) through a 12-lead EEG system, according to recommendations of the American Academy of Sleep Medicine (AASM) [35]. Sleep stages and respiratory events were also analyzed based on AASM recommendations [35].

Selection of central apnea and normal breathing segments

Participants were connected to non-invasive hemodynamic and respiratory monitoring systems and spent the following night in the academic sleep laboratory at Münster University Hospital. Figure 1a shows a patient in a sleep lab bed being connected to the non-invasive hemodynamic and respiratory monitoring systems. Sleep stages, respiration, hemodynamic, and autonomic nervous system parameters were evaluated online and on a “beat-to-beat” basis by NH and JS.
From each subject, three fragments were chosen by visual identification, each of them 10 min in duration. The first segment was characterized by normal breathing (NB) during the awake state, the second by NB during stable N2 sleep, and the third segment by CA during stable N2 sleep. With regard to HRV analysis, all segments selected were required to show sinus rhythm with fewer than 5% ectopic beats. SVB and BRS were computed in the same time windows of 10 min.
Normal breathing was defined as the absence of apneas, hypopneas, and respiratory event-related arousals from sleep (RERA). Variability of peripheral oxygen saturation (SpO2) was required to be < 2% over > 80% of the time during the 10-min segment. Selection of 10-min segments of N2 sleep with CA was based on the presence of CAs for ≥ 20% of the time. A CA was scored when flow amplitude was decreased by ≥ 90% for ≥ 10 s in the absence of any ventilatory effort. CA were visually identified and manually selected for in-depth analysis of associated changes in hemodynamic measures.
Average values for HR, sBP, and dBP were calculated during the CA and compared with the 15 s preceding the apnea. The 15 s preceding the apnea had to be characterized by NB, as defined above. Representative tracings of a CA breathing pattern and NB during stable N2 sleep are shown Fig. 1b and c.

Statistical analyses

All analyses were performed using the Sigma Plot® software (Version 13.0, Systat Software GmbH, Erkrath, Germany). Assuming a two-sided significance level of 0.05 (alpha) and 80% power (beta), a sample size of ten patients per group was calculated to allow detection of a 20% change in the LF component of dBPV. Values for the mean and standard deviation of LF component of dBPV during 10 min of N2 sleep were obtained from previous preliminary data [36, 37]. Furthermore, it was also known from our previous measurements that intra-individual variation in the LF component of dBPV during N2 sleep is ~ 5% at most [37]. Results were expressed as mean and standard deviation for continuous variables with normal distribution, and median and interquartile range for continuous variables with a skewed distribution. Categorical variables were expressed as percentages, unless otherwise specified. Differences between groups (HF and ICA patients) were analyzed using the unpaired t test or the Mann-Whitney rank sum test, when appropriate, while differences in categorical data were compared using the χ2 test. Measures reflecting SVB and hemodynamic parameters were compared with baseline values using a paired t test or Wilcoxon rank sum test, as appropriate. For all tests, a two-tailed p value ≤ 0.05 was considered statistically significant.

Results

Subjects

A total of 20 subjects were enrolled, 10 with CA due to HF and 10 with ICA (Table 1). Patients with HF had lower LVEF and higher plasma levels of NT-proBNP, consistent with the inclusion criteria. Etiology of HF was ischemic in 6/10 patients. There were no significant differences in age, gender, body mass index, and basic lung function or basic objective sleep data of the study night (Supplemental Table S2) between the two patient groups. The majority of patients with systolic HF were taking a β-blocker and angiotensin-converting enzyme (ACE) inhibitor, as recommended by current guidelines [10]. Etiology of HF was ischemic in 6/10 patients. Among patients with ICA, 4/10 were taking a β-blocker and/or ACE inhibitor for pharmacological treatment of arterial hypertension. Supplemental Table 1 reports the overall sleep-related characteristics of the study night.
Table 1
Demographic and clinical characteristics of the study population at baseline
 
CA-HF (n = 10)
ICA (n = 10)
p value
Male (n (%))
9 (90)
9 (90)
1.0
Age (years)
65.2 ± 10.8
57.7 ± 13.2
0.22
BMI (kg/m2)
29.1 ± 4.6
29.6 ± 5.3
0.84
BSA (m2)
2.0 ± 0.2
2.1 ± 0.2
0.47
NYHA class I (n (%))
0 (0)
n.a.
 
NYHA class II (n (%))
7 (70)
n.a.
 
NYHA class III (n (%))
3 (30)
n.a.
 
NYHA class IV (n (%))
0 (0)
n.a.
 
LVEF (%)
36.6 ± 11.8
63.3 ± 5.5
< 0.001
Mitral regurgitation (grade III) (n (%))
0 (0)
0 (0)
1.0
TAPSE (mm)
19.7 ± 4.5
25.1 ± 3.1
0.008
NT-proBNP (ng/L)
962.5 (799.5–1738.0)
60.5 (50.0–113.0)
< 0.001
Lung function data
  FVC (L)
3.6 ± 1.1
4.5 ± 1.4
0.16
  FVC (% predicted)
93.7 ± 24.2
104.2 ± 29.8
0.41
  FEV1 (L)
3.0 ± 0.7
3.6 ± 1.0
0.15
  FEV1 (% predicted)
96.4 ± 19.2
103.9 ± 25.1
0.48
  FEV1/VC (%)
76.0 ± 7.5
70.5 ± 7.1
0.12
  PEF (L/s)
418.9 ± 125.4
442.3 ± 159.6
0.73
Medication (n (%))
  ACEI/ARB
9 (90)
4 (40)
0.018
  β-Blocker
10 (100)
3 (30)
< 0.001
  Diuretics
8 (80)
1 (10)
< 0.001
  Aldosterone antagonists
6 (60)
-
0.002
Italicized means p < 0.05
Values are mean ± standard deviation, median (interquartile range), or number of patients (%)
ACEI, angiotensin-converting enzyme inhibitor; ARB, angiotensin receptor blocker; BMI, body mass index; BSA, body surface area; CA-HF, central apnea and heart failure patients; FEV1, forced expiratory volume; FVC, forced vital capacity; ICA, idiopathic central sleep apnea patients; LVEF, left ventricular ejection fraction; NT-proBNP, N-terminal pro brain natriuretic peptide; NYHA, New York Heart Association; PEF, peak expiratory flow; TAPSE, tricuspid annular plane systolic excursion; VC, vital capacity

Baseline sympathovagal balance and hemodynamics during free breathing

Measures reflecting SVB, BRS, and hemodynamic parameters during free breathing are detailed in Table 2. Patients with HF and CA showed a reduced BRS (slope < 15 ms/mmHg) in 40% and 50% of cases considering up events and down events, respectively. Subjects with ICA showed reduced BRS in 90% and 70% of cases again considering up events and down events, respectively. [38]. HF patients had lower blood pressure, SVI, and CI compared with ICA patients (all p < 0.01). There were no significant differences in HR or total peripheral resistance index (TPRI) between HF and ICA patients (all p > 0.05).
Table 2
Sympathovagal balance and hemodynamics in patients with heart failure-related central apnea and patients with idiopathic central apnea during free breathing in the awake state
 
CA-HF
ICA
p value
Sympathovagal balance parameters
  HF nuRRI (%)
71.2 ± 15.0
39.9 ± 29.5
0.01
  LF nuRRI (%)
28.8 ± 15.0
60.1 ± 29.5
0.01
  LF/HF RRI
0.5 ± 0.4
2.8 ± 2.3
0.006
  HF nudBPV (%)
30.1 ± 12.0
22.5 ± 20.6
0.34
  LF nudBPV (%)
30.5 ± 8.6
37.6 ± 14.0
0.20
  LF/HF dBPV
1.1 ± 0.4
4.1 ± 5.0
0.07
  VLF nudBP (%)
39.4 ± 18.2
39.9 ± 17.2
0.95
BRS slope
  Up events (n)
52
91
0.16
    Up events (ms/mmHg)
33.9 ± 28.6
11.5 ± 9.9
0.07
    BR slope < 15 ms/mmHg (n (%))
40
90
0.04
  Down events (n)
58
105
0.04
    Down events (ms/mmHg)
36.1 ± 30.3
14.1 ± 11.0
0.07
    % with BRS slope < 15 ms/mmHg (n (%))
50
70
0.096
Hemodynamic parameters
  Heart rate (min−1)
68.5 ± 8.3
62.4 ± 8.8
0.15
  Systolic BP (mmHg)
98.3 ± 25.1
104.9 ± 11.5
0.51
  Diastolic BP (mmHg)
59.8 ± 23.5
63.3 ± 13.0
0.71
  Stroke volume index (mL/m2)
26.3 ± 5.4
35.0 ± 4.9
0.003
  Cardiac index (L/min/m2)
1.8 ± 0.4
2.2 ± 0.4
0.05
  TPRI (dyne s m2 cm−5)
3608.9 ± 1482.3
2997.9 ± 511.7
0.28
Italicized means p < 0.05
Values are mean ± standard deviation or number of patients (%)
BP, blood pressure; BRS slope, slope of baroreceptor reflex sensitivity (up events and down events); CA-HF, central apnea and heart failure patients; HF nudBPV, high frequency component of diastolic blood pressure variability normalized for total power spectra (normalized units); HF nuRRI, high frequency component of heart rate variability normalized for total power spectra (normalized units); ICA, idiopathic central sleep apnea patients; LF nudBPV, low frequency component of diastolic blood pressure variability normalized for total power spectra (normalized units); LF/HF dBPV, relative ratio of low frequency and high frequency component of diastolic blood pressure variability; LF nuRRI, low frequency component of heart rate variability normalized for total power spectra (normalized units); LF/HF RRI, relative ratio of low frequency and high frequency component of heart rate variability; TPRI, total peripheral resistance index

Impact of CA on sympathovagal balance and hemodynamics in patients with HF

In HF patients, transition from the awake state to N2 sleep under NB was not associated with statistically significant changes in any SVB parameters, but there were significant reductions from baseline to sleep in hemodynamic variables with a reduction in heart rate (by ~ 8%, p = 0.009), SVI (by ~ 7%, p = 0.02), and CI (by ~ 15%, p = 0.002) (Table 3).
Table 3
Effect of central apnea on sympathovagal balance and hemodynamic parameters in patients with heart failure
 
Baseline awake (NB)
N2 sleep (NB)
p value*
N2 sleep (CA)
p value**
Sympathovagal balance parameters
  HF nuRRI (%)
71.2 ± 15.0
65.3 ± 19.0
0.50
62.3 ± 24.0
0.28
  LF nuRRI (%)
28.8 ± 15.0
34.7 ± 19.0
0.50
37.7 ± 24.0
0.28
  LF/HF RRI
0.5 ± 0.4
0.8 ± 0.8
0.35
1.1 ± 1.3
0.19
  HF nudBPV (%)
30.1 ± 12.0
28.6 ± 13.2
0.61
27.6 ± 12.9
0.65
  LF nudBPV (%)
30.5 ± 8.6
27.4 ± 10.3
0.19
27.0 ± 7.6
0.86
  LF/HF dBPV
1.1 ± 0.4
1.1 ± 0.3
0.54
1.1 ± 0.3
0.73
  VLF nudBP (%)
39.4 ± 18.2
44.0 ± 22.4
0.97
45.4 ± 18.9
0.75
BRS slope
  Up events (ms/mmHg)
33.9 ± 28.6
41.8 ± 31.4
0.66
41.7 ± 32.7
0.96
  Down events (ms/mmHg)
36.1 ± 30.3
55.8 ± 38.6
0.14
36.0 ± 31.7
0.27
Hemodynamic parameters
  Heart rate (min−1)
68.5 ± 8.3
63.0 ± 10.2
0.009
64.1 ± 8.7
0.42
  Systolic BP (mmHg)
98.3 ± 25.1
100.1 ± 23.8
0.81
94.8 ± 19.0
0.38
  Diastolic BP (mmHg)
59.8 ± 23.5
63.8 ± 23.9
0.60
58.4 ± 17.9
0.41
  Stroke volume index (mL/m2)
26.3 ± 5.4
24.3 ± 5.3
0.02
24.9 ± 5.0
0.47
  Cardiac index (L/min/m2)
1.8 ± 0.4
1.5 ± 0.3
0.002
1.6 ± 0.3
0.19
  TPRI (dyne s m2 cm−5)
3608.9 ± 1482.3
4053.4 ± 1067.7
0.13
3844.0 ± 1435.7
0.09
Italicized means p < 0.05
Values are mean ± standard deviation or number of patients (%)
*For comparison vs. baseline
**For comparison CA vs. NB
BP, blood pressure; BRS slope, slope of baroreceptor reflex sensitivity (up events and down events); CA, central apnea; HF nudBPV, high frequency component of diastolic blood pressure variability normalized for total power spectra (normalized units); HF nuRRI, high frequency component of heart rate variability normalized for total power spectra (normalized units); LF nudBPV, low frequency component of diastolic blood pressure variability normalized for total power spectra (normalized units); LF/HF dBPV, relative ratio of low frequency and high frequency component of diastolic blood pressure variability; LF nuRRI, low frequency component of heart rate variability normalized for total power spectra (normalized units); LF/HF RRI, relative ratio of low frequency and high frequency component of heart rate variability; NB, normal breathing; TPRI, total peripheral resistance index
During stable N2 sleep, the phase of unstable breathing (25 ± 5% of the 10-min time segments chosen, i.e., corresponding to mild-moderate CA severity) was not associated with any significant variations in SVB and hemodynamics compared with the phase of NB (Table 3, Fig. 2 for a summary of the effects seen in the entire group of HF patients). When comparing hyperventilation and CA within the unstable breathing phase, there was a ~ 5% decrease in HR (p < 0.001), sBP (p < 0.001), and dBP (p < 0.001) during CA (n = 1233; 23 ± 5 s in length) compared with the last 15 s of hyperventilation preceding apnea onset (Fig. 2c).

Impact of CA on sympathovagal balance and hemodynamics in patients with ICA

In ICA patients, transition from the awake state to N2 sleep was associated with a reduction in LF/HF RRI (by ~ 40%) under NB conditions (p = 0.025). HR decreased slightly (~ by 5%, p = 0.098), and the mean slope of the BRS for up events was increased slightly, but these changes did not reach statistical significance (Table 4). In addition, the TPRI was markedly increased during N2 as compared with baseline values (by ~ 30%, p = 0.018) (Table 4). During stable N2 sleep, the VLF component of dBPV significantly increased during the CA segment (25 ± 5% of the 10-min time segments chosen, again and similar to HF patients enrolled corresponding to mild-moderate CA severity) vs. NB (p = 0.001), with otherwise neutral effects on other SVB and hemodynamic parameters (Table 4). Similar to our findings in HF patients, data analysis showed that in patients with ICA, central apneas (n = 1426; 21 ± 4 s in length) were associated with a ~ 5% decrease in HR (p < 0.001), sBP (p < 0.001), and dBP (p < 0.001) compared with the last 15 s of hyperventilation preceding apnea onset (Fig. 3c).
Table 4
Effect of CA on sympathovagal balance parameters and hemodynamic parameters in ICA patients
 
Baseline awake (NB)
N2 sleep (NB)
p value*
N2 sleep (ICA)
p value**
Sympathovagal balance parameters
  HF nuRRI (%)
39.9 ± 29.5
44.2 ± 23.2
0.07
41.0 ± 23.2
0.59
  LF nuRRI (%)
60.1 ± 29.5
55.8 ± 23.2
0.07
59.0 ± 23.2
0.59
  LF/HF RRI
3.7 ± 3.3
2.2 ± 2.0
0.025
2.3 ± 1.5
0.86
  HF nudBPV (%)
22.5 ± 20.6
17.6 ± 15.0
0.51
10.8 ± 15.6
0.13
  LF nudBPV (%)
37.6 ± 14.0
43.2 ± 11.5
0.47
30.5 ± 15.3
0.002
  LF/HF dBPV
2.3(1.5–4.4)
4.1(1.5–6.0)
0.84
5.3(2.7–6.9)
0.88
  VLF nudBP (%)
39.9 ± 17.2
39.2 ± 8.7
0.98
62.6 ± 17.9
0.001
BRS slope
  Up events (ms/mmHg)
11.5 ± 9.9
16.3 ± 10.9
0.09
15.2 ± 11.0
0.77
  Down events (ms/mmHg)
14.1 ± 11.0
17.4 ± 8.7
0.62
19.9 ± 14.3
0.52
Hemodynamic parameters
  Heart rate (min−1)
62.4 ± 8.8
58.0 ± 6.1
0.10
55.9 ± 7.8
0.14
  Systolic BP (mmHg)
104.9 ± 11.5
110.4 ± 11.8
0.41
110.2 ± 5.3
0.74
  Diastolic BP (mmHg)
63.3 ± 13.0
73.2 ± 9.2
0.12
70.1 ± 8.4
0.19
  Stroke volume index (mL/m2)
35.0 ± 4.9
31.8 ± 4.2
0.29
36.8 ± 15.1
0.26
  Cardiac index (L/min/m2)
2.2 ± 0.4
1.8 ± 0.3
0.17
1.8 ± 0.3
0.87
  TPRI (dyne s m2 cm−5)
2997.9 ± 511.7
4002.5 ± 956.3
0.018
3842.2 ± 936.8
0.14
Italicized means p < 0.05
Values are mean ± standard deviation, median (interquartile range), or number of patients (%)
*For comparison vs. baseline
**For comparison CA vs. NB
BP, blood pressure; BRS slope, slope of baroreceptor reflex sensitivity (up events and down events); HF nudBPV, high frequency component of diastolic blood pressure variability normalized for total power spectra (normalized units); HF nuRRI, high frequency component of heart rate variability normalized for total power spectra (normalized units); ICA, idiopathic central apnea; LF nudBPV, low frequency component of diastolic blood pressure variability normalized for total power spectra (normalized units); LF/HF dBPV, relative ratio of low frequency and high frequency component of diastolic blood pressure variability; LF nuRRI, low frequency component of heart rate variability normalized for total power spectra (normalized units); LF/HF RRI, relative ratio of low frequency and high frequency component of heart rate variability; NB, normal breathing; TPRI, total peripheral resistance index

Discussion/conclusion

This is the first study to comparatively investigate the effects of spontaneous CA on SVB and hemodynamics in both HF and ICA patients at night, after adjusting for sleep stage. The most important finding of this study is that the impact of CA on SVB seems to be different depending on the underlying hemodynamic conditions or the presence of HF. In patients with HF, CA had almost neutral effects on SVB compared with NB during N2 sleep. In contrast, in subjects with ICA, CAs were associated with an increase in the VLF component of dBPV. While some oscillatory variations of dBP can be detected within a single cycle of hyperventilation and apnea (either CA in the context of HF, or ICA), when comparing the phases of ventilatory instability to those of NB, an overall neutral hemodynamic effect was observed in both HF patients and heart-healthy subjects.
Transition from wakefulness to sleep was characterized by a decrease in HR and by a shift in HRV (but not BPV) towards vagal predominance over sympathetic control (increased high frequency component, decreased LF, and decreased HF/LF) in heart-healthy subjects, as expected and consistent with a previous study [39]. The same transition was accompanied by a decrease in only HR in patients with HF, while no changes were observed in HRV and BPV, a finding which may be in line with an overall decreased circadian modulation of different physiological signals in HF [3, 6, 9].
It has been hypothesized that CA further increases sympathetic drive in the setting of HF, thus exerting harmful effects if left untreated [4, 40]. Nonetheless, a previous study which applied both polysomnography and right heart catheterization showed that cardiac norepinephrine spillover was mainly related to HF severity rather than to CA severity (based on the AHI) [41]. Moreover, the SERVE-HF trial surprisingly showed that effective suppression of CA during treatment with adaptive servo-ventilation might result in increased mortality [1921]. This finding gave rise to the hypothesis that CSR may exert compensatory rather than detrimental effects in patients with HF [20, 21]. In particular, Naughton suggested that the hyperventilation phase may enhance stroke volume and attenuate excessive SNA with “physiological features more likely to be compensatory and beneficial than injurious in HF” [22].
We tested this hypothesis by measuring the immediate effects of a CA breathing pattern on SVB and hemodynamics during N2 sleep in HF patients. At least in this cohort of HF patients with only mild to moderate CA, these effects turned out to be neutral; i.e., no significant differences between relevant parameters were found when periods of CA were compared with NB. Patients with HF typically show a CSR pattern, and unstable breathing is characterized by repetitive cycles of hyperventilatory and apneic phases. A possible interpretation of this finding may therefore be that whatever the effect of CA on SVB and cardiac hemodynamics is, this is compensated and counterbalanced during the hyperventilation phase. However, in heart-healthy subjects with normal cardiac output and ICA, an increase in the VLF component was still observed when the same sleep stage was compared with NB.
Three explanations may help explain this finding. Firstly, CA characteristics are different in patients with and without HF (e. g., cycle duration in HF-related CA is 1–2 min vs. 30–45 s in ICA) [5, 24, 25]. Since hypocapnic hyperventilation is likely to attenuate excessive sympathetic drive, longer hyperventilatory phases in HF would explain why the net effect on SVB was neutral in CA due to HF but not in ICA (i.e., due to longer compensatory hyperventilation in HF). Secondly, even similar CAs may result in different autonomic response in patients with and without HF, depending on the background chemoreflex activity and its interaction with other cardiopulmonary reflexes during oscillatory breathing. In particular, HF leads to resetting of several feedback circuits due to forward (hypoperfusion of carotid bodies and kidneys) and backward failure (increased left atrial pressure). Indeed, a previous study from our group pointed towards this theory. Applying simulation of a CSR breathing pattern, it was shown that CSR results in neutral changes effects on SVB in patients with HF [36]. Thirdly, the effects of HF treatment, especially β-blockers, have been previously been shown to dump the firing of sympathetic outflow on peripheral targets (i.e., heart and vessels). Similar effects of β-blocker treatment were obvious in the current study since sympathetic drive was decreased in the awake state in HF compared with ICA patients.
In both patients with HF-related CA and those with ICA, CAs were found to be associated with slight decreases in HR, sBP, and dBP compared with pre-apneic (hyperventilation) levels. These findings support data from physiologic studies of Burnum et al. and a previous study from our group [42, 43], and are also in line with the oscillatory behavior observed in the pulmonary circulation [42]. It is currently unknown whether the net effect of alternating phases of apneas and hyperventilation will always be neutral on sympathovagal balance (as observed in our HF cohort) or also cause a drift in some physiological variables in subjects with different chemoreflex and/or plant gain, circulatory delay, or in different sleep stages. This should be addressed by future studies that should also ideally include some direct measurement of sympathetic outflow, such as cardiac sympathetic nerve recordings in animals or muscle SNA in humans.
Despite its comprehensive approach, this study has, in our opinion, five main limitations. Firstly, previous validation studies using the same non-invasive hemodynamic monitor were only performed in free breathing conditions. Therefore, the results obtained during different breathing maneuvers should be interpreted with caution, especially SVI, CI, and TPRI [2729]. Nonetheless, even invasive hemodynamic recordings would include uncertainties compared with a steady-state condition, especially when rapid transition from one phase to the other is evaluated (i.e., from hyperventilation to apnea). Secondly, SNA was not measured invasively in this study, i.e., by direct recording of MSNA. Non-invasive recording of HRV and BPV can only provide an indirect measure of SNA. Therefore, it is impossible to understand whether there is a problem of nerve firing or organ response. However, close correlation between MSNA and non-invasive SNA surrogate markers has previously been shown, both in heart-healthy volunteers [44] and HF patients [45], and performance of MSNA during sleep is technically difficult, especially for long recordings. Thirdly, rather short recordings of CA breathing patterns (i.e., 10 min in length) were chosen to assess the effects of CA on SVB. Hence, we cannot exclude that longer periods of CA would increase sympathetic drive in patients with HF. Fourthly, these results are not applicable to different sleep stages and the CA alteration of sleep architecture may also be a cause of HRV/BPV changes during the whole night, as previously described [46]. Finally, our findings may be different in larger populations with more severe CA, different LVEF, background hemodynamics, and feedback set point ideally also measuring MSNA and corrrelating it with the extent of hypocapnia in HF [47]. Hence, neutral findings in the present study must be interpreted again with caution considering the small sample size and pre-selection for the presence of CAs that were “only” present in 20–30% of the 10 minutes of the 10-min time segments evaluated. Taking into account the statistical assumption behind the calculation of the Bonferroni post hoc corrections for multiple t tests, actual p values reported in Tables 3 and 4 should be interpreted with caution.
In conclusion, CA during sleep is associated with acute hemodynamic effects compared with hyperventilation in both patients with HF and subjects with ICA. When comparing phases of unstable ventilation to those of stable ventilation, no significant change in SVB and hemodynamics was observed in patients with systolic HF, while an increased VLF of dBPV was observed in heart-healthy subjects with ICA. The different effects of CA in different clinical conditions should be thoroughly investigated in future studies to understand if, and in which cases, they may represent a risk and consequently a promising therapeutic target.

Acknowledgments

We gratefully acknowledge Mister Christian Glatz’s and Mister Salvador Perez’s kind technical help in collecting our data, Dr. Tobias Brix’s help in editing tables and figures, and Dr. Jürgen Fortin’s help in analyzing and interpreting autonomic nervous system and hemodynamic data. English language editing assistance was provided by Nicola Ryan, an independent medical writer.

Compliance with ethical standards

Statement of ethics

The study protocol conforms to the 1975 Declaration of Helsinki and was approved by the Institution’s human research committee (Ethikkommission der Ärztekammer Westfalen-Lippe und der Medizinischen Fakultät der Westfälischen Wilhelms-Universität Münster, Az. 2017-188-f-S). All participants gave written informed consent to participate in the study.

Conflict of interest

JS is supported by the Else-Kröner-Fresenius Stiftung (Grant SP A109) for this work, by Kommission für Innovative Medizinische Forschung an der Medizinischen Fakultät Muenster (IMF Grant SP 11 18 15), by Deutsche Herzstiftung (DHS Grant S/01/19), and by the young investigator research support from Scuola Superiore Sant’Anna Pisa (Curriculum Ph.D. in Translational Medicine) and has received travel grants and lecture honoraria from Boehringer Ingelheim and Chiesi (all) outside the present work. MB is supported by Sanofi Genzyme outside this work and has received speaker honoraria and travel grants from Sanofi Genzyme and Loewenstein Medical. WR reports no conflicts of interest in context with the paper, but has received travel grants and speaking fees outside the submitted work from Loewenstein Medical, Weinmann, Philips Respironics, Novartis, Inspire, and Boehringer Ingelheim. MD reports to have received travel grants and/or speaking fees and/or fees for consulting from Actelion, Astra Zeneca, Bayer, Berlin Chemie, Boehringer, Chiesi, GSK, Hamilton, Heinen und Löwenstein, Intermune, Linde, Novartis, Pfizer, Philips Respironics, ResMed, Roche, and Weinmann outside the present work. NH, MFG, RR, IT, XJ, SK, PS, and CP declare no conflicts of interest in the context of this article.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.
Comments
Spiesshoefer et al. describe different autonomic changes during central apneas in patients with systolic heart failure compared to central apneas in patients with idiopathic central apneas. These findings are interesting and create the basis for future investigations. Further detailed description of central respiratory events and autonomic challenge testing may provide further information to assess, whether differences in central respiratory events or differences in autonomic responses to comparable central respiratory events contribute to the observation described in this manuscript. Additionally, further sleep and autonomic intervention studies are required to finally show, whether central sleep apneas in heart failure patients are really modifiable risk factors or a marker of underlying disease and autonomic imbalance.
Dominik Linz
Mainz, Germany

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

e.Dent – Das Online-Abo der Zahnmedizin

Online-Abonnement

Mit e.Dent erhalten Sie Zugang zu allen zahnmedizinischen Fortbildungen und unseren zahnmedizinischen und ausgesuchten medizinischen Zeitschriften.

e.Med Innere Medizin

Kombi-Abonnement

Mit e.Med Innere Medizin erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes Innere Medizin, den Premium-Inhalten der internistischen Fachzeitschriften, inklusive einer gedruckten internistischen Zeitschrift Ihrer Wahl.

Anhänge

Electronic supplementary material

Literatur
1.
Zurück zum Zitat Oldenburg O, Lamp B, Faber L, Teschler H, Horstkotte D, Töpfer V (2007) Sleep-disordered breathing in patients with symptomatic heart failure. A contemporary study of prevalence in and characteristics of 700 patients. Eur J Heart Fail 9:251–257PubMed Oldenburg O, Lamp B, Faber L, Teschler H, Horstkotte D, Töpfer V (2007) Sleep-disordered breathing in patients with symptomatic heart failure. A contemporary study of prevalence in and characteristics of 700 patients. Eur J Heart Fail 9:251–257PubMed
2.
Zurück zum Zitat Lanfranchi PA, Somers VK, Braghiroli A, Corra U, Eleuteri E, Giannuzzi P (2003) Central sleep apnea in left ventricular dysfunction: prevalence and implications for arrhythmic risk. Circulation 107:727–732PubMed Lanfranchi PA, Somers VK, Braghiroli A, Corra U, Eleuteri E, Giannuzzi P (2003) Central sleep apnea in left ventricular dysfunction: prevalence and implications for arrhythmic risk. Circulation 107:727–732PubMed
3.
Zurück zum Zitat Emdin M, Mirizzi G, Giannoni A, Poletti R, Iudice G, Bramanti F, Passino C (2017) Prognostic significance of central apneas throughout a 24-hour period in patients with heart failure. J Am Coll Cardiol 70:1351–1364PubMed Emdin M, Mirizzi G, Giannoni A, Poletti R, Iudice G, Bramanti F, Passino C (2017) Prognostic significance of central apneas throughout a 24-hour period in patients with heart failure. J Am Coll Cardiol 70:1351–1364PubMed
4.
Zurück zum Zitat Kara T, Narkiewicz K, Somers VK (2003) Chemoreflexes-physiology and clinical implications. Acta Physiol Scand 177:377–384PubMed Kara T, Narkiewicz K, Somers VK (2003) Chemoreflexes-physiology and clinical implications. Acta Physiol Scand 177:377–384PubMed
5.
Zurück zum Zitat Spiesshoefer J, Spieker M, Klose et al (2019) Reduction of sleep-disordered breathing following effective percutaneous mitral valve repair with the MitraClip system. Sleep Breath 23:815–824PubMed Spiesshoefer J, Spieker M, Klose et al (2019) Reduction of sleep-disordered breathing following effective percutaneous mitral valve repair with the MitraClip system. Sleep Breath 23:815–824PubMed
6.
Zurück zum Zitat Solin P, Kaye DM, Little PJ et al (2003) Impact of sleep apnea on sympathetic nervous system activity in heart failure. Chest 123:1119–1126PubMed Solin P, Kaye DM, Little PJ et al (2003) Impact of sleep apnea on sympathetic nervous system activity in heart failure. Chest 123:1119–1126PubMed
7.
Zurück zum Zitat Giannoni A, Raglianti V, Mirizzi G, Taddei C, del Franco A, Iudice G, Bramanti F, Aimo A, Pasanisi E, Emdin M, Passino C (2016) Influence of central apneas and chemo reflex activation on pulmonary artery pressure in chronic heart failure. Int J Cardiol 202:200–206PubMed Giannoni A, Raglianti V, Mirizzi G, Taddei C, del Franco A, Iudice G, Bramanti F, Aimo A, Pasanisi E, Emdin M, Passino C (2016) Influence of central apneas and chemo reflex activation on pulmonary artery pressure in chronic heart failure. Int J Cardiol 202:200–206PubMed
8.
Zurück zum Zitat Giannoni A, Raglianti V, Taddei C, Borrelli C, Chubuchny V, Vergaro G, Mirizzi G, Valleggi A, Cameli M, Pasanisi E, Emdin M, Passino C (2019) Cheyne-Stokes respiration related oscillations in cardiopulmonary hemodynamics in patients with heart failure. Int J Cardiol 289:76–82PubMed Giannoni A, Raglianti V, Taddei C, Borrelli C, Chubuchny V, Vergaro G, Mirizzi G, Valleggi A, Cameli M, Pasanisi E, Emdin M, Passino C (2019) Cheyne-Stokes respiration related oscillations in cardiopulmonary hemodynamics in patients with heart failure. Int J Cardiol 289:76–82PubMed
9.
Zurück zum Zitat Floras JS (2009) Sympathetic nervous system activation in human heart failure. Clinical implications of an updated model. J Am Coll Cardiol 54:375–385PubMed Floras JS (2009) Sympathetic nervous system activation in human heart failure. Clinical implications of an updated model. J Am Coll Cardiol 54:375–385PubMed
10.
Zurück zum Zitat Ponikowski P, Voors A, Anker S, Bueno H, Cleland JGF, Coats AJS, Falk V, González-Juanatey JR, Harjola VP, Jankowska EA, Jessup M, Linde C, Nihoyannopoulos P, Parissis JT, Pieske B, Riley JP, Rosano GMC, Ruilope LM, Ruschitzka F, Rutten FH, van der Meer P, ESC Scientific Document Group (2016) 2016 ESC guidelines for the diagnosis and treatment of acute and chronic heart failure: the Task Force for the diagnosis and treatment of acute and chronic heart failure of the European Society of Cardiology (ESC) developed with the special contribution of the Heart Failure Association (HFA) of the ESC. Eur Heart J 37:2129–2200PubMed Ponikowski P, Voors A, Anker S, Bueno H, Cleland JGF, Coats AJS, Falk V, González-Juanatey JR, Harjola VP, Jankowska EA, Jessup M, Linde C, Nihoyannopoulos P, Parissis JT, Pieske B, Riley JP, Rosano GMC, Ruilope LM, Ruschitzka F, Rutten FH, van der Meer P, ESC Scientific Document Group (2016) 2016 ESC guidelines for the diagnosis and treatment of acute and chronic heart failure: the Task Force for the diagnosis and treatment of acute and chronic heart failure of the European Society of Cardiology (ESC) developed with the special contribution of the Heart Failure Association (HFA) of the ESC. Eur Heart J 37:2129–2200PubMed
11.
Zurück zum Zitat Lorenzi-Filho G, Rankin F, Bies I, Douglas Bradley T (1999) Effects of inhaled carbon dioxide and oxygen on Cheyne-Stokes respiration in patients with heart failure. Am J Respir Crit Care Med 159:1490–1498PubMed Lorenzi-Filho G, Rankin F, Bies I, Douglas Bradley T (1999) Effects of inhaled carbon dioxide and oxygen on Cheyne-Stokes respiration in patients with heart failure. Am J Respir Crit Care Med 159:1490–1498PubMed
12.
Zurück zum Zitat Giannoni A, Baruah R, Willson K, Mebrate Y, Mayet J, Emdin M, Hughes AD, Manisty CH, Francis DP (2010) Real-time dynamic carbon dioxide administration a novel treatment strategy for stabilization of periodic breathing with potential application to central sleep apnea. J Am Coll Cardiol 56:1832–1837PubMed Giannoni A, Baruah R, Willson K, Mebrate Y, Mayet J, Emdin M, Hughes AD, Manisty CH, Francis DP (2010) Real-time dynamic carbon dioxide administration a novel treatment strategy for stabilization of periodic breathing with potential application to central sleep apnea. J Am Coll Cardiol 56:1832–1837PubMed
13.
Zurück zum Zitat Schumacher DS, Müller-Mottet S, Hasler ED et al (2014) Effect of oxygen and acetazolamide on nocturnal cardiac conduction, repolarization, and arrhythmias in precapillary pulmonary hypertension and sleep-disturbed breathing. Chest 146:1226–1236PubMed Schumacher DS, Müller-Mottet S, Hasler ED et al (2014) Effect of oxygen and acetazolamide on nocturnal cardiac conduction, repolarization, and arrhythmias in precapillary pulmonary hypertension and sleep-disturbed breathing. Chest 146:1226–1236PubMed
14.
Zurück zum Zitat Fontana M, Emdin M, Giannoni A, Iudice G, Baruah R, Passino C (2011) Effect of acetazolamide on chemosensitivity, Cheyne-Stokes respiration, and response to effort in patients with heart failure. Am J Cardiol 107:1675–1680PubMed Fontana M, Emdin M, Giannoni A, Iudice G, Baruah R, Passino C (2011) Effect of acetazolamide on chemosensitivity, Cheyne-Stokes respiration, and response to effort in patients with heart failure. Am J Cardiol 107:1675–1680PubMed
15.
Zurück zum Zitat Javaheri S (2006) Acetazolamide improves central sleep apnea in heart failure: a double-blind, prospective study. Am J Respir Crit Care Med 173:234–237PubMed Javaheri S (2006) Acetazolamide improves central sleep apnea in heart failure: a double-blind, prospective study. Am J Respir Crit Care Med 173:234–237PubMed
16.
Zurück zum Zitat Naughton MT, Benard DC, Liu PP, Rutherford R, Rankin FBT (1995) Effects of nasal CPAP on sympathetic activity in patients with heart failure and central sleep apnea. Am J Respir Crit Care Med 152:473–479PubMed Naughton MT, Benard DC, Liu PP, Rutherford R, Rankin FBT (1995) Effects of nasal CPAP on sympathetic activity in patients with heart failure and central sleep apnea. Am J Respir Crit Care Med 152:473–479PubMed
17.
Zurück zum Zitat Koyama T, Watanabe H, Tamura Y, Oguma Y, Kosaka T, Ito H (2013) Adaptive servo-ventilation therapy improves cardiac sympathetic nerve activity in patients with heart failure. Eur J Heart Fail 15:902–909PubMed Koyama T, Watanabe H, Tamura Y, Oguma Y, Kosaka T, Ito H (2013) Adaptive servo-ventilation therapy improves cardiac sympathetic nerve activity in patients with heart failure. Eur J Heart Fail 15:902–909PubMed
18.
Zurück zum Zitat Joho S, Oda Y, Ushijima R, Hirai T, Inoue H (2012) Effect of adaptive servoventilation on muscle sympathetic nerve activity in patients with chronic heart failure and central sleep apnea. J Card Fail 18:769–775PubMed Joho S, Oda Y, Ushijima R, Hirai T, Inoue H (2012) Effect of adaptive servoventilation on muscle sympathetic nerve activity in patients with chronic heart failure and central sleep apnea. J Card Fail 18:769–775PubMed
19.
Zurück zum Zitat Cowie MR, Woehrle H, Wegscheider K et al (2018) Adaptive servo-ventilation for central sleep apnoea in systolic heart failure : results of the major substudy of SERVE-HF. Eur J Heart Fail 20:536–544PubMed Cowie MR, Woehrle H, Wegscheider K et al (2018) Adaptive servo-ventilation for central sleep apnoea in systolic heart failure : results of the major substudy of SERVE-HF. Eur J Heart Fail 20:536–544PubMed
20.
Zurück zum Zitat Mehra R, Gottlieb DJ (2015) A paradigm shift in the treatment of central sleep apnea in heart failure. Chest 148:848–851PubMed Mehra R, Gottlieb DJ (2015) A paradigm shift in the treatment of central sleep apnea in heart failure. Chest 148:848–851PubMed
21.
Zurück zum Zitat Linz D, Fox H, Bitter T et al (2016) Impact of SERVE-HF on management of sleep disordered breathing in heart failure: a call for further studies. Clin Res Cardiol 105:563–570PubMed Linz D, Fox H, Bitter T et al (2016) Impact of SERVE-HF on management of sleep disordered breathing in heart failure: a call for further studies. Clin Res Cardiol 105:563–570PubMed
22.
Zurück zum Zitat Naughton MT (2012) Cheyne–Stokes respiration: friend or foe? Thorax 67:357–360PubMed Naughton MT (2012) Cheyne–Stokes respiration: friend or foe? Thorax 67:357–360PubMed
23.
Zurück zum Zitat Oldenburg O, Spießhöfer J, Fox H et al (2015) Cheyne-Stokes respiration in heart failure: friend or foe? Hemodynamic effects of hyperventilation in heart failure patients and healthy volunteers. Clin Res Cardiol 104:328–333PubMed Oldenburg O, Spießhöfer J, Fox H et al (2015) Cheyne-Stokes respiration in heart failure: friend or foe? Hemodynamic effects of hyperventilation in heart failure patients and healthy volunteers. Clin Res Cardiol 104:328–333PubMed
25.
Zurück zum Zitat Solin P, Roebuck T, Johns DP, Walters EH, Naughton MT (2000) Peripheral and central ventilatory responses in central sleep apnea with and without congestive heart failure. Am J Respir Crit Care Med 162:2194–2200PubMed Solin P, Roebuck T, Johns DP, Walters EH, Naughton MT (2000) Peripheral and central ventilatory responses in central sleep apnea with and without congestive heart failure. Am J Respir Crit Care Med 162:2194–2200PubMed
26.
Zurück zum Zitat Baumgartner H, Falk V, Bax JJ et al (2017) 2017 ESC/EACTS guidelines for the management of valvular heart disease. Eur Heart J 38:2739–2786PubMed Baumgartner H, Falk V, Bax JJ et al (2017) 2017 ESC/EACTS guidelines for the management of valvular heart disease. Eur Heart J 38:2739–2786PubMed
27.
Zurück zum Zitat Fortin J, Habenbacher W, Heller A, Hacker A, Grüllenberger R, Innerhofer J, Passath H, Wagner Ch, Haitchi G, Flotzinger D, Pacher R, Wach P (2006) Non-invasive beat-to-beat cardiac output monitoring by an improved method of transthoracic bioimpedance measurement. Comput Biol Med 36:1185–1203PubMed Fortin J, Habenbacher W, Heller A, Hacker A, Grüllenberger R, Innerhofer J, Passath H, Wagner Ch, Haitchi G, Flotzinger D, Pacher R, Wach P (2006) Non-invasive beat-to-beat cardiac output monitoring by an improved method of transthoracic bioimpedance measurement. Comput Biol Med 36:1185–1203PubMed
28.
Zurück zum Zitat Fortin J, Marte W, Grüllenberger R, Hacker A, Habenbacher W, Heller A, Wagner Ch, Wach P, Skrabal F (2006) Continuous non-invasive blood pressure monitoring using concentrically interlocking control loops. Comput Biol Med 36:941–957PubMed Fortin J, Marte W, Grüllenberger R, Hacker A, Habenbacher W, Heller A, Wagner Ch, Wach P, Skrabal F (2006) Continuous non-invasive blood pressure monitoring using concentrically interlocking control loops. Comput Biol Med 36:941–957PubMed
29.
Zurück zum Zitat Spießhöfer J, Fox H, Lehmann R et al (2016) Heterogenous haemodynamic effects of adaptive servoventilation therapy in sleeping patients with heart failure and Cheyne–Stokes respiration compared to healthy volunteers. Heart Vessel 31:1117–1130 Spießhöfer J, Fox H, Lehmann R et al (2016) Heterogenous haemodynamic effects of adaptive servoventilation therapy in sleeping patients with heart failure and Cheyne–Stokes respiration compared to healthy volunteers. Heart Vessel 31:1117–1130
30.
Zurück zum Zitat Gratze G, Fortin J, Holler A, Grasenick K, Pfurtscheller G, Wach P, Schönegger J, Kotanko P, Skrabal F (1998) A software package for non-invasive, real-time beat-to-beat monitoring of stroke volume, blood pressure, total peripheral resistance and for assessment of autonomic function. Comput Biol Med 28:121–142PubMed Gratze G, Fortin J, Holler A, Grasenick K, Pfurtscheller G, Wach P, Schönegger J, Kotanko P, Skrabal F (1998) A software package for non-invasive, real-time beat-to-beat monitoring of stroke volume, blood pressure, total peripheral resistance and for assessment of autonomic function. Comput Biol Med 28:121–142PubMed
31.
Zurück zum Zitat Nakata A, Takata S, Yuasa T et al (1998) Spectral analysis of heart rate, arterial pressure, and muscle sympathetic nerve activity in normal humans. Power 274:1211–1217 Nakata A, Takata S, Yuasa T et al (1998) Spectral analysis of heart rate, arterial pressure, and muscle sympathetic nerve activity in normal humans. Power 274:1211–1217
32.
Zurück zum Zitat Guidelines (1996) Guidelines Heart rate variability. Eur Heart J 17:354–381 Guidelines (1996) Guidelines Heart rate variability. Eur Heart J 17:354–381
33.
Zurück zum Zitat Carrasco-Sosa S, Gaitán-González MJ, González-Camarena R, Yáñez-Suárez O (2005) Baroreflex sensitivity assessment and heart rate variability: relation to maneuver and technique. Eur J Appl Physiol 95:265–275PubMed Carrasco-Sosa S, Gaitán-González MJ, González-Camarena R, Yáñez-Suárez O (2005) Baroreflex sensitivity assessment and heart rate variability: relation to maneuver and technique. Eur J Appl Physiol 95:265–275PubMed
34.
Zurück zum Zitat Davies LC, Francis D, Jurák P et al (1999) Reproducibility of methods for assessing baroreflex sensitivity in normal controls and in patients with chronic heart failure. Clin Sci (Lond) 97:515–522 Davies LC, Francis D, Jurák P et al (1999) Reproducibility of methods for assessing baroreflex sensitivity in normal controls and in patients with chronic heart failure. Clin Sci (Lond) 97:515–522
35.
Zurück zum Zitat Berry RB, Gamaldo CE, Harding SM, Lloyd RM, Marcus CL, Vaughn BV, for the American Academy of Sleep Medicine BR (2012) The AASM manual for the scoring of sleep and associated events. J Clin Sleep Med 8:597–619PubMedPubMedCentral Berry RB, Gamaldo CE, Harding SM, Lloyd RM, Marcus CL, Vaughn BV, for the American Academy of Sleep Medicine BR (2012) The AASM manual for the scoring of sleep and associated events. J Clin Sleep Med 8:597–619PubMedPubMedCentral
36.
Zurück zum Zitat Spiesshoefer J, Becker S, Tuleta I, Mohr M, Diller GP, Emdin M, Florian AR, Yilmaz A, Boentert M, Giannoni A (2019) Impact of simulated hyperventilation and periodic breathing on sympatho-vagal balance and hemodynamics in patients with and without heart failure. Respiration 98:482–494PubMed Spiesshoefer J, Becker S, Tuleta I, Mohr M, Diller GP, Emdin M, Florian AR, Yilmaz A, Boentert M, Giannoni A (2019) Impact of simulated hyperventilation and periodic breathing on sympatho-vagal balance and hemodynamics in patients with and without heart failure. Respiration 98:482–494PubMed
38.
Zurück zum Zitat Karemaker JM (2017) An introduction into autonomic nervous function. Physiol Meas 38:89–118 Karemaker JM (2017) An introduction into autonomic nervous function. Physiol Meas 38:89–118
39.
Zurück zum Zitat Penzel T, Wessel N, Riedl M et al (2007) Cardiovascular and respiratory dynamics during normal and pathological sleep. Chaos 17:015116PubMed Penzel T, Wessel N, Riedl M et al (2007) Cardiovascular and respiratory dynamics during normal and pathological sleep. Chaos 17:015116PubMed
40.
Zurück zum Zitat Poletti R, Passino C, Giannoni A, Zyw L, Prontera C, Bramanti F, Clerico A, Piepoli M, Emdin M (2009) Risk factors and prognostic value of daytime Cheyne–Stokes respiration in chronic heart failure patients. Int J Cardiol 137:47–53PubMed Poletti R, Passino C, Giannoni A, Zyw L, Prontera C, Bramanti F, Clerico A, Piepoli M, Emdin M (2009) Risk factors and prognostic value of daytime Cheyne–Stokes respiration in chronic heart failure patients. Int J Cardiol 137:47–53PubMed
41.
Zurück zum Zitat Mansfield D, Kaye DM, La Rocca HB et al (2003) Raised sympathetic nerve activity in heart failure and central sleep apnea is due to heart failure severity. Circulation 107:1396–1400PubMed Mansfield D, Kaye DM, La Rocca HB et al (2003) Raised sympathetic nerve activity in heart failure and central sleep apnea is due to heart failure severity. Circulation 107:1396–1400PubMed
42.
Zurück zum Zitat Maze SS, Kotler MN, Parry WR (1989) Doppler evaluation of changing cardiac dynamics during Cheyne-Stokes respiration. Chest 95:525–529PubMed Maze SS, Kotler MN, Parry WR (1989) Doppler evaluation of changing cardiac dynamics during Cheyne-Stokes respiration. Chest 95:525–529PubMed
43.
Zurück zum Zitat Xie A, Wong B, Phillipson EA et al (1994) Interaction of hyperventilation and arousal in the pathogenesis of idiopathic central sleep apnea. Am J Respir Crit Care Med 150:489–495PubMed Xie A, Wong B, Phillipson EA et al (1994) Interaction of hyperventilation and arousal in the pathogenesis of idiopathic central sleep apnea. Am J Respir Crit Care Med 150:489–495PubMed
44.
Zurück zum Zitat Van de Borne P, Montano N, Zimmerman B et al (1997) Relationship between repeated measures of hemodynamics, muscle sympathetic nerve activity, and their spectral oscillations. Circulation 96:4326–4332PubMed Van de Borne P, Montano N, Zimmerman B et al (1997) Relationship between repeated measures of hemodynamics, muscle sympathetic nerve activity, and their spectral oscillations. Circulation 96:4326–4332PubMed
45.
Zurück zum Zitat Kienzle MG, Ferguson DW, Birkett CL et al (1992) Clinical, hemodynamic and sympathetic neural correlates of heart rate variability in congestive heart failure. Am J Cardiol 69:761–767PubMed Kienzle MG, Ferguson DW, Birkett CL et al (1992) Clinical, hemodynamic and sympathetic neural correlates of heart rate variability in congestive heart failure. Am J Cardiol 69:761–767PubMed
46.
Zurück zum Zitat Türoff A, Thiem U, Fox H, Spießhöfer J, Bitter T, Tamisier R, Punjabi NM, Horstkotte D, Oldenburg O (2017) Sleep duration and quality in heart failure patients. Sleep Breath 21:919–927PubMed Türoff A, Thiem U, Fox H, Spießhöfer J, Bitter T, Tamisier R, Punjabi NM, Horstkotte D, Oldenburg O (2017) Sleep duration and quality in heart failure patients. Sleep Breath 21:919–927PubMed
47.
Zurück zum Zitat Burnum JF, Hickam JB, McIntosh HD (1954) The effect of hypocapnia on arterial blood pressure. Circulation 9:89–95PubMed Burnum JF, Hickam JB, McIntosh HD (1954) The effect of hypocapnia on arterial blood pressure. Circulation 9:89–95PubMed
Metadaten
Titel
Effects of central apneas on sympathovagal balance and hemodynamics at night: impact of underlying systolic heart failure
verfasst von
Jens Spiesshoefer
Nora Hegerfeld
Malte Frank Gerdes
Sören Klemm
Martha Gorbachevski
Robert Radke
Izabela Tuleta
Claudio Passino
Xiaoyi Jiang
Paolo Sciarrone
Winfried Randerath
Michael Dreher
Matthias Boentert
Alberto Giannoni
Publikationsdatum
22.07.2020
Verlag
Springer International Publishing
Erschienen in
Sleep and Breathing / Ausgabe 2/2021
Print ISSN: 1520-9512
Elektronische ISSN: 1522-1709
DOI
https://doi.org/10.1007/s11325-020-02144-8

Weitere Artikel der Ausgabe 2/2021

Sleep and Breathing 2/2021 Zur Ausgabe

Leitlinien kompakt für die Innere Medizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Update Innere Medizin

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.